Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-23T06:39:18.468Z Has data issue: false hasContentIssue false

Part III - Neuroimaging Methods and Findings

Published online by Cambridge University Press:  11 June 2021

Aron K. Barbey
Affiliation:
University of Illinois, Urbana-Champaign
Sherif Karama
Affiliation:
McGill University, Montréal
Richard J. Haier
Affiliation:
University of California, Irvine
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2021

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

References

Allin, M. P. G., Kontis, D., Walshe, M., Wyatt, J., Barker, G. J., Kanaan, R. A. A., … Nosarti, C. (2011). White matter and cognition in adults who were born preterm. PLoS One, 6(10), e24525.CrossRefGoogle ScholarPubMed
Atkinson, D. S., Abou-Khalil, B., Charles, P. D., & Welch, L. (1996). Midsagittal corpus callosum area, intelligence and language in epilepsy. Journal of Neuroimaging, 6(4), 235239.CrossRefGoogle ScholarPubMed
Barbey, A. K. (2018). Network neuroscience theory of human intelligence. Trends in Cognitive Sciences, 22(1), 820.CrossRefGoogle ScholarPubMed
Basser, P. J., & Pierpaoli, C. (1996). Microstructural and physiological features of tissues elucidated by quantitative-diffusion-tensor MRI. Journal of Magnetic Resonance, Series B, 111(3), 209219.Google Scholar
Beaulieu, C. (2002). The basis of anisotropic water diffusion in the nervous system – A technical review. NMR in Biomedicine, 15(7–8), 435455.Google Scholar
Behrens, T. E., Berg, H. J., Jbabdi, S., Rushworth, M. F. S., & Woolrich, M. W. (2007). Probabilistic diffusion tractography with multiple fibre orientations: What can we gain? NeuroImage, 34(1), 144155.CrossRefGoogle ScholarPubMed
Behrens, T. E., Woolrich, M. W., Jenkinson, M., Johansen-Berg, H., Nunes, R. G., Clare, S., … Smith, S. M. (2003). Characterization and propagation of uncertainty in diffusion-weighted MR imaging. Magnetic Resonance in Medicine, 50(5), 10771088.Google Scholar
Booth, T., Bastin, M. E., Penke, L., Maniega, S. M., Murray, C., Royle, N. A., … Hernández, M. (2013). Brain white matter tract integrity and cognitive abilities in community-dwelling older people: The Lothian Birth Cohort, 1936. Neuropsychology, 27(5), 595607.CrossRefGoogle ScholarPubMed
Campbell, J. S. W., & Pike, G. B. (2014). Potential and limitations of diffusion MRI tractography for the study of language. Brain and Language, 131, 6573.CrossRefGoogle Scholar
Catani, M., & Thiebaut de Schotten, M. (2008). A diffusion tensor imaging tractography atlas for virtual in vivo dissections. Cortex, 44(8), 11051132.Google Scholar
Chiang, M. C., Barysheva, M., Shattuck, D. W., Lee, A. D., Madsen, S. K., Avedissian, C., … Thompson, P. M. (2009). Genetics of brain fiber architecture and intellectual performance. Journal of Neuroscience, 29(7), 22122224.Google Scholar
Cremers, L. G. M., de Groot, M., Hofman, A., Krestin, G. P., van der Lugt, A., Niessen, W. J., … Ikram, M. A. (2016). Altered tract-specific white matter microstructure is related to poorer cognitive performance: The Rotterdam Study. Neurobiology of Aging, 39, 108117.CrossRefGoogle ScholarPubMed
Deary, I. J., Bastin, M. E., Pattie, A., Clayden, J. D., Whalley, L. J., Starr, J. M., & Wardlaw, J. M. (2006). White matter integrity and cognition in childhood and old age. Neurology, 66(4), 505512.Google Scholar
Deary, I. J., Penke, L., & Johnson, W. (2010). The neuroscience of human intelligence differences. Nature Reviews Neuroscience, 11(3), 201211.CrossRefGoogle ScholarPubMed
Dunst, B., Benedek, M., Koschutnig, K., Jauk, E., & Neubauer, A. C. (2014). Sex differences in the IQ-white matter microstructure relationship: A DTI study. Brain and Cognition, 91, 7178.CrossRefGoogle ScholarPubMed
Ferrer, E., Whitaker, K. J., Steele, J. S., Green, C. T., Wendelken, C., & Bunge, S. A. (2013). White matter maturation supports the development of reasoning ability through its influence on processing speed. Developmental Science, 16(6), 941951.CrossRefGoogle ScholarPubMed
Filley, C. (2012). The behavioral neurology of white matter. New York: Oxford University Press.CrossRefGoogle ScholarPubMed
Fischer, F. U., Wolf, D., Scheurich, A., & Fellgiebel, A. (2014). Association of structural global brain network properties with intelligence in normal aging. PLoS One, 9(1), e86258.Google Scholar
Galton, F. (1888). Head growth in students at the University of Cambridge. Nature, 38(996), 1415.Google Scholar
Genc, E., Fraenz, C., Schlüter, C., Friedrich, P., Hossiep, R., Voelkle, M. C., … Jung, R. E. (2018). Diffusion markers of dendritic density and arborization in gray matter predict differences in intelligence. Nature Communications, 9(1), 1905.Google Scholar
Genc, E., Fraenz, C., Schlüter, C., Friedrich, P., Voelkle, M. C., Hossiep, R., & Güntürkün, O. (2019). The neural architecture of general knowledge. European Journal of Personality, 33(5), 589605.CrossRefGoogle Scholar
Goriounova, N. A., Heyer, D. B., Wilbers, R., Verhoog, M. B., Giugliano, M., Verbist, C., … Verberne, M. (2018). Large and fast human pyramidal neurons associate with intelligence. eLife, 7(1), e41714.Google Scholar
Goriounova, N. A., & Mansvelder, H. D. (2019). Genes, cells and brain areas of intelligence. Frontiers in Human Neuroscience, 13, 14.CrossRefGoogle ScholarPubMed
Haász, J., Westlye, E. T., Fjær, S., Espeseth, T., Lundervold, A., & Lundervold, A. J. (2013). General fluid-type intelligence is related to indices of white matter structure in middle-aged and old adults. NeuroImage, 83, 372383.CrossRefGoogle ScholarPubMed
Hulshoff-Pol, H. E., Schnack, H. G., Posthuma, D., Mandl, R. C. W., Baare, W. F., van Oel, C., … Kahn, R. S. (2006). Genetic contributions to human brain morphology and intelligence. Journal of Neuroscience, 26(40), 1023510242.CrossRefGoogle ScholarPubMed
Jung, R. E., & Haier, R. J. (2007). The Parieto-Frontal Integration Theory (P-FIT) of intelligence: Converging neuroimaging evidence. Behavioral and Brain Sciences, 30(2), 135154.Google Scholar
Kievit, R. A., Davis, S. W., Mitchell, D. J., Taylor, J. R., Duncan, J., Tyler, L. K., … Cusack, R. (2014). Distinct aspects of frontal lobe structure mediate age-related differences in fluid intelligence and multitasking. Nature Communications, 5, 5658.Google Scholar
Kim, D. J., Davis, E. P., Sandman, C. A., Sporns, O., O’Donnell, B. F., Buss, C., & Hetrick, W. P. (2016). Children’s intellectual ability is associated with structural network integrity. NeuroImage, 124, 550556.CrossRefGoogle ScholarPubMed
Koenis, M. M. G., Brouwer, R. M., Swagerman, S. C., van Soelen, I. L. C., Boomsma, D. I., & Hulshoff Pol, H. E. (2018). Association between structural brain network efficiency and intelligence increases during adolescence. Human Brain Mapping, 39(2), 822836.CrossRefGoogle ScholarPubMed
Kontis, D., Catani, M., Cuddy, M., Walshe, M., Nosarti, C., Jones, D., … Allin, M. (2009). Diffusion tensor MRI of the corpus callosum and cognitive function in adults born preterm. Neuroreport, 20(4), 424428.Google Scholar
Kuznetsova, K. A., Maniega, S. M., Ritchie, S. J., Cox, S. R., Storkey, A. J., Starr, J. M., … Bastin, M. E. (2016). Brain white matter structure and information processing speed in healthy older age. Brain Structure and Function, 221(6), 32233235.Google Scholar
Le Bihan, D. (2003). Looking into the functional architecture of the brain with diffusion MRI. Nature Reviews Neuroscience, 4(6), 469480.Google Scholar
Li, Y., Liu, Y., Li, J., Qin, W., Li, K., Yu, C., & Jiang, T. (2009). Brain anatomical network and intelligence. PLoS Computational Biology, 5(5), e1000395.CrossRefGoogle ScholarPubMed
Luders, E., Narr, K. L., Bilder, R. M., Thompson, P. M., Szeszko, P. R., Hamilton, L., & Toga, A. W. (2007). Positive correlations between corpus callosum thickness and intelligence. NeuroImage, 37(4), 14571464.Google Scholar
Ma, J., Kang, H. J., Kim, J. Y., Jeong, H. S., Im, J. J., Namgung, E., … Oh, J. K. (2017). Network attributes underlying intellectual giftedness in the developing brain. Scientific Reports, 7(1), 11321.CrossRefGoogle ScholarPubMed
MacKay, A. L., & Laule, C. (2016). Magnetic resonance of myelin water: An in vivo marker for myelin. Brain Plasticity, 2(1), 7191.Google Scholar
Malpas, C. B., Genc, S., Saling, M. M., Velakoulis, D., Desmond, P. M., & O’Brien, T. J. (2016). MRI correlates of general intelligence in neurotypical adults. Journal of Clinical Neuroscience, 24, 128134.CrossRefGoogle ScholarPubMed
McDaniel, M. A. (2005). Big-brained people are smarter: A meta-analysis of the relationship between in vivo brain volume and intelligence. Intelligence, 33(4), 337346.Google Scholar
Mori, S. (2007). Introduction to diffusion tensor imaging. Oxford: Elsevier.Google Scholar
Morris, D. M., Embleton, K. V., & Parker, G. J. M. (2008). Probabilistic fibre tracking: Differentiation of connections from chance events. NeuroImage, 42(4), 13291339.Google Scholar
Muetzel, R. L., Mous, S. E., van der Ende, J., Blanken, L. M. E., van der Lugt, A., Jaddoe, V. W. V., … White, T. (2015). White matter integrity and cognitive performance in school-age children: A population-based neuroimaging study. NeuroImage, 119, 119128.Google Scholar
Narr, K. L., Woods, R. P., Thompson, P. M., Szeszko, P., Robinson, D., Dimtcheva, T., … Bilder, R. M. (2007). Relationships between IQ and regional cortical gray matter thickness in healthy adults. Cerebral Cortex, 17(9), 21632171.Google Scholar
Neubauer, A., & Fink, A. (2009). Intelligence and neural efficiency. Neuroscience and Biobehavioral Reviews, 33(7), 10041023.Google Scholar
Nusbaum, F., Hannoun, S., Kocevar, G., Stamile, C., Fourneret, P., Revol, O., & Sappey-Marinier, D. (2017). Hemispheric differences in white matter microstructure between two profiles of children with high intelligence quotient vs. controls: A tract-based spatial statistics study. Frontiers in Neuroscience, 11, 173.Google Scholar
Ocklenburg, S., Anderson, C., Gerding, W. M., Fraenz, C., Schluter, C., Friedrich, P., … Genc, E. (2018). Myelin water fraction imaging reveals hemispheric asymmetries in human white matter that are associated with genetic variation in PLP1. Molecular Neurobiology, 56(6), 39994012.Google Scholar
Pakkenberg, B., & Gundersen, H. J. G. (1997). Neocortical neuron number in humans: Effect of sex and age. Journal of Comparative Neurology, 384(2), 312320.Google Scholar
Penke, L., Maniega, S. M., Bastin, M. E., Hernandez, M. C. V., Murray, C., Royle, N. A., … Deary, I. J. (2012). Brain white matter tract integrity as a neural foundation for general intelligence. Molecular Psychiatry, 17(10), 10261030.Google Scholar
Penke, L., Maniega, S. M., Murray, C., Gow, A. J., Hernandez, M. C., Clayden, J. D., … Deary, I. J. (2010). A general factor of brain white matter integrity predicts information processing speed in healthy older people. Journal of Neuroscience, 30(22), 75697574.Google Scholar
Pietschnig, J., Penke, L., Wicherts, J. M., Zeiler, M., & Voracek, M. (2015). Meta-analysis of associations between human brain volume and intelligence differences: How strong are they and what do they mean? Neuroscience and Biobehavioral Reviews, 57, 411432.CrossRefGoogle ScholarPubMed
Pineda-Pardo, J. A., Martínez, K., Román, F. J., & Colom, R. (2016). Structural efficiency within a parieto-frontal network and cognitive differences. Intelligence, 54, 105116.Google Scholar
Ryman, S. G., Yeo, R. A., Witkiewitz, K., Vakhtin, A. A., van den Heuvel, M., de Reus, M., … Jung, R. E. (2016). Fronto-Parietal gray matter and white matter efficiency differentially predict intelligence in males and females. Human Brain Mapping, 37(11), 40064016.Google Scholar
Sampaio-Baptista, C., Khrapitchev, A. A., Foxley, S., Schlagheck, T., Scholz, J., Jbabdi, S., … Thomas, N. (2013). Motor skill learning induces changes in white matter microstructure and myelination. Journal of Neuroscience, 33(50), 1949919503.Google Scholar
Schmithorst, V. J. (2009). Developmental sex differences in the relation of neuroanatomical connectivity to intelligence. Intelligence, 37(2), 164173.CrossRefGoogle ScholarPubMed
Schmithorst, V. J., Wilke, M., Dardzinski, B. J., & Holland, S. K. (2005). Cognitive functions correlate with white matter architecture in a normal pediatric population: A diffusion tensor MRI study. Human Brain Mapping, 26(2), 139147.Google Scholar
Smith, S. M., Jenkinson, M., Johansen-Berg, H., Rueckert, D., Nichols, T. E., Mackay, C. E., … Matthews, P. M. (2006). Tract-based spatial statistics: Voxelwise analysis of multi-subject diffusion data. NeuroImage, 31(4), 14871505.CrossRefGoogle ScholarPubMed
Tamnes, C. K., Østby, Y., Walhovd, K. B., Westlye, L. T., Due‐Tønnessen, P., & Fjell, A. M. (2010). Intellectual abilities and white matter microstructure in development: A diffusion tensor imaging study. Human Brain Mapping, 31(10), 16091625.CrossRefGoogle ScholarPubMed
Tang, C. Y., Eaves, E. L., Ng, J. C., Carpenter, D. M., Mai, X., Schroeder, D. H., … Haier, R. J. (2010). Brain networks for working memory and factors of intelligence assessed in males and females with fMRI and DTI. Intelligence, 38(3), 293303.Google Scholar
Tuch, D. S. (2004). Q-ball imaging. Magnetic Resonance in Medicine, 52(6), 13581372.Google Scholar
Urger, S. E., De Bellis, M. D., Hooper, S. R., Woolley, D. P., Chen, S. D., & Provenzale, J. (2015). The superior longitudinal fasciculus in typically developing children and adolescents: Diffusion tensor imaging and neuropsychological correlates. Journal of Child Neurology, 30(1), 920.Google Scholar
Wang, Y., Adamson, C., Yuan, W., Altaye, M., Rajagopal, A., Byars, A. W., & Holland, S. K. (2012). Sex differences in white matter development during adolescence: A DTI study. Brain Research, 1 478, 115.Google Scholar
Wen, W., Zhu, W., He, Y., Kochan, N. A., Reppermund, S., Slavin, M. J., … Sachdev, P. (2011). Discrete neuroanatomical networks are associated with specific cognitive abilities in old age. Journal of Neuroscience, 31(4), 12041212.Google Scholar
Wiseman, S. J., Booth, T., Ritchie, S. J., Cox, S. R., Muñoz Maniega, S., Valdés Hernández, M., … Deary, I. J. (2018). Cognitive abilities, brain white matter hyperintensity volume, and structural network connectivity in older age. Human Brain Mapping, 39(2), 622632.CrossRefGoogle ScholarPubMed
Wolff, S. D., & Balaban, R. S. (1989). Magnetization transfer contrast (MTC) and tissue water proton relaxation in vivo. Magnetic Resonance in Medicine, 10(1), 135144.CrossRefGoogle ScholarPubMed
Yu, C., Li, J., Liu, Y., Qin, W., Li, Y., Shu, N., … Li, K. (2008). White matter tract integrity and intelligence in patients with mental retardation and healthy adults. NeuroImage, 40(4), 15331541.Google Scholar
Zhang, H., Schneider, T., Wheeler-Kingshott, C. A., & Alexander, D. C. (2012). NODDI: Practical in vivo neurite orientation dispersion and density imaging of the human brain. NeuroImage, 61(4), 10001016.CrossRefGoogle ScholarPubMed

References

Aiken, L. S., & West, S. G. (1991). Multiple regression: Testing and interpreting interactions. Thousand Oaks, CA: Sage Publications, Inc.Google Scholar
Aleman-Gomez, Y., Janssen, J., Schnack, H., Balaban, E., Pina-Camacho, L., Alfaro-Almagro, F., … Desco, M. (2013). The human cerebral cortex flattens during adolescence. Journal of Neuroscience, 33(38), 1500415010.Google Scholar
Andreasen, N. C., Flaum, M., Swayze, V., 2nd, O’Leary, D. S., Alliger, R., Cohen, G., … Yuh, W. T. (1993). Intelligence and brain structure in normal individuals. American Journal of Psychiatry, 150(1), 130134.Google Scholar
Ashburner, J., & Friston, K. J. (2000). Voxel-based morphometry – The methods. Neuroimage, 11(6 Pt 1), 805821.Google Scholar
Aydin, K., Ucar, A., Oguz, K. K., Okur, O. O., Agayev, A., Unal, Z., Yilmaz, S., and Ozturk, C. (2007). Increased gray matter density in the parietal cortex of mathematicians: A voxel-based morphometry study. AJNR American Journal of Neuroradiology, 28(10), 18591864.CrossRefGoogle ScholarPubMed
Bajaj, S., Raikes, A., Smith, R., Dailey, N. S., Alkozei, A., Vanuk, J. R., & Killgore, W. D. S. (2018). The relationship between general intelligence and cortical structure in healthy individuals. Neuroscience, 388, 3644.Google Scholar
Bassett, D. S., Bullmore, E., Verchinski, B. A., Mattay, V. S., Weinberger, D. R., & Meyer-Lindenberg, A. (2008). Hierarchical organization of human cortical networks in health and schizophrenia. Journal of Neuroscience, 28(37), 92399248.Google Scholar
Basten, U., Hilger, K., & Fiebach, C. J. (2015). Where smart brains are different: A quantitative meta-analysis of functional and structural brain imaging studies on intelligence. Intelligence, 51, 1027.Google Scholar
Bedford, S. A., Park, M. T. M., Devenyi, G. A., Tullo, S., Germann, J., Patel, R., … Consortium, Mrc Aims (2020). Large-scale analyses of the relationship between sex, age and intelligence quotient heterogeneity and cortical morphometry in autism spectrum disorder. Molecular Psychiatry, 25(3), 614628.Google Scholar
Bjuland, K. J., Løhaugen, G. C., Martinussen, M., & Skranes, J. (2013). Cortical thickness and cognition in very-low-birth-weight late teenagers. Early Human Development, 89(6), 371380.Google Scholar
Bourgeois, J. P., Goldman-Rakic, P. S., & Rakic, P. (1994). Synaptogenesis in the prefrontal cortex of rhesus monkeys. Cerebral Cortex, 4(1), 7896.Google Scholar
Breslau, N., Chilcoat, H. D., Susser, E. S., Matte, T., Liang, K.-Y., & Peterson, E. L. (2001). Stability and change in children’s intelligence quotient scores: A comparison of two socioeconomically disparate communities. American Journal of Epidemiology, 154(8), 711717.Google Scholar
Brouwer, R. M., Hedman, A. M., van Haren, N. E. M., Schnack, H. G., Brans, R. G. H., Smit, D. J. A., … Hulshoff Pol, H. E. (2014). Heritability of brain volume change and its relation to intelligence. Neuroimage, 100, 676683.Google Scholar
Budde, J., Shajan, G., Scheffler, K., & Pohmann, R. (2014). Ultra-high resolution imaging of the human brain using acquisition-weighted imaging at 9.4T. Neuroimage, 86, 592598.CrossRefGoogle ScholarPubMed
Burgaleta, M., Johnson, W., Waber, D. P., Colom, R., & Karama, S. (2014). Cognitive ability changes and dynamics of cortical thickness development in healthy children and adolescents. Neuroimage, 84, 810819.Google Scholar
Burgaleta, M., MacDonald, P. A., Martínez, K., Román, F. J., Álvarez-Linera, J., Ramos González, A., … Colom, R. (2014). Subcortical regional morphology correlates with fluid and spatial intelligence. Human Brain Mapping, 35(5), 19571968.Google Scholar
Button, K. S., Ioannidis, J. P. A., Mokrysz, C., Nosek, B. A., Flint, J., Robinson, E. S. J., & Munafò, M. R. (2013). Power failure: Why small sample size undermines the reliability of neuroscience. Nature Reviews Neuroscience, 14(5), 365376.CrossRefGoogle ScholarPubMed
Chance, S. A., Casanova, M. F., Switala, A. E., & Crow, T. J. (2008). Auditory cortex asymmetry, altered minicolumn spacing and absence of ageing effects in schizophrenia. Brain, 131(Pt 12), 31783192.Google Scholar
Chen, Z. J., He, Y., Rosa-Neto, P., Germann, J., & Evans, A. C. (2008). Revealing modular architecture of human brain structural networks by using cortical thickness from MRI. Cerebral Cortex, 18(10), 23742381.CrossRefGoogle ScholarPubMed
Chklovskii, D. B., Mel, B. W., & Svoboda, K. (2004). Cortical rewiring and information storage. Nature, 431(7010), 782788.Google Scholar
Choi, Y. Y., Shamosh, N. A., Cho, S. H., DeYoung, C. G., Lee, M. J., Lee, J. M., … Lee, K. H. (2008). Multiple bases of human intelligence revealed by cortical thickness and neural activation. Journal of Neuroscience, 28(41), 1032310329.Google Scholar
Cocosco, C. A., Zijdenbos, A. P., & Evans, A. C. (2003). A fully automatic and robust brain MRI tissue classification method. Medical Image Analysis, 7(4), 513527.Google Scholar
Collins, D. L., Neelin, P., Peters, T. M., & Evans, A. C. (1994). Automatic 3D intersubject registration of MR volumetric data in standardized Talairach space. Journal of Computer Assisted Tomography, 18(2), 192205.Google Scholar
Colom, R., Burgaleta, M., Román, F. J., Karama, S., Alvarez-Linera, J., Abad, F. J., … Haier, R. J. (2013). Neuroanatomic overlap between intelligence and cognitive factors: Morphometry methods provide support for the key role of the frontal lobes. Neuroimage, 72, 143152.Google Scholar
Colom, R., Haier, R. J., Head, K., Álvarez-Linera, J., Quiroga, M. Á., Shih, P. C., & Jung, R. E. (2009). Gray matter correlates of fluid, crystallized, and spatial intelligence: Testing the P-FIT model. Intelligence, 37(2), 124135.Google Scholar
Colom, R., Jung, R. E., & Haier, R. J. (2006). Distributed brain sites for the g-factor of intelligence. Neuroimage, 31(3), 13591365.Google Scholar
DeFelipe, J. (2011). The evolution of the brain, the human nature of cortical circuits, and intellectual creativity. Frontiers in Neuroanatomy, 5, 29.Google Scholar
Dubois, J., Galdi, P., Paul, L. K., & Adolphs, R. (2018). A distributed brain network predicts general intelligence from resting-state human neuroimaging data. Philosophical Transactions of the Royal Society B: Biological Sciences, 373(1756), 20170284.Google Scholar
Ducharme, S., Albaugh, M. D., Nguyen, T. V., Hudziak, J. J., Mateos-Perez, J. M., Labbe, A., … Brain Development Cooperative Group (2016). Trajectories of cortical thickness maturation in normal brain development – The importance of quality control procedures. Neuroimage, 125, 267279.CrossRefGoogle ScholarPubMed
Eickhoff, S. B., Constable, R. T., & Yeo, B. T. T. (2018). Topographic organization of the cerebral cortex and brain cartography. Neuroimage, 170, 332347.Google Scholar
Escorial, S., Román, F. J., Martínez, K., Burgaleta, M., Karama, S., & Colom, R. (2015). Sex differences in neocortical structure and cognitive performance: A surface-based morphometry study. Neuroimage, 104, 355365.Google Scholar
Estrada, E., Ferrer, E., Román, F. J., Karama, S., & Colom, R. (2019). Time-lagged associations between cognitive and cortical development from childhood to early adulthood. Developmental Psychology, 55(6), 13381352.Google Scholar
Evans, A. C., & Brain Development Cooperative Group (2006). The NIH MRI study of normal brain development. Neuroimage, 30(1), 184202.CrossRefGoogle ScholarPubMed
Evans, A. C., Janke, A. L., Collins, D. L., & Baillet, S. (2012). Brain templates and atlases. Neuroimage, 62(2), 911922.Google Scholar
Fjell, A. M., Westlye, L. T., Amlien, I., Tamnes, C. K., Grydeland, H., Engvig, A., … Walhovd, K. B. (2015). High-expanding cortical regions in human development and evolution are related to higher intellectual abilities. Cerebral Cortex, 25(1), 2634.Google Scholar
Flashman, L. A., Andreasen, N. C., Flaum, M., & Swayze, V. W. (1997). Intelligence and regional brain volumes in normal controls. Intelligence, 25(3), 149160.CrossRefGoogle Scholar
Frangou, S., Chitins, X., & Williams, S. C. (2004). Mapping IQ and gray matter density in healthy young people. Neuroimage, 23(3), 800805.Google Scholar
Ganjavi, H., Lewis, J. D., Bellec, P., MacDonald, P. A., Waber, D. P., Evans, A. C., … Brain Development Cooperative Group (2011). Negative associations between corpus callosum midsagittal area and IQ in a representative sample of healthy children and adolescents. PLoS One, 6(5), e19698.Google Scholar
Gautam, P., Anstey, K. J., Wen, W., Sachdev, P. S., & Cherbuin, N. (2015). Cortical gyrification and its relationships with cortical volume, cortical thickness, and cognitive performance in healthy mid-life adults. Behavioural Brain Research, 287, 331339.Google Scholar
Goh, S., Bansal, R., Xu, D., Hao, X., Liu, J., & Peterson, B. S. (2011). Neuroanatomical correlates of intellectual ability across the life span. Developmental Cognitive Neuroscience, 1(3), 305312.Google Scholar
Good, C. D., Johnsrude, I. S., Ashburner, J., Henson, R. N., Friston, K. J., & Frackowiak, R. S. (2001). A voxel-based morphometric study of ageing in 465 normal adult human brains. Neuroimage, 14(1 Pt 1), 2136.Google Scholar
Green, S., Blackmon, K., Thesen, T., DuBois, J., Wang, X., Halgren, E., & Devinsky, O. (2018). Parieto-frontal gyrification and working memory in healthy adults. Brain Imaging Behavior, 12(2), 303308.CrossRefGoogle ScholarPubMed
Gregory, M. D., Kippenhan, J. S., Dickinson, D., Carrasco, J., Mattay, V. S., Weinberger, D. R., & Berman, K. F. (2016). Regional variations in brain gyrification are associated with general cognitive ability in humans. Current Biology, 26(10), 13011305.Google Scholar
Gur, R. C., Turetsky, B. I., Matsui, M., Yan, M., Bilker, W., Hughett, P., & Gur, R. E. (1999). Sex differences in brain gray and white matter in healthy young adults: Correlations with cognitive performance. Journal of Neuroscience, 19(10), 40654072.Google Scholar
Haier, R. J. (2016). The neuroscience of intelligence. Cambridge University Press.Google Scholar
Haier, R. J., Jung, R. E., Yeo, R. A., Head, K., & Alkire, M. T. (2004). Structural brain variation and general intelligence. Neuroimage, 23(1), 425433.Google Scholar
Haier, R. J., Jung, R. E., Yeo, R. A., Head, K., & Alkire, M. T. (2005). The neuroanatomy of general intelligence: Sex matters. Neuroimage, 25(1), 320327.Google Scholar
Haier, R. J., Karama, S., Colom, R., Jung, R., & Johnson, W. (2014). Yes, but flaws remain. Intelligence, 46, 341344.Google Scholar
He, Y., Chen, Z. J., & Evans, A. C. (2007). Small-world anatomical networks in the human brain revealed by cortical thickness from MRI. Cerebral Cortex, 17(10), 24072419.CrossRefGoogle ScholarPubMed
Hogstrom, L. J., Westlye, L. T., Walhovd, K. B., & Fjell, A. M. (2013). The structure of the cerebral cortex across adult life: Age-related patterns of surface area, thickness, and gyrification. Cerebral Cortex, 23(11), 25212530.Google Scholar
Huttenlocher, P. R. (1990). Morphometric study of human cerebral cortex development. Neuropsychologia, 28(6), 517527.Google Scholar
Jung, R. E., & Haier, R. J. (2007). The Parieto-Frontal Integration Theory (P-FIT) of intelligence: Converging neuroimaging evidence. Behavioral and Brain Sciences, 30(2), 135154.Google Scholar
Kabani, N., Le Goualher, G., MacDonald, D., & Evans, A. C. (2001). Measurement of cortical thickness using an automated 3-D algorithm: A validation study. Neuroimage, 13(2), 375380.Google Scholar
Karama, S., Ad-Dab’bagh, Y., Haier, R. J., Deary, I. J., Lyttelton, O. C., Lepage, C., … Brain Development Cooperative Group (2009). Positive association between cognitive ability and cortical thickness in a representative US sample of healthy 6 to 18 year-olds. Intelligence, 37(2), 145155.Google Scholar
Karama, S., Bastin, M. E., Murray, C., Royle, N. A., Penke, L., Muñoz Maniega, S., … Deary, I. J. (2014). Childhood cognitive ability accounts for associations between cognitive ability and brain cortical thickness in old age. Molecular Psychiatry, 19(5), 555559.CrossRefGoogle ScholarPubMed
Karama, S., Colom, R., Johnson, W., Deary, I. J., Haier, R., Waber, D. P., … Brain Development Cooperative Group (2011). Cortical thickness correlates of specific cognitive performance accounted for by the general factor of intelligence in healthy children aged 6 to 18. Neuroimage, 55(4), 14431453.Google Scholar
Kennedy, D. N., Lange, N., Makris, N., Bates, J., Meyer, J., & Caviness, V. S. Jr. (1998). Gyri of the human neocortex: An MRI-based analysis of volume and variance. Cerebral Cortex, 8(4), 372384.Google Scholar
Khundrakpam, B. S., Reid, A., Brauer, J., Carbonell, F., Lewis, J., Ameis, S., … Brain Development Cooperative Group (2013). Developmental changes in organization of structural brain networks. Cerebral Cortex, 23(9), 20722085.Google Scholar
Kim, J. S., Singh, V., Lee, J. K., Lerch, J., Ad-Dab’bagh, Y., MacDonald, D., … Evans, A. C. (2005). Automated 3-D extraction and evaluation of the inner and outer cortical surfaces using a Laplacian map and partial volume effect classification. Neuroimage, 27(1), 210221.Google Scholar
la Fougere, C., Grant, S., Kostikov, A., Schirrmacher, R., Gravel, P., Schipper, H. M., … Thiel, A. (2011). Where in-vivo imaging meets cytoarchitectonics: The relationship between cortical thickness and neuronal density measured with high-resolution [18F]flumazenil-PET. Neuroimage, 56(3), 951960.Google Scholar
Lemaitre, H., Goldman, A. L., Sambataro, F., Verchinski, B. A., Meyer-Lindenberg, A., Weinberger, D. R., & Mattay, V. S. (2012). Normal age-related brain morphometric changes: Nonuniformity across cortical thickness, surface area and gray matter volume? Neurobiology of Aging, 33(3), 617.e1–617.e9.Google Scholar
Lenroot, R. K., Gogtay, N., Greenstein, D. K., Wells, E. M., Wallace, G. L., Clasen, L. S., … Giedd, J. N. (2007). Sexual dimorphism of brain developmental trajectories during childhood and adolescence. Neuroimage, 36(4), 10651073.Google Scholar
Lerch, J. P., & Evans, A. C. (2005). Cortical thickness analysis examined through power analysis and a population simulation. Neuroimage, 24(1), 163173.Google Scholar
Lerch, J. P., Worsley, K., Shaw, W. P., Greenstein, D. K., Lenroot, R. K., Giedd, J., & Evans, A. C. (2006). Mapping anatomical correlations across cerebral cortex (MACACC) using cortical thickness from MRI. Neuroimage, 31(3), 9931003.Google Scholar
Li, W., Yang, C., Shi, F., Wu, S., Wang, Q., Nie, Y., & Zhang, X. (2017). Construction of individual morphological brain networks with multiple morphometric features. Frontiers in Neuroanatomy, 11, 34.CrossRefGoogle ScholarPubMed
Lo, C. Y., He, Y., & Lin, C. P. (2011). Graph theoretical analysis of human brain structural networks. Reviews Neuroscience, 22(5), 551563.Google Scholar
Luders, E., Narr, K. L., Bilder, R. M., Thompson, P. M., Szeszko, P. R., Hamilton, L., & Toga, A. W. (2007). Positive correlations between corpus callosum thickness and intelligence. Neuroimage, 37(4), 14571464.Google Scholar
Luders, E., Narr, K. L., Thompson, P. M., & Toga, A. W. (2009). Neuroanatomical correlates of intelligence. Intelligence, 37(2), 156163.Google Scholar
Luders, E., Thompson, P. M., Narr, K. L., Zamanyan, A., Chou, Y. Y., Gutman, B., … Toga, A. W. (2011). The link between callosal thickness and intelligence in healthy children and adolescents. Neuroimage, 54(3), 18231830.Google Scholar
Lyttelton, O. C., Karama, S., Ad-Dab’bagh, Y., Zatorre, R. J., Carbonell, F., Worsley, K., & Evans, A. C. (2009). Positional and surface area asymmetry of the human cerebral cortex. Neuroimage, 46(4), 895903.Google Scholar
MacDonald, D., Kabani, N., Avis, D., & Evans, A. C. (2000). Automated 3-D extraction of inner and outer surfaces of cerebral cortex from MRI. Neuroimage, 12(3), 340356.Google Scholar
MacDonald, P. A., Ganjavi, H., Collins, D. L., Evans, A. C., & Karama, S. (2014). Investigating the relation between striatal volume and IQ. Brain Imaging and Behavior, 8(1), 5259.Google Scholar
McDaniel, M. A. (2005). Big-brained people are smarter: A meta-analysis of the relationship between in vivo brain volume and intelligence. Intelligence, 33(4), 337346.Google Scholar
Menary, K., Collins, P. F., Porter, J. N., Muetzel, R., Olson, E. A., Kumar, V., … Luciana, M. (2013). Associations between cortical thickness and general intelligence in children, adolescents and young adults. Intelligence, 41(5), 597606.Google Scholar
Modroño, C., Navarrete, G., Nicolle, A., González-Mora, J. L., Smith, K. W., Marling, M., & Goel, V. (2019). Developmental grey matter changes in superior parietal cortex accompany improved transitive reasoning. Thinking & Reasoning, 25(2), 151170.Google Scholar
Moffitt, T. E., Caspi, A., Harkness, A. R., & Silva, P. A. (1993). The natural history of change in intellectual performance: Who changes? How much? Is it meaningful? Journal of Child Psychology and Psychiatry, 34(4), 455506.Google Scholar
Narr, K. L., Woods, R. P., Thompson, P. M., Szeszko, P., Robinson, D., Dimtcheva, T., … Bilder, R. M. (2007). Relationships between IQ and regional cortical gray matter thickness in healthy adults. Cerebral Cortex, 17(9), 21632171.Google Scholar
Panizzon, M. S., Fennema-Notestine, C., Eyler, L. T., Jernigan, T. L., Prom-Wormley, E., Neale, M., … Kremen, W. S. (2009). Distinct genetic influences on cortical surface area and cortical thickness. Cerebral Cortex, 19(11), 27282735.Google Scholar
Paradiso, S., Andreasen, N. C., O’Leary, D. S., Arndt, S., & Robinson, R. G. (1997). Cerebellar size and cognition: Correlations with IQ, verbal memory and motor dexterity. Neuropsychiatry, Neuropsychology, and Behavioral Neurology, 10(1), 18.Google Scholar
Paul, E. J., Larsen, R. J., Nikolaidis, A., Ward, N., Hillman, C. H., Cohen, N. J., … Barbey, A. K. (2016). Dissociable brain biomarkers of fluid intelligence. Neuroimage, 137, 201211.Google Scholar
Paus, T., Zijdenbos, A., Worsley, K., Collins, D. L., Blumenthal, J., Giedd, J. N., … Evans, A. C. (1999). Structural maturation of neural pathways in children and adolescents: In vivo study. Science, 283(5409), 19081911.Google Scholar
Pietschnig, J., Penke, L., Wicherts, J. M., Zeiler, M., & Voracek, M. (2015). Meta-analysis of associations between human brain volume and intelligence differences: How strong are they and what do they mean? Neuroscience & Biobehavioral Reviews, 57, 411432.Google Scholar
Rakic, P. (1988). Specification of cerebral cortical areas. Science, 241(4862), 170176.Google Scholar
Raznahan, A., Shaw, P., Lalonde, F., Stockman, M., Wallace, G. L., Greenstein, D., … Giedd, J. N. (2011). How does your cortex grow? The Journal of Neuroscience: The Official Journal of the Society for Neuroscience, 31(19), 71747177.Google Scholar
Regis, J., Mangin, J. F., Ochiai, T., Frouin, V., Riviere, D., Cachia, A., … Samson, Y. (2005). “Sulcal root” generic model: A hypothesis to overcome the variability of the human cortex folding patterns. Neurologia Medico-Chirurgica (Tokyo), 45(1), 117.Google Scholar
Reiss, A. L., Abrams, M. T., Singer, H. S., Ross, J. L., & Denckla, M. B. (1996). Brain development, gender and IQ in children. A volumetric imaging study. Brain, 119(Pt 5), 17631774.Google Scholar
Reuter, M., Tisdall, M. D., Qureshi, A., Buckner, R. L., van der Kouwe, A. J. W., & Fischl, B. (2015). Head motion during MRI acquisition reduces gray matter volume and thickness estimates. Neuroimage, 107, 107115.Google Scholar
Riahi, F., Zijdenbos, A., Narayanan, S., Arnold, D., Francis, G., Antel, J., & Evans, A. C. (1998). Improved correlation between scores on the expanded disability status scale and cerebral lesion load in relapsing-remitting multiple sclerosis. Results of the application of new imaging methods. Brain, 121(Pt 7), 13051312.Google Scholar
Richman, D. P., Stewart, R. M., Hutchinson, J. W., & Caviness, V. S. Jr. (1975). Mechanical model of brain convolutional development. Science, 189(4196), 1821.Google Scholar
Rilling, J. K., & Insel, T. R. (1999). The primate neocortex in comparative perspective using magnetic resonance imaging. Journal of Human Evolution, 37(2), 191223.Google Scholar
Ritchie, S. J., Booth, T., Valdes Hernandez, M. D., Corley, J., Maniega, S. M., Gow, A. J., … Deary, I. J. (2015). Beyond a bigger brain: Multivariable structural brain imaging and intelligence. Intelligence, 51, 4756.Google Scholar
Riva, D., & Giorgi, C. (2000). The cerebellum contributes to higher functions during development: Evidence from a series of children surgically treated for posterior fossa tumours. Brain, 123(5), 10511061.Google Scholar
Román, F. J., Morillo, D., Estrada, E., Escorial, S., Karama, S., & Colom, R. (2018). Brain-intelligence relationships across childhood and adolescence: A latent-variable approach. Intelligence, 68, 2129.Google Scholar
Roth, G., & Dicke, U. (2005). Evolution of the brain and intelligence. Trends in Cognitive Sciences, 9(5), 250257.Google Scholar
Rushton, J. P., & Ankney, C. D. (2009). Whole brain size and general mental ability: A review. International Journal of Neuroscience, 119(5), 691731.Google Scholar
Sanabria-Diaz, G., Melie-Garcia, L., Iturria-Medina, Y., Aleman-Gomez, Y., Hernandez-Gonzalez, G., Valdes-Urrutia, L., … Valdes-Sosa, P. (2010). Surface area and cortical thickness descriptors reveal different attributes of the structural human brain networks. Neuroimage, 50(4), 14971510.CrossRefGoogle ScholarPubMed
Schmahmann, J. D. (2004). Disorders of the cerebellum: Ataxia, dysmetria of thought, and the cerebellar cognitive affective syndrome. The Journal of Neuropsychiatry and Clinical Neurosciences, 16(3), 367378.Google Scholar
Schmitt, J. E., Neale, M. C., Clasen, L. S., Liu, S., Seidlitz, J., Pritikin, J. N., … Raznahan, A. (2019). A comprehensive quantitative genetic analysis of cerebral surface area in youth. Journal of Neuroscience, 39(16), 30283040.Google Scholar
Schmitt, J. E., Raznahan, A., Clasen, L. S., Wallace, G. L., Pritikin, J. N., Lee, N. R., … Neale, M. C. (2019). The dynamic associations between cortical thickness and general intelligence are genetically mediated. Cerebral Cortex, 29(11), 47434752.Google Scholar
Schoenemann, P. T., Budinger, T. F., Sarich, V. M., & Wang, W. S. Y. (2000). Brain size does not predict general cognitive ability within families. Proceedings of the National Academy of Sciences, 97(9), 49324937.Google Scholar
Schulte, T., & Muller-Oehring, E. M. (2010). Contribution of callosal connections to the interhemispheric integration of visuomotor and cognitive processes. Neuropsychology Review, 20(2), 174190.Google Scholar
Shaw, P., Greenstein, D., Lerch, J., Clasen, L., Lenroot, R., Gogtay, N., … Giedd, J. (2006). Intellectual ability and cortical development in children and adolescents. Nature, 440(7084), 676679.Google Scholar
Sowell, E. R., Thompson, P. M., Leonard, C. M., Welcome, S. E., Kan, E., & Toga, A. W. (2004). Longitudinal mapping of cortical thickness and brain growth in normal children. The Journal of Neuroscience, 24(38), 8223.Google Scholar
Stonnington, C. M., Tan, G., Klöppel, S., Chu, C., Draganski, B., Jack, C. R. Jr., … Frackowiak, R. S. (2008). Interpreting scan data acquired from multiple scanners: a study with Alzheimer’s disease. Neuroimage, 39(3), 11801185.Google Scholar
Storsve, A. B., Fjell, A. M., Tamnes, C. K., Westlye, L. T., Overbye, K., Aasland, H. W., & Walhovd, K. B. (2014). Differential longitudinal changes in cortical thickness, surface area and volume across the adult life span: Regions of accelerating and decelerating change. Journal of Neuroscience, 34(25), 84888498.Google Scholar
Stucht, D., Danishad, K. A., Schulze, P., Godenschweger, F., Zaitsev, M., & Speck, O. (2015). Highest resolution in vivo human brain MRI using prospective motion correction. PLoS One, 10(7), e0133921.Google Scholar
Sur, M., & Rubenstein, J. L. (2005). Patterning and plasticity of the cerebral cortex. Science, 310(5749), 805810.Google Scholar
Tadayon, E., Pascual-Leone, A., & Santarnecchi, E. (2019). Differential contribution of cortical thickness, surface area, and gyrification to fluid and crystallized intelligence. Cerebral Cortex, 30(1).Google Scholar
Tamnes, C. K., Fjell, A. M., Østby, Y., Westlye, L. T., Due-Tønnessen, P., Bjørnerud, A., & Walhovd, K. B. (2011). The brain dynamics of intellectual development: Waxing and waning white and gray matter. Neuropsychologia, 49(13), 36053611.Google Scholar
Thompson, P. M., Hayashi, K. M., Dutton, R. A., Chiang, M.-C., Leow, A. D., Sowell, E. R., … Toga, A. W. (2007). Tracking Alzheimer’s disease. Annals of the New York Academy of Science, 1 097, 183214.Google Scholar
Thompson, P. (2020). ENIGMA and global neuroscience: A decade of large-scale studies of the brain in health and disease across more than 40 countries. Biological Psychiatry, 87(9, Suppl), S56.Google Scholar
Turin, G. (1960). An introduction to matched filters. IRE Transactions on Information Theory, 6(3), 311329.Google Scholar
Van Essen, D. C. (2005). A population-average, landmark- and surface-based (PALS) atlas of human cerebral cortex. Neuroimage, 28(3), 635662.CrossRefGoogle ScholarPubMed
Vuoksimaa, E., Panizzon, M. S., Chen, C.-H., Fiecas, M., Eyler, L. T., Fennema-Notestine, C., … Kremen, W. S. (2015). The genetic association between neocortical volume and general cognitive ability is driven by global surface area rather than thickness. Cerebral Cortex, 25(8), 21272137.Google Scholar
Watson, P. D., Paul, E. J., Cooke, G. E., Ward, N., Monti, J. M., Horecka, K. M., … Barbey, A. K. (2016). Underlying sources of cognitive-anatomical variation in multi-modal neuroimaging and cognitive testing. Neuroimage, 129, 439449.Google Scholar
Westerhausen, R., Friesen, C. M., Rohani, D. A., Krogsrud, S. K., Tamnes, C. K., Skranes, J. S., … Walhovd, K. B. (2018). The corpus callosum as anatomical marker of intelligence? A critical examination in a large-scale developmental study. Brain Structure and Function, 223(1), 285296.Google Scholar
Westlye, L. T., Walhovd, K. B., Dale, A. M., Bjørnerud, A., Due-Tønnessen, P., Engvig, A., … Fjell, A. M. (2009). Life-span changes of the human brain white matter: Diffusion tensor imaging (DTI) and volumetry. Cerebral Cortex, 20(9), 20552068.Google Scholar
Wickett, J. C., Vernon, P. A., & Lee, D. H. (2000). Relationships between factors of intelligence and brain volume. Personality and Individual Differences, 29(6), 10951122.Google Scholar
Winkler, A. M., Kochunov, P., Blangero, J., Almasy, L., Zilles, K., Fox, P. T., … Glahn, D. C. (2010). Cortical thickness or grey matter volume? The importance of selecting the phenotype for imaging genetics studies. Neuroimage, 53(3), 11351146.Google Scholar
Winkler, A. M., Sabuncu, M. R., Yeo, B. T., Fischl, B., Greve, D. N., Kochunov, P., … Glahn, D. C. (2012). Measuring and comparing brain cortical surface area and other areal quantities. Neuroimage, 61(4), 14281443.Google Scholar
Worsley, K. J., Marrett, S., Neelin, P., Vandal, A. C., Friston, K. J., & Evans, A. C. (1996). A unified statistical approach for determining significant signals in images of cerebral activation. Human Brain Mapping, 4(1), 5873.Google Scholar
Xie, Y., Chen, Y. A., & De Bellis, M. D. (2012). The relationship of age, gender, and IQ with the brainstem and thalamus in healthy children and adolescents: A magnetic resonance imaging volumetric study. Journal of Child Neurology, 27(3), 325331.Google Scholar
Zatorre, R. J., Fields, R. D., & Johansen-Berg, H. (2012). Plasticity in gray and white: Neuroimaging changes in brain structure during learning. Nature Neuroscience, 15(4), 528536.Google Scholar
Zijdenbos, A. P., Lerch, J. P., Bedell, B. J., & Evans, A. C. (2005). Brain imaging in drug R&D. Biomarkers 10(Suppl 1), S58S68.Google Scholar
Zilles, K., Armstrong, E., Schleicher, A., & Kretschmann, H. J. (1988). The human pattern of gyrification in the cerebral cortex. Anatomy and Embryology (Berlin), 179(2), 173179.Google Scholar

References

Barbey, A. K. (2018). Network neuroscience theory of human intelligence. Trends in Cognitive Science, 22(1), 820.Google Scholar
Barbey, A. K., Colom, R., & Grafman, J. (2013). Dorsolateral prefrontal contributions to human intelligence. Neuropsychologia, 51(7), 13611369.Google Scholar
Barbey, A. K., Colom, R., Paul, E. J., & Grafman, J. (2014). Architecture of fluid intelligence and working memory revealed by lesion mapping. Brain Structure and Function, 219, 485494.Google Scholar
Barbey, A. K., Colom, R., Solomon, J., Krueger, F., Forbes, C., & Grafman, J. (2012). An integrative architecture for general intelligence and executive function revealed by lesion mapping. Brain, 135(4), 11541164. doi: 10.1093/brain/aws021.Google Scholar
Basten, U., Hilger, K., & Fiebach, C. J. (2015). Where smart brains are different: A quantitative meta-analysis of functional and structural brain imaging studies on intelligence. Intelligence, 51, 1027. doi: 10.1016/j.intell.2015.04.009.Google Scholar
Basten, U., Stelzel, C., & Fiebach, C. J. (2011). Trait anxiety modulates the neural efficiency of inhibitory control. Journal of Cognitive Neuroscience, 23(10), 31323145. doi: 10.1162/jocn_a_00003.Google Scholar
Basten, U., Stelzel, C., & Fiebach, C. J. (2012). Trait anxiety and the neural efficiency of manipulation in working memory. Cognitive, Affective, & Behavioral Neuroscience, 12(3), 571588. doi: 10.3758/s13415–012-0100-3.Google Scholar
Basten, U., Stelzel, C., & Fiebach, C. J. (2013). Intelligence is differentially related to neural effort in the task-positive and the task-negative brain network. Intelligence, 41(5), 517528. doi: 10.1016/j.intell.2013.07.006.Google Scholar
Berent, S., Giordani, B., Lehtinen, S., Markel, D., Penney, J. B., Buchtel, H. A., … Young, A. B. (1988). Positron emission tomographic scan investigations of Huntington’s disease: Cerebral metabolic correlates of cognitive function. Annals of Neurology, 23(6), 541546. doi: 10.1002/ana.410230603.Google Scholar
Burgess, G. C., Gray, J. R., Conway, A. R. A., & Braver, T. S. (2011). Neural mechanisms of interference control underlie the relationship between fluid intelligence and working memory span. Journal of Experimental Psychology: General, 140(4), 674692. doi: 10.1037/a0024695.Google Scholar
Button, K. S., Ioannidis, J. P. A., Mokrysz, C., Nosek, B. A., Flint, J., Robinson, E. S. J., & Munafò, M. R. (2013). Power failure: Why small sample size undermines the reliability of neuroscience. Nature Reviews Neuroscience, 14(5), 365376. doi: 10.1038/nrn3475.Google Scholar
Cabeza, R., & Nyberg, L. (2000). Imaging cognition II: An empirical review of 275 PET and fMRI studies. Journal of Cognitive Neuroscience, 12(1), 147. doi: 10.1162/08989290051137585.Google Scholar
Cattell, R. B. (1963). Theory of fluid and crystallized intelligence: A critical experiment. Journal of Educational Psychology, 54(1), 122. doi: 10.1037/h0046743.Google Scholar
Chang, L. J., Yarkoni, T., Khaw, M. W., & Sanfey, A. G. (2013). Decoding the role of the insula in human cognition: Functional parcellation and large-scale reverse inference. Cerebral Cortex, 23(3), 739749. doi: 10.1093/cercor/bhs065.Google Scholar
Choi, Y. Y., Shamosh, N. A., Cho, S. H., DeYoung, C. G., Lee, M. J., Lee, J.-M., … Lee, K. H. (2008). Multiple bases of human intelligence revealed by cortical thickness and neural activation. Journal of Neuroscience, 28(41), 1032310329. doi: 10.1523/JNEUROSCI.3259-08.2008.Google Scholar
Cole, M. W., & Schneider, W. (2007). The cognitive control network: Integrated cortical regions with dissociable functions. NeuroImage, 37(1), 343360. doi: 10.1016/j.neuroimage.2007.03.071.Google Scholar
Corbetta, M., Patel, G., & Shulman, G. L. (2008). The reorienting system of the human brain: From environment to theory of mind. Neuron, 58(3), 306324. doi: 10.1016/j.neuron.2008.04.017.Google Scholar
Cremers, H. R., Wager, T. D., & Yarkoni, T. (2017). The relation between statistical power and inference in fMRI. PLoS One, 12(11), e0184923. doi: 10.1371/journal.pone.0184923.Google Scholar
Daugherty, A. M., Sutton, B. P., Hillman, C. H., Kramer, A. F., Cohen, N. J., & Barbey, A. K. (2020). Individual differences in the neurobiology of fluid intelligence predict responsiveness to training: Evidence from a comprehensive cognitive, mindfulness meditation, and aerobic fitness intervention. Trends in Neuroscience and Education, 18, 100123. doi: 10.1016/j.tine.2019.100123.Google Scholar
Daugherty, A. M., Zwilling, C., Paul, E. J., Sherepa, N., Allen, C., Kramer, A. F., … Barbey, A. K. (2018). Multi-modal fitness and cognitive training to enhance fluid intelligence. Intelligence, 66, 3243.Google Scholar
Derrfuss, J., Vogt, V. L., Fiebach, C. J., von Cramon, D. Y., & Tittgemeyer, M. (2012). Functional organization of the left inferior precentral sulcus: Dissociating the inferior frontal eye field and the inferior frontal junction. NeuroImage, 59(4), 38293837. doi: 10.1016/j.neuroimage.2011.11.051.Google Scholar
DeYoung, C. G., Shamosh, N. A., Green, A. E., Braver, T. S., & Gray, J. R. (2009). Intellect as distinct from openness: Differences revealed by fMRI of working memory. Journal of Personality and Social Psychology, 97(5), 883892. doi: 10.1037/a0016615.Google Scholar
Dosenbach, N. U. F., Fair, D. A., Miezin, F. M., Cohen, A. L., Wenger, K. K., Dosenbach, R. A. T., … Petersen, S. E. (2007). Distinct brain networks for adaptive and stable task control in humans. Proceedings of the National Academy of Sciences, 104(26), 1107311078. doi: 10.1073/pnas.0704320104.Google Scholar
Duncan, J. (1995). Attention, intelligence, and the frontal lobes. In Gazzaniga, M. S. (ed.), The cognitive neurosciences (pp. 721733). Cambridge, MA: The MIT Press.Google Scholar
Duncan, J., Seitz, R. J., Kolodny, J., Bor, D., Herzog, H., Ahmed, A., … Emslie, H. (2000). A neural basis for general intelligence. Science, 289(5478), 457460. doi: 10.1126/science.289.5478.457.Google Scholar
Duncan, J. (2005). Frontal lobe function and general intelligence: Why it matters. Cortex, 41(2), 215217. doi: 10.1016/S0010–9452(08)70896-7.Google Scholar
Duncan, J. (2010). The multiple-demand (MD) system of the primate brain: Mental programs for intelligent behaviour. Trends in Cognitive Sciences, 14(4), 172179. doi: 10.1016/j.tics.2010.01.004.Google Scholar
Duncan, J., Burgess, P., & Emslie, H. (1995). Fluid intelligence after frontal lobe lesions. Neuropsychologia, 33(3), 261268. doi: 10.1016/0028-3932(94)00124-8.Google Scholar
Duncan, J., Emslie, H., Williams, P., Johnson, R., & Freer, C. (1996). Intelligence and the frontal lobe: The organization of goal-directed behavior. Cognitive Psychology, 30(3), 257303. doi: 10.1006/cogp.1996.0008.Google Scholar
Ebisch, S. J., Perrucci, M. G., Mercuri, P., Romanelli, R., Mantini, D., Romani, G. L., … Saggino, A. (2012). Common and unique neuro-functional basis of induction, visualization, and spatial relationships as cognitive components of fluid intelligence. NeuroImage, 62(1), 331342. doi: 10.1016/j.neuroimage.2012.04.053.Google Scholar
Esposito, G., Kirkby, B. S., Van Horn, J. D., Ellmore, T. M., & Berman, K. F. (1999). Context-dependent, neural system-specific neurophysiological concomitants of ageing: Mapping PET correlates during cognitive activation. Brain: A Journal of Neurology, 122(Pt 5), 963979. doi: 10.1093/brain/122.5.963.Google Scholar
Euler, M. J., Weisend, M. P., Jung, R. E., Thoma, R. J., & Yeo, R. A. (2015). Reliable activation to novel stimuli predicts higher fluid intelligence. NeuroImage, 114, 311319. doi: 10.1016/j.neuroimage.2015.03.078.Google Scholar
Fox, M. D., Snyder, A. Z., Vincent, J. L., Corbetta, M., Van Essen, D. C., & Raichle, M. E. (2005). The human brain is intrinsically organized into dynamic, anticorrelated functional networks. Proceedings of the National Academy of Sciences of the United States of America, 102(27), 9673. doi: 10.1073/pnas.0504136102.Google Scholar
Geake, J. G., & Hansen, P. C. (2005). Neural correlates of intelligence as revealed by fMRI of fluid analogies. NeuroImage, 26(2), 555564. doi: 10.1016/j.neuroimage.2005.01.035.Google Scholar
Genç, E., Fraenz, C., Schlüter, C., Friedrich, P., Hossiep, R., Voelkle, M. C., … Jung, R. E. (2018). Diffusion markers of dendritic density and arborization in gray matter predict differences in intelligence. Nature Communications, 9(1), 1905. doi: 10.1038/s41467–018-04268-8.Google Scholar
Ghatan, P. H., Hsieh, J. C., Wirsén-Meurling, A., Wredling, R., Eriksson, L., Stone-Elander, S., … Ingvar, M. (1995). Brain activation induced by the perceptual maze test: A PET study of cognitive performance. NeuroImage, 2(2), 112124.CrossRefGoogle ScholarPubMed
Gläscher, J., Rudrauf, D., Colom, R., Paul, L. K., Tranel, D., Damasio, H., & Adolphs, R. (2010). Distributed neural system for general intelligence revealed by lesion mapping. Proceedings of the National Academy of Sciences, 107(10), 47054709. doi: 10.1073/pnas.0910397107.Google Scholar
Goel, V., & Dolan, R. J. (2001). Functional neuroanatomy of three-term relational reasoning. Neuropsychologia, 39(9), 901909.Google Scholar
Goel, V., Gold, B., Kapur, S., & Houle, S. (1998). Neuroanatomical correlates of human reasoning. Journal of Cognitive Neuroscience, 10(3), 293302. doi: 10.1162/089892998562744.Google Scholar
Grabner, R. H., Neubauer, A. C., & Stern, E. (2006). Superior performance and neural efficiency: The impact of intelligence and expertise. Brain Research Bulletin, 69(4), 422439. doi: 10.1016/j.brainresbull.2006.02.009.Google Scholar
Grabner, R. H., Stern, E., & Neubauer, A. C. (2003). When intelligence loses its impact: Neural efficiency during reasoning in a familiar area. International Journal of Psychophysiology, 49(2), 8998. doi: 10.1016/S0167–8760(03)00095-3.Google Scholar
Gray, J. R., Chabris, C. F., & Braver, T. S. (2003). Neural mechanisms of general fluid intelligence. Nature Neuroscience, 6(3), 316322. doi: 10.1038/nn1014.Google Scholar
Gregory, M. D., Kippenhan, J. S., Dickinson, D., Carrasco, J., Mattay, V. S., Weinberger, D. R., & Berman, K. F. (2016). Regional variations in brain gyrification are associated with general cognitive ability in humans. Current Biology, 26(10), 13011305. doi: 10.1016/j.cub.2016.03.021.Google Scholar
Haier, R. (2016). The neuroscience of intelligence (Cambridge fundamentals of neuroscience in psychology). Cambridge University Press. doi: 10.1017/9781316105771.Google Scholar
Haier, R. J., Siegel, B. V., Nuechterlein, K. H., Hazlett, E., Wu, J. C., Paek, J., … Buchsbaum, M. S. (1988). Cortical glucose metabolic rate correlates of abstract reasoning and attention studied with positron emission tomography. Intelligence, 12(2), 199217. doi: 10.1016/0160-2896(88)90016-5.Google Scholar
Haier, R. J., Siegel, B., Tang, C., Abel, L., & Buchsbaum, M. S. (1992). Intelligence and changes in regional cerebral glucose metabolic rate following learning. Intelligence, 16(3–4), 415426. do: 10.1016/0160-2896(92)90018-M.Google Scholar
Hammer, R., Paul, E. J., Hillman, C. H., Kramer, A. F., Cohen, N. J., & Barbey, A. K. (2019). Individual differences in analogical reasoning revealed by multivariate task-based functional brain imaging. Neuroimage, 184, 9931004.Google Scholar
Hilger, K., Ekman, M., Fiebach, C. J., & Basten, U. (2017a). Efficient hubs in the intelligent brain: Nodal efficiency of hub regions in the salience network is associated with general intelligence. Intelligence, 60, 1025. doi: 10.1016/j.intell.2016.11.001.Google Scholar
Hilger, K., Ekman, M., Fiebach, C. J., & Basten, U. (2017b). Intelligence is associated with the modular structure of intrinsic brain networks. Scientific Reports, 7(1), 16088. doi: 10.1038/s41598–017-15795-7.Google Scholar
Ioannidis, J. P. A. (2008). Why most discovered true associations are inflated. Epidemiology, 19(5), 640648. doi: 10.1097/EDE.0b013e31818131e7.Google Scholar
Jaušovec, N. (2000). Differences in cognitive processes between gifted, intelligent, creative, and average individuals while solving complex problems: An EEG study. Intelligence, 28(3), 213237. doi: 10.1016/S0160–2896(00)00037-4.Google Scholar
Jaušovec, N., & Jaušovec, K. (2004). Differences in induced brain activity during the performance of learning and working-memory tasks related to intelligence. Brain and Cognition, 54(1), 6574. doi: 10.1016/S0278–2626(03)00263-X.Google Scholar
Jung, R. E., & Haier, R. J. (2007). The Parieto-Frontal Integration Theory (P-FIT) of intelligence: Converging neuroimaging evidence. Behavioral and Brain Sciences, 30(2), 135154. doi: 10.1017/S0140525X07001185.Google Scholar
Kievit, R. A., Davis, S. W., Griffiths, J., Correia, M. M., Cam-Can, , & Henson, R. N. (2016). A watershed model of individual differences in fluid intelligence. Neuropsychologia, 91, 186198. doi: 10.1016/j.neuropsychologia.2016.08.008.Google Scholar
Knauff, M., Mulack, T., Kassubek, J., Salih, H. R., & Greenlee, M. W. (2002). Spatial imagery in deductive reasoning: A functional MRI study. Brain Research Cognitive Brain Research, 13(2), 203212.Google Scholar
Kruschwitz, J. D., Waller, L., Daedelow, L. S., Walter, H., & Veer, I. M. (2018). General, crystallized and fluid intelligence are not associated with functional global network efficiency: A replication study with the human connectome project 1200 data set. NeuroImage, 171, 323331. doi: 10.1016/j.neuroimage.2018.01.018.Google Scholar
Lee, K. H., Choi, Y. Y., Gray, J. R., Cho, S. H., Chae, J.-H., Lee, S., & Kim, K. (2006). Neural correlates of superior intelligence: Stronger recruitment of posterior parietal cortex. NeuroImage, 29(2), 578586. doi: 10.1016/j.neuroimage.2005.07.036.Google Scholar
Lipp, I., Benedek, M., Fink, A., Koschutnig, K., Reishofer, G., Bergner, S., … Neubauer, A. C. (2012). Investigating neural efficiency in the visuo-spatial domain: An FMRI study. PLoS One, 7(12), e51316. doi: 10.1371/journal.pone.0051316.Google Scholar
McKiernan, K. A., Kaufman, J. N., Kucera-Thompson, J., & Binder, J. R. (2003). A parametric manipulation of factors affecting task-induced deactivation in functional neuroimaging. Journal of Cognitive Neuroscience, 15(3), 394408. doi: 10.1162/089892903321593117.Google Scholar
Mennes, M., Biswal, B. B., Castellanos, F. X., & Milham, M. P. (2013). Making data sharing work: The FCP/INDI experience. NeuroImage, 82, 683691. doi: 10.1016/j.neuroimage.2012.10.064.Google Scholar
Miller, D. I., & Halpern, D. F. (2014). The new science of cognitive sex differences. Trends in Cognitive Sciences, 18(1), 3745. doi: 10.1016/j.tics.2013.10.011.Google Scholar
Miller, E. M. (1994). Intelligence and brain myelination: A hypothesis. Personality and Individual Differences, 17(6), 803832. doi: 10.1016/0191-8869(94)90049-3.Google Scholar
Miller, K. L., Alfaro-Almagro, F., Bangerter, N. K., Thomas, D. L., Yacoub, E., Xu, J., … Smith, S. M. (2016). Multimodal population brain imaging in the UK Biobank prospective epidemiological study. Nature Neuroscience, 19(11), 15231536. doi: 10.1038/nn.4393.Google Scholar
Neubauer, A. C., & Fink, A. (2009). Intelligence and neural efficiency. Neuroscience & Biobehavioral Reviews, 33(7), 10041023. doi: 10.1016/j.neubiorev.2009.04.001.Google Scholar
Neubauer, A. C., Fink, A., & Schrausser, D. G. (2002). Intelligence and neural efficiency: The influence of task content and sex on the brain–IQ relationship. Intelligence, 30(6), 515536. doi: 10.1016/S0160–2896(02)00091-0.Google Scholar
Neubauer, A. C., Freudenthaler, H. H., & Pfurtscheller, G. (1995). Intelligence and spatiotemporal patterns of event-related desynchronization (ERD). Intelligence, 20(3), 249266. doi: 10.1016/0160-2896(95)90010-1.Google Scholar
Neubauer, A. C., Grabner, R. H., Fink, A., & Neuper, C. (2005). Intelligence and neural efficiency: Further evidence of the influence of task content and sex on the brain–IQ relationship. Cognitive Brain Research, 25(1), 217225. doi: 10.1016/j.cogbrainres.2005.05.011.Google Scholar
Neubauer, A. C., Grabner, R. H., Freudenthaler, H. H., Beckmann, J. F., & Guthke, J. (2004). Intelligence and individual differences in becoming neurally efficient. Acta Psychologica, 116(1), 5574. doi: 10.1016/j.actpsy.2003.11.005.Google Scholar
Neuper, C., Grabner, R. H., Fink, A., & Neubauer, A. C. (2005). Long-term stability and consistency of EEG event-related (de-)synchronization across different cognitive tasks. Clinical Neurophysiology, 116(7), 16811694. doi: 10.1016/j.clinph.2005.03.013.Google Scholar
Niendam, T. A., Laird, A. R., Ray, K. L., Dean, Y. M., Glahn, D. C., & Carter, C. S. (2012). Meta-analytic evidence for a superordinate cognitive control network subserving diverse executive functions. Cognitive, Affective, & Behavioral Neuroscience, 12(2), 241268. doi: 10.3758/s13415–011-0083-5.Google Scholar
O’Boyle, M. W., Cunnington, R., Silk, T. J., Vaughan, D., Jackson, G., Syngeniotis, A., & Egan, G. F. (2005). Mathematically gifted male adolescents activate a unique brain network during mental rotation. Cognitive Brain Research, 25(2), 583587. doi: 10.1016/j.cogbrainres.2005.08.004.Google Scholar
Parks, R. W., Loewenstein, D. A., Dodrill, K. L., Barker, W. W., Yoshii, F., Chang, J. Y., … Duara, R. (1988). Cerebral metabolic effects of a verbal fluency test: A PET scan study. Journal of Clinical and Experimental Neuropsychology, 10(5), 565575. doi: 10.1080/01688638808402795.Google Scholar
Paul, E. J., Larsen, R. J., Nikolaidis, A., Ward, N., Hillman, C. H., Cohen, N. J., … Barbey, A. K. (2016). Dissociable brain biomarkers of fluid intelligence. Neuroimage, 137, 201211.Google Scholar
Penke, L., Maniega, S. M., Bastin, M. E., Valdés Hernández, M. C., Murray, C., Royle, N. A., … Deary, I. J. (2012). Brain white matter tract integrity as a neural foundation for general intelligence. Molecular Psychiatry, 17(10), 10261030. doi: 10.1038/mp.2012.66.Google Scholar
Pfurtscheller, G., & Aranibar, A. (1977). Event-related cortical desynchronization detected by power measurements of scalp EEG. Electroencephalography and Clinical Neurophysiology, 42(6), 817826. doi: 10.1016/0013-4694(77)90235-8.Google Scholar
Poldrack, R.A. (2015). Is “efficiency” a useful concept in cognitive neuroscience? Developments in Cognitive Neuroscience, 11, 1217.Google Scholar
Prabhakaran, V., Rypma, B., & Gabrieli, J. D. E. (2001). Neural substrates of mathematical reasoning: A functional magnetic resonance imaging study of neocortical activation during performance of the necessary arithmetic operations test. Neuropsychology, 15(1), 115127. doi: 10.1037/0894-4105.15.1.115.Google Scholar
Prabhakaran, V., Smith, J. A. L., Desmond, J. E., Glover, G. H., & Gabrieli, J. D. E. (1997). Neural substrates of fluid reasoning: An fMRI study of neocortical activation during performance of the Raven’s progressive matrices test. Cognitive Psychology, 33(1), 4363. doi: 10.1006/cogp.1997.0659.Google Scholar
Santarnecchi, E., Emmendorfer, A., & Pascual-Leone, A. (2017). Dissecting the parieto-frontal correlates of fluid intelligence: A comprehensive ALE meta-analysis study. Intelligence, 63, 928. doi: 10.1016/j.intell.2017.04.008.Google Scholar
Santarnecchi, E., Emmendorfer, A., Tadayon, S., Rossi, S., Rossi, A., & Pascual-Leone, A. (2017). Network connectivity correlates of variability in fluid intelligence performance. Intelligence, 65, 3547. doi: 10.1016/j.intell.2017.10.002.Google Scholar
Spearman, C. (1904). “General intelligence,” objectively determined and measured. The American Journal of Psychology, 15(2), 201293. doi: 10.2307/1412107.Google Scholar
Sripada, C., Angstadt, M., & Rutherford, S. (2018). Towards a “treadmill test” for cognition: Reliable prediction of intelligence from whole-brain task activation patterns. BioRxiv, 412056. doi: 10.1101/412056.Google Scholar
Sudlow, C., Gallacher, J., Allen, N., Beral, V., Burton, P., Danesh, J., … Collins, R. (2015). UK Biobank: An open access resource for identifying the causes of a wide range of complex diseases of middle and old age. PLoS Medicine, 12(3), e1001779. doi: 10.1371/journal.pmed.1001779.Google Scholar
Takeuchi, H., Taki, Y., Nouchi, R., Yokoyama, R., Kotozaki, Y., Nakagawa, S., … Kawashima, R. (2018). General intelligence is associated with working memory-related brain activity: New evidence from a large sample study. Brain Structure and Function, 223(9), 42434258. doi: 10.1007/s00429–018-1747-5.Google Scholar
Toffanin, P., Johnson, A., de Jong, R., & Martens, S. (2007). Rethinking neural efficiency: Effects of controlling for strategy use. Behavioral Neuroscience, 121(5), 854870. doi: 10.1037/0735-7044.121.5.854.Google Scholar
Turner, B. O., Paul, E. J., Miller, M. B., & Barbey, A. K. (2018). Small sample sizes reduce the replicability of task-based fMRI studies. Communications Biology, 1, 62. doi: 10.1038/s42003-018-0073-z.Google Scholar
van den Heuvel, M. P., Stam, C. J., Kahn, R. S., & Hulshoff Pol, H. E. (2009). Efficiency of functional brain networks and intellectual performance. Journal of Neuroscience, 29(23), 76197624. doi: 10.1523/JNEUROSCI.1443-09.2009.Google Scholar
Van Essen, D. C., Smith, S. M., Barch, D. M., Behrens, T. E. J., Yacoub, E., & Ugurbil, K. (2013). The WU-Minn Human Connectome Project: An overview. NeuroImage, 80(15), 6279. doi: 10.1016/j.neuroimage.2013.05.041.Google Scholar
Woolgar, A., Duncan, J., Manes, F., & Fedorenko, E. (2018). Fluid intelligence is supported by the multiple-demand system not the language system. Nature Human Behaviour, 2(3), 200204. doi: 10.1038/s41562–017-0282-3.Google Scholar
Woolgar, A., Parr, A., Cusack, R., Thompson, R., Nimmo-Smith, I., Torralva, T., … Duncan, J. (2010). Fluid intelligence loss linked to restricted regions of damage within frontal and parietal cortex. Proceedings of the National Academy of Sciences, 107(33), 1489914902. doi: 10.1073/pnas.1007928107.Google Scholar
Yarkoni, T. (2009). Big correlations in little studies: Inflated fMRI correlations reflect low statistical power – Commentary on Vul et al. (2009). Perspectives on Psychological Science, 4(3), 294298. doi: 10.1111/j.1745-6924.2009.01127.x.Google Scholar
Yarkoni, T., Poldrack, R. A., Van Essen, D. C., & Wager, T. D. (2010). Cognitive neuroscience 2.0: Building a cumulative science of human brain function. Trends in Cognitive Sciences, 14(11), 489496. doi: 10.1016/j.tics.2010.08.004.Google Scholar
Yarkoni, T., & Westfall, J. (2017). Choosing prediction over explanation in psychology: Lessons from machine learning. Perspectives on Psychological Science, 12(6), 11001122. doi: 10.1177/1745691617693393.Google Scholar
Yeo, B. T., Krienen, F. M., Sepulcre, J., Sabuncu, M. R., Lashkari, D., Hollinshead, M., … Buckner, R. L. (2011). The organization of the human cerebral cortex estimated by intrinsic functional connectivity. Journal of Neurophysiology, 106(3), 11251165. doi: 10.1152/jn.00338.2011.Google Scholar

References

Alstott, J., Breakspear, M., Hagmann, P., Cammoun, L., & Sporns, O. (2009). Modeling the impact of lesions in the human brain. PLoS Computational Biology, 5(6), e1000408.Google Scholar
Baddeley, A. D., & Hitch, G. (1974). Working memory. In Bower, G. H. (ed.) Psychology of learning and motivation, 8th ed. (pp. 4789). New York: Academic Press.Google Scholar
Barbey, A. K. (2018). Network neuroscience theory of human intelligence. Trends in Cognitive Sciences, 22(1), 820.Google Scholar
Barbey, A. K., Colom, R., Solomon, J., Krueger, F., Forbes, C., & Grafman, J. (2012). An integrative architecture for general intelligence and executive function revealed by lesion mapping. Brain, 135(Pt 4), 11541164.Google Scholar
Bassett, D. S., & Bullmore, E. (2006). Small-world brain networks. The Neuroscientist, 12(6), 512523.Google Scholar
Bassett, D. S., Bullmore, E. T., Meyer-Lindenberg, A., Apud, J. A., Weinberger, D. R., & Coppola, R. (2009). Cognitive fitness of cost-efficient brain functional networks. Proceedings of the National Academy of Sciences of the United States of America, 106(28), 1174711752.Google Scholar
Basten, U., Hilger, K., & Fiebach, C. J. (2015). Where smart brains are different: A quantitative meta-analysis of functional and structural brain imaging studies on intelligence. Intelligence, 51, 1027.Google Scholar
Betzel, R. F., Gu, S., Medaglia, J. D., Pasqualetti, F., & Bassett, D. S. (2016). Optimally controlling the human connectome: The role of network topology. Scientific Reports, 6, 30770.Google Scholar
Bohlken, M. M., Brouwer, R. M., Mandl, R. C. W., Hedman, A. M., van den Heuvel, M. P., van Haren, N. E. M., … Hulshoff Pol, H. E. (2016). Topology of genetic associations between regional gray matter volume and intellectual ability: Evidence for a high capacity network. Neuroimage, 124(Pt A), 10441053.Google Scholar
Breakspear, M. (2017). Dynamic models of large-scale brain activity. Nature Neuroscience, 20(3), 340352.Google Scholar
Bullmore, E., & Sporns, O. (2009). Complex brain networks: Graph theoretical analysis of structural and functional systems. Nature Reviews Neuroscience, 10(3), 186198.Google Scholar
Bullmore, E., & Sporns, O. (2012). The economy of brain network organization. Nature Reviews Neuroscience, 13(5), 336349.Google Scholar
Cabral, J., Kringelbach, M. L., & Deco, G. (2017). Functional connectivity dynamically evolves on multiple time-scales over a static structural connectome: Models and mechanisms. Neuroimage, 160, 8496.Google Scholar
Calhoun, V. D., Miller, R., Pearlson, G., & Adalı, T. (2014). The chronnectome: Time-varying connectivity networks as the next frontier in fMRI data discovery. Neuron, 84(2), 262274.Google Scholar
Cao, M., Wang, Z., & He, Y. (2015). Connectomics in psychiatric research: Advances and applications. Neuropsychiatric Disease and Treatment, 11, 28012810.Google Scholar
Cattell, R. B. (1971). Abilities: Their structure, growth and action. Boston, MA: Houghton Mifflin.Google Scholar
Cattell, R. B., & Horn, J. D. (1978). A check on the theory of fluid and crystallized intelligence with description of new subtest designs. Journal of Educational Measurement, 15(3), 139164.Google Scholar
Chen, T., Cai, W., Ryali, S., Supekar, K., & Menon, V. (2016). Distinct global brain dynamics and spatiotemporal organization of the salience network. PLoS Biology, 14(6), e1002469.Google Scholar
Chen, J. E., Chang, C., Greicius, M. D., & Glover, G. H. (2015). Introducing co-activation pattern metrics to quantify spontaneous brain network dynamics. Neuroimage, 111, 476488.Google Scholar
Chen, Y., Spagna, A., Wu, T., Kim, T. H., Wu, Q., Chen, C., … Fan, J. (2019). Testing a cognitive control model of human intelligence. Scientific Reports, 9, 2898.Google Scholar
Cohen, J. R. (2018). The behavioral and cognitive relevance of time-varying, dynamic changes in functional connectivity. Neuroimage, 180(Pt B), 515525.Google Scholar
Cohen, J. R., & D’Esposito, M. (2016). The segregation and integration of distinct brain networks and their relationship to cognition. The Journal of Neuroscience, 36(48), 1208312094.Google Scholar
Cohen, J. R., Gallen, C. L., Jacobs, E. G., Lee, T. G., & D’Esposito, M. (2014). Quantifying the reconfiguration of intrinsic networks during working memory. PLoS One, 9(9), e106636.Google Scholar
Cole, M. W., Ito, T., Bassett, D. S., & Schultz, D. H. (2016). Activity flow over resting-state networks shapes cognitive task activations. Nature Neuroscience, 19(12), 17181726.Google Scholar
Cole, M. W., Ito, T., & Braver, T. S. (2015). Lateral prefrontal cortex contributes to fluid intelligence through multinetwork connectivity. Brain Connectivity, 5(8), 497504.Google Scholar
Cole, M. W., Laurent, P., & Stocco, A. (2013). Rapid instructed task learning: A new window into the human brain’s unique capacity for flexible cognitive control. Cognitive, Affective & Behavioral Neuroscience, 13(1), 122.Google Scholar
Cole, M. W., Pathak, S., & Schneider, W. (2010). Identifying the brain’s most globally connected regions. Neuroimage, 49(4), 31323148.Google Scholar
Cole, M. W., Yarkoni, T., Repovš, G., Anticevic, A., & Braver, T. S. (2012). Global connectivity of prefrontal cortex predicts cognitive control and intelligence. The Journal of Neuroscience, 32(26), 89888999.Google Scholar
Conway, A. R. A., Getz, S. J., Macnamara, B., & Engel de Abreu, P. M. J. (2011). Working memory and intelligence. In Sternberg, R. J., & Kaufman, S. B. (eds.), The Cambridge handbook of intelligence (pp. 394418). New York: Cambridge University Press.Google Scholar
Conway, A. R. A., Kane, M. J., & Engle, R. W. (2003). Working memory capacity and its relation to general intelligence. Trends in Cognitive Sciences, 7(12), 547552.Google Scholar
Deco, G., & Corbetta, M. (2011). The dynamical balance of the brain at rest. The Neuroscientist, 17(1), 107123.Google Scholar
Deco, G., Jirsa, V. K., & McIntosh, A. R. (2013). Resting brains never rest: Computational insights into potential cognitive architectures. Trends in Neurosciences, 36(5), 268274.Google Scholar
Deco, G., Jirsa, V., McIntosh, A. R., Sporns, O., & Kötter, R. (2009). Key role of coupling, delay, and noise in resting brain fluctuations. Proceedings of the National Academy of Sciences of the United States of America, 106(25), 1030210307.Google Scholar
Deco, G., & Kringelbach, M. L. (2014). Great expectations: Using whole-brain computational connectomics for understanding neuropsychiatric disorders. Neuron, 84(5), 892905.Google Scholar
Deco, G., Tononi, G., Boly, M., & Kringelbach, M. L. (2015). Rethinking segregation and integration: Contributions of whole-brain modelling. Nature Reviews Neuroscience, 16(7), 430439.Google Scholar
Dehaene, S., Kerszberg, M., & Changeux, J. P. (1998). A neuronal model of a global workspace in effortful cognitive tasks. Proceedings of the National Academy of Sciences of the United States of America, 95(24), 1452914534.Google Scholar
Dosenbach, N. U. F., Fair, D. A., Cohen, A. L., Schlaggar, B. L., & Petersen, S. E. (2008). A dual-networks architecture of top-down control. Trends in Cognitive Sciences, 12(3), 99105.Google Scholar
Dubin, M. (2017). Imaging TMS: Antidepressant mechanisms and treatment optimization. International Review of Psychiatry, 29(2), 8997.Google Scholar
Dubois, J., Galdi, P., Paul, L. K., & Adolphs, R. (2018). A distributed brain network predicts general intelligence from resting-state human neuroimaging data. Philosophical Transactions of the Royal Society B: Biological Sciences, 373, 20170284.Google Scholar
Duncan, J. (2001). An adaptive coding model of neural function in prefrontal cortex. Nature Reviews Neuroscience, 2(11), 820829.Google Scholar
Duncan, J. (2010). The multiple-demand (MD) system of the primate brain: Mental programs for intelligent behaviour. Trends in Cognitive Sciences, 14(4), 172179.Google Scholar
Ekman, M., Derrfuss, J., Tittgemeyer, M., & Fiebach, C. J. (2012). Predicting errors from reconfiguration patterns in human brain networks. Proceedings of the National Academy of Sciences of the United States of America, 109(41), 1671416719.Google Scholar
Elton, A., & Gao, W. (2014). Divergent task-dependent functional connectivity of executive control and salience networks. Cortex, 51, 5666.Google Scholar
Elton, A., & Gao, W. (2015). Task-related modulation of functional connectivity variability and its behavioral correlations. Human Brain Mapping, 36(8), 32603272.Google Scholar
Euler, M. J. (2018). Intelligence and uncertainty: Implications of hierarchical predictive processing for the neuroscience of cognitive ability. Neuroscience and Biobehavioral Reviews, 94, 93112.Google Scholar
Finc, K., Bonna, K., Lewandowska, M., Wolak, T., Nikadon, J., Dreszer, J., … Kühn, S. (2017). Transition of the functional brain network related to increasing cognitive demands. Human Brain Mapping, 38(7), 36593674.Google Scholar
Finn, E. S., Shen, X., Scheinost, D., Rosenberg, M. D., Huang, J., Chun, M. M., … Constable, R. T. (2015). Functional connectome fingerprinting: Identifying individuals using patterns of brain connectivity. Nature Neuroscience, 18(11), 16641671.Google Scholar
Fornito, A., Zalesky, A., & Breakspear, M. (2015). The connectomics of brain disorders. Nature Reviews Neuroscience, 16(3), 159172.Google Scholar
Fox, M. D., Buckner, R. L., Liu, H., Chakravarty, M. M., Lozano, A. M., & Pascual-Leone, A. (2014). Resting-state networks link invasive and noninvasive brain stimulation across diverse psychiatric and neurological diseases. Proceedings of the National Academy of Sciences of the United States of America, 111(41), E4367E4375.Google Scholar
Friedman, N. P., & Miyake, A. (2017). Unity and diversity of executive functions: Individual differences as a window on cognitive structure. Cortex, 86, 186204.Google Scholar
Friedman, N. P., Miyake, A., Corley, R. P., Young, S. E., Defries, J. C., & Hewitt, J. K. (2006). Not all executive functions are related to intelligence. Psychological Science, 17(2), 172179.Google Scholar
Gallen, C. L., & D’Esposito, M. (2019). Modular brain network organization: A biomarker of cognitive plasticity. Trends in Cognitive Sciences, 23(4), 293304.Google Scholar
Gallen, C. L., Turner, G. R., Adnan, A., & D’Esposito, M. (2016). Reconfiguration of brain network architecture to support executive control in aging. Neurobiology of Aging, 44, 4252.Google Scholar
Garlick, D. (2002). Understanding the nature of the general factor of intelligence: The role of individual differences in neural plasticity as an explanatory mechanism. Psychological Review, 109(1), 116136.Google Scholar
Girn, M., Mills, C., & Christoff, K. (2019). Linking brain network reconfiguration and intelligence: Are we there yet? Trends in Neuroscience and Education, 15, 6270.Google Scholar
Gläscher, J., Rudrauf, D., Colom, R., Paul, L. K., Tranel, D., Damasio, H., & Adolphs, R. (2010). Distributed neural system for general intelligence revealed by lesion mapping. Proceedings of the National Academy of Sciences of the United States of America, 107(10), 47054709.Google Scholar
Godwin, D., Barry, R. L., & Marois, R. (2015). Breakdown of the brain’s functional network modularity with awareness. Proceedings of the National Academy of Sciences of the United States of America, 112(12), 37993804.Google Scholar
Gonzalez-Castillo, J., & Bandettini, P. A. (2018). Task-based dynamic functional connectivity: Recent findings and open questions. Neuroimage, 180(Pt B), 526533.Google Scholar
Goodkind, M., Eickhoff, S. B., Oathes, D. J., Jiang, Y., Chang, A., Jones-Hagata, L. B., … Etkin, A. (2015). Identification of a common neurobiological substrate for mental illness. JAMA Psychiatry, 72(4), 305315.Google Scholar
Gordon, E. M., Stollstorff, M., & Vaidya, C. J. (2012). Using spatial multiple regression to identify intrinsic connectivity networks involved in working memory performance. Human Brain Mapping, 33(7), 15361552.Google Scholar
Goschke, T. (2014). Dysfunctions of decision-making and cognitive control as transdiagnostic mechanisms of mental disorders: Advances, gaps, and needs in current research. International Journal of Methods in Psychiatric Research, 23(Suppl 1), 4157.Google Scholar
Gratton, C., Laumann, T. O., Nielsen, A. N., Greene, D. J., Gordon, E. M., Gilmore, A. W., … Petersen, S. E. (2018). Functional brain networks are dominated by stable group and individual factors, not cognitive or daily variation. Neuron, 98(2), 439452.e5.Google Scholar
Gratton, C., Lee, T. G., Nomura, E. M., & D’Esposito, M. (2013). The effect of theta-burst TMS on cognitive control networks measured with resting state fMRI. Frontiers in Systems Neuroscience, 7, 124.Google Scholar
Gratton, C., Nomura, E. M., Perez, F., & D’Esposito, M. (2012). Focal brain lesions to critical locations cause widespread disruption of the modular organization of the brain. Journal of Cognitive Neuroscience, 24(6), 12751285.Google Scholar
Gratton, C., Sun, H., & Petersen, S. E. (2018). Control networks and hubs. Psychophysiology, 55(3), e13032.Google Scholar
Greene, A. S., Gao, S., Scheinost, D., & Constable, R. T. (2018). Task-induced brain state manipulation improves prediction of individual traits. Nature Communications, 9(1), 2807.Google Scholar
Gu, S., Pasqualetti, F., Cieslak, M., Telesford, Q. K., Yu, A. B., Kahn, A. E., … Bassett, D. S. (2015). Controllability of structural brain networks. Nature Communications, 6, 8414.Google Scholar
Guimerà, R., Mossa, S., Turtschi, A., & Amaral, L. A. N. (2005). The worldwide air transportation network: Anomalous centrality, community structure, and cities’ global roles. Proceedings of the National Academy of Sciences of the United States of America, 102(22), 77947799.Google Scholar
Haier, R. J., Siegel, B. V., Nuechterlein, K. H., Hazlett, E., Wu, J. C., Paek, J., … Buchsbaum, M. S. (1988). Cortical glucose metabolic rate correlates of abstract reasoning and attention studied with positron emission tomography. Intelligence, 12, 199217.Google Scholar
Hart, M. G., Ypma, R. J. F., Romero-Garcia, R., Price, S. J., & Suckling, J. (2016). Graph theory analysis of complex brain networks: New concepts in brain mapping applied to neurosurgery. Journal of Neurosurgery, 124(6), 16651678.Google Scholar
Hearne, L. J., Mattingley, J. B., & Cocchi, L. (2016). Functional brain networks related to individual differences in human intelligence at rest. Scientific Reports, 6, 32328.Google Scholar
Hilger, K., Ekman, M., Fiebach, C. J., & Basten, U. (2017a). Efficient hubs in the intelligent brain: Nodal efficiency of hub regions in the salience network is associated with general intelligence. Intelligence, 60, 1025.Google Scholar
Hilger, K., Ekman, M., Fiebach, C. J., & Basten, U. (2017b). Intelligence is associated with the modular structure of intrinsic brain networks. Scientific Reports, 7(1), 16088.Google Scholar
Honey, C. J., Kötter, R., Breakspear, M., & Sporns, O. (2007). Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proceedings of the National Academy of Sciences of the United States of America, 104(24), 1024010245.Google Scholar
Hutchison, R. M., & Morton, J. B. (2015). Tracking the brain’s functional coupling dynamics over development. The Journal of Neuroscience, 35(17), 68496859.Google Scholar
Jung, R. E., & Haier, R. J. (2007). The Parieto-Frontal Integration Theory (P-FIT) of intelligence: Converging neuroimaging evidence. Behavioral and Brain Sciences, 30(2), 135154, discussion 154–187.Google Scholar
Kane, M. J., & Engle, R. W. (2002). The role of prefrontal cortex in working-memory capacity, executive attention, and general fluid intelligence: An individual-differences perspective. Psychonomic Bulletin & Review, 9(4), 637671.Google Scholar
Kenett, Y. N., Medaglia, J. D., Beaty, R. E., Chen, Q., Betzel, R. F., Thompson-Schill, S. L., & Qiu, J. (2018). Driving the brain towards creativity and intelligence: A network control theory analysis. Neuropsychologia, 118(Pt A), 7990.Google Scholar
Kitzbichler, M. G., Henson, R. N. A., Smith, M. L., Nathan, P. J., & Bullmore, E. T. (2011). Cognitive effort drives workspace configuration of human brain functional networks. The Journal of Neuroscience, 31(22), 82598270.Google Scholar
Kovacs, K., & Conway, A. R. A. (2016). Process overlap theory: A unified account of the general factor of intelligence. Psychological Inquiry, 27(3), 151177.Google Scholar
Kucyi, A., Tambini, A., Sadaghiani, S., Keilholz, S., & Cohen, J. R. (2018). Spontaneous cognitive processes and the behavioral validation of time-varying brain connectivity. Network Neuroscience, 2(4), 397417.Google Scholar
Langer, N., Pedroni, A., Gianotti, L. R. R., Hänggi, J., Knoch, D., & Jäncke, L. (2012). Functional brain network efficiency predicts intelligence. Human Brain Mapping, 33(6), 13931406.Google Scholar
Li, Y., Liu, Y., Li, J., Qin, W., Li, K., Yu, C., & Jiang, T. (2009). Brain anatomical network and intelligence. PLoS Computational Biology, 5(5), e1000395.Google Scholar
Liang, X., Zou, Q., He, Y., & Yang, Y. (2016). Topologically reorganized connectivity architecture of default-mode, executive-control, and salience networks across working memory task loads. Cerebral Cortex, 26(4), 15011511.Google Scholar
Liu, H., Yu, H., Li, Y., Qin, W., Xu, L., Yu, C., & Liang, M. (2017). An energy-efficient intrinsic functional organization of human working memory: A resting-state functional connectivity study. Behavioural Brain Research, 316, 6673.Google Scholar
Malpas, C. B., Genc, S., Saling, M. M., Velakoulis, D., Desmond, P. M., & O’Brien, T. J. (2016). MRI correlates of general intelligence in neurotypical adults. Journal of Clinical Neuroscience, 24, 128134.Google Scholar
McTeague, L. M., Goodkind, M. S., & Etkin, A. (2016). Transdiagnostic impairment of cognitive control in mental illness. Journal of Psychiatric Research, 83, 3746.Google Scholar
McTeague, L. M., Huemer, J., Carreon, D. M., Jiang, Y., Eickhoff, S. B., & Etkin, A. (2017). Identification of common neural circuit disruptions in cognitive control across psychiatric disorders. American Journal of Psychiatry, 174(7), 676685.Google Scholar
Mercado, E. III. (2008). Neural and cognitive plasticity: From maps to minds. Psychological Bulletin, 134(1), 109137.Google Scholar
Mesulam, M.-M. (1990). Large-scale neurocognitive networks and distributed processing for attention, language, and memory. Annals of Neurology, 28(5), 597613.Google Scholar
Meunier, D., Lambiotte, R., & Bullmore, E. T. (2010). Modular and hierarchically modular organization of brain networks. Frontiers in Neuroscience, 4, 200.Google Scholar
Mill, R. D., Ito, T., & Cole, M. W. (2017). From connectome to cognition: The search for mechanism in human functional brain networks. Neuroimage, 160, 124139.Google Scholar
Miller, E. K., & Cohen, J. D. (2001). An integrative theory of prefrontal cortex function. Annual Review of Neuroscience, 24, 167202.Google Scholar
Mitra, A., Snyder, A. Z., Blazey, T., & Raichle, M. E. (2015). Lag threads organize the brain’s intrinsic activity. Proceedings of the National Academy of Sciences of the United States of America, 112(17), E2235E2244.Google Scholar
Newman, M. E. J., & Girvan, M. (2004). Finding and evaluating community structure in networks. Physical Review E, 69(2), 026113.Google Scholar
O’Reilly, R. C., Herd, S. A., & Pauli, W. M. (2010). Computational models of cognitive control. Current Opinion in Neurobiology, 20(2), 257261.Google Scholar
Opitz, A., Fox, M. D., Craddock, R. C., Colcombe, S., & Milham, M. P. (2016). An integrated framework for targeting functional networks via transcranial magnetic stimulation. Neuroimage, 127, 8696.Google Scholar
Power, J. D., & Petersen, S. E. (2013). Control-related systems in the human brain. Current Opinion in Neurobiology, 23(2), 223228.Google Scholar
Raven, J. (2000). The Raven’s progressive matrices: Change and stability over culture and time. Cognitive Psychology, 41(1), 148.Google Scholar
Sadaghiani, S., Poline, J. B., Kleinschmidt, A., & D’Esposito, M. (2015). Ongoing dynamics in large-scale functional connectivity predict perception. Proceedings of the National Academy of Sciences of the United States of America, 112(27), 84638468.Google Scholar
Santarnecchi, E., Emmendorfer, A., Tadayon, S., Rossi, S., Rossi, A., Pascual-Leone, A., & Honeywell SHARP Team Authors. (2017). Network connectivity correlates of variability in fluid intelligence performance. Intelligence, 65, 3547.Google Scholar
Schultz, D. H., & Cole, M. W. (2016). Higher intelligence is associated with less task-related brain network reconfiguration. The Journal of Neuroscience, 36(33), 85518561.Google Scholar
Shanmugan, S., Wolf, D. H., Calkins, M. E., Moore, T. M., Ruparel, K., Hopson, R. D., … Satterthwaite, T. D. (2016). Common and dissociable mechanisms of executive system dysfunction across psychiatric disorders in youth. American Journal of Psychiatry, 173(5), 517526.Google Scholar
Shine, J. M., Bissett, P. G., Bell, P. T., Koyejo, O., Balsters, J. H., Gorgolewski, K. J., … Poldrack, R. A. (2016). The dynamics of functional brain networks: Integrated network states during cognitive task performance. Neuron, 92(2), 544554.Google Scholar
Shine, J. M., & Poldrack, R. A. (2018). Principles of dynamic network reconfiguration across diverse brain states. Neuroimage, 180(Pt B), 396405.Google Scholar
Snyder, H. R., Miyake, A., & Hankin, B. L. (2015). Advancing understanding of executive function impairments and psychopathology: Bridging the gap between clinical and cognitive approaches. Frontiers in Psychology, 6, 328.Google Scholar
Song, M., Zhou, Y., Li, J., Liu, Y., Tian, L., Yu, C., & Jiang, T. (2008). Brain spontaneous functional connectivity and intelligence. Neuroimage, 41(3), 11681176.Google Scholar
Spadone, S., Della Penna, S., Sestieri, C., Betti, V., Tosoni, A., Perrucci, M. G., … Corbetta, M. (2015). Dynamic reorganization of human resting-state networks during visuospatial attention. Proceedings of the National Academy of Sciences of the United States of America, 112(26), 81128117.Google Scholar
Sporns, O. (2010). Networks of the brain. Cambridge, MA: MIT Press.Google Scholar
Sporns, O. (2013). Network attributes for segregation and integration in the human brain. Current Opinion in Neurobiology, 23(2), 162171.Google Scholar
Stanley, M. L., Dagenbach, D., Lyday, R. G., Burdette, J. H., & Laurienti, P. J. (2014). Changes in global and regional modularity associated with increasing working memory load. Frontiers in Human Neuroscience, 8, 954.Google Scholar
Sternberg, R. J., & Kaufman, S. B. (eds.) (2011). The Cambridge handbook of intelligence. New York: Cambridge University Press.Google Scholar
Thompson, G. J., Magnuson, M. E., Merritt, M. D., Schwarb, H., Pan, W.-J., McKinley, A., … Keilholz, S. D. (2013). Short-time windows of correlation between large-scale functional brain networks predict vigilance intraindividually and interindividually. Human Brain Mapping, 34(12), 32803298.Google Scholar
van den Heuvel, M. P., & Sporns, O. (2011). Rich-club organization of the human connectome. The Journal of Neuroscience, 31(44), 1577515786.Google Scholar
van den Heuvel, M. P., Stam, C. J., Kahn, R. S., & Hulshoff Pol, H. E. (2009). Efficiency of functional brain networks and intellectual performance. The Journal of Neuroscience, 29(23), 76197624.Google Scholar
Vatansever, D., Menon, D. K., Manktelow, A. E., Sahakian, B. J., & Stamatakis, E. A. (2015). Default mode dynamics for global functional integration. The Journal of Neuroscience, 35(46), 1525415262.Google Scholar
Wang, C., Ong, J. L., Patanaik, A., Zhou, J., & Chee, M. W. L. (2016). Spontaneous eyelid closures link vigilance fluctuation with fMRI dynamic connectivity states. Proceedings of the National Academy of Sciences of the United States of America, 113(34), 96539658.Google Scholar
Wang, L., Song, M., Jiang, T., Zhang, Y., & Yu, C. (2011). Regional homogeneity of the resting-state brain activity correlates with individual intelligence. Neuroscience Letters, 488(3), 275278.Google Scholar
Wechsler, D. (2008). Wechsler Adult Intelligence Scale – Fourth edition (WAIS-IV). San Antonio, TX: Pearson.Google Scholar
Wechsler, D. (2011). Wechsler Abbreviated Scale of Intelligence – Second edition (WASI-II). San Antonio, TX: Pearson.Google Scholar
Xia, M., & He, Y. (2011). Magnetic resonance imaging and graph theoretical analysis of complex brain networks in neuropsychiatric disorders. Brain Connectivity, 1(5), 349365.Google Scholar
Xiao, L., Stephen, J. M., Wilson, T. W., Calhoun, V. D., & Wang, Y. (2019). Alternating diffusion map based fusion of multimodal brain connectivity networks for IQ prediction. IEEE Transactions on Biomedical Engineering, 68(8), 21402151.Google Scholar
Yin, S., Wang, T., Pan, W., Liu, Y., & Chen, A. (2015). Task-switching cost and intrinsic functional connectivity in the human brain: Toward understanding individual differences in cognitive flexibility. PLoS One, 10(12), e0145826.Google Scholar
Zippo, A. G., Della Rosa, P. A., Castiglioni, I., & Biella, G. E. M. (2018). Alternating dynamics of segregation and integration in human EEG functional networks during working-memory task. Neuroscience, 371, 191206.Google Scholar

References

Aboitiz, F., Scheibel, A. B., Fisher, R. S., & Zaidel, E. (1992). Fiber composition of the human corpus callosum. Brain Research, 598(1–2), 143153.Google Scholar
Andreasen, N. C., Flaum, M., Swayze, V. D., O’Leary, D. S., Alliger, R., Cohen, G., … Yuh, W. T. (1993). Intelligence and brain structure in normal individuals. American Journal of Psychiatry, 150(1), 130134.Google Scholar
Anon., (2002). WAIS-III WMS-III technical manual. New York: The Psychological Corporation.Google Scholar
Aydin, K., Uysal, S., Yakut, A., Emiroglu, B., & Yilmaz, F. (2012). N-Acetylaspartate concentration in corpus callosum is positively correlated with intelligence in adolescents. NeuroImage, 59(2), 10581064.Google Scholar
Basten, U., Hilger, K., & Fiebach, C. J. (2015). Where smart brains are different: A quantitative meta-analysis of functional and structural brain imaging studies on intelligence. Intelligence, 51(1), 1027.Google Scholar
Bates, T. E., Strangward, M., Keelan, J., Davey, G. P., Munro, P. M. G. G., & Clark, J. B. 1996. Inhibition of N-acetylaspartate production: Implications for 1H MRS studies in vivo. Neuroreport, 7(8), 13971400.Google Scholar
Blakely, R. D., & Coyle, J. T. (1988). The neurobiology of N-acetylasparty. International Review of Neurobiology, 30, 39100.Google Scholar
Brooks, W. M., Friedman, S. D., & Gasparovic, C. (2001). Magnetic resonance spectroscopy in traumatic brain injury. Journal of Head Trauma Rehabilitation, 16(2), 149164.Google Scholar
Button, K. S., Ioannidis, J. P. A., Mokrysz, C., Nosek, B. A. Flint, J., Robinson, E. S. J., & Munafò, M. R. (2013). Power failure: Why small sample size undermines the reliability of neuroscience. Nature Reviews Neuroscience, 14, 365376.Google Scholar
Cabeza, R., & Nyberg, L. (2000). Imaging cognition II: An empirical review of 275 PET and FMRI studies. Journal of Cognitive Neuroscience, 12(1), 147.Google Scholar
Callaway, E. (1973). Correlations between averaged evoked potentials and measures of intelligence: An overview. Archives of General Psychiatry, 29(4), 553558.Google Scholar
Charlton, R. A., McIntyre, D. J. O. O., Howe, F. A., Morris, R. G., & Markus, H. S. (2007). The relationship between white matter brain metabolites and cognition in normal aging: The GENIE study. Brain Research, 1 164, 108116.Google Scholar
D’Adamo, A. F., & Yatsu, F. M. (1966). Acetate metabolism in the nervous system. N-acetyl-l-aspartic acid and the biosynthesis of brain lipids. Journal of Neurochemistry, 13(10), 961965.Google Scholar
Dubois, J., & Adolphs, R. (2016). Building a science of individual differences from FMRI. Trends in Cognitive Sciences, 20(6), 425443.Google Scholar
Ellis, F. R. (1969). Some effects of PCO2 and PH on nerve tissue. British Journal of Pharmacology, 35(1), 197201.Google Scholar
Ertl, J. P., & Schafer, E. W. P. (1969). Brain response correlates of psychometric intelligence. Nature, 223, 421422.Google Scholar
Ferguson, K. J., MacLullich, A. M. J., Marshall, I., Deary, I. J., Starr, J. M., Seckl, J. R., & Wardlaw, J. M. (2002). Magnetic resonance spectroscopy and cognitive function in healthy elderly men. Brain, 125(Pt. 12), 27432749.Google Scholar
Flynn, J. R. (1987). Massive IQ gains in 14 nations: What IQ tests really measure. Psychological Bulletin, 101(2), 171191.Google Scholar
Friedman, S. D., Brooks, W. M., Jung, R. E., Blaine, B. L. L., Hart, L., & Yeo, R. A. (1998). Proton MR spectroscopic findings correspond to neuropsychological function in traumatic brain injury. American Journal of Neuroradiology, 19(10), 18791885.Google Scholar
Gadian, D. G. (1995). NMR and its applications to living systems. Oxford University Press.Google Scholar
Giménez, M., Junqué, C., Narberhaus, A., Caldú, X., Segarra, D., Vendrell, P., … Mercader, J. M. (2004). Medial temporal MR spectroscopy is related to memory performance in normal adolescent subjects. Neuroreport, 15(4), 703707.Google Scholar
Glasser, M. F., Coalson, T. S., Robinson, E. C., Hacker, C. D., Harwell, J., Yacoub, E., … Van Essen, D. C. (2016). A multi-modal parcellation of human cerebral cortex. Nature, 536, 171178.Google Scholar
Graff-Radford, J., & Kantarci, K. (2013). Magnetic resonance spectroscopy in Alzheimer’s disease. Neuropsychiatric Disease and Treatment, 9, 687–696.Google Scholar
Grazioplene, R. G., Rachael, G., Ryman, S. G., Gray, J. R., Rustichini, A., Jung, R. E., & DeYoung, C. G. (2015). Subcortical intelligence: Caudate volume predicts IQ in healthy adults. Human Brain Mapping, 36(4), 14071416.Google Scholar
Gur, R. C., Turetsky, B. I., Matsui, M., Yan, M., Bilker, W., Hughett, P., & Gur, R. E. (1999). Sex differences in brain gray and white matter in healthy young adults: Correlations with cognitive performance. Journal of Neuroscience, 19(10), 40654072.Google Scholar
Haász, J., Westlye, E. T., Fjær, S., Espeseth, T., Lundervold, A., & Lundervold, A. J. (2013). General fluid-type intelligence is related to indices of white matter structure in middle-aged and old adults. NeuroImage, 83, 372383.Google Scholar
Harvey, I., Persaud, R., Ron, M. A., Baker, G., & Murray, R. M. (1994). Volumetric MRI measurements in bipolars compared with schizophrenics and healthy controls. Psychological Medicine, 24(3), 689699.Google Scholar
Hashimoto, T., Tayama, M., Miyazaki, M., Yoneda, Y., Yoshimoto, T., Harada, M., … Kuroda, Y. (1995). Reduced N-acetylaspartate in the brain observed on in vivo proton magnetic resonance spectroscopy in patients with mental retardation. Pediatric Neurology, 13(3), 205208.Google Scholar
Jensen, A. R. (1982). Reaction time and psychometric g. In Eysenck, H. J. (ed.), A model for intelligence (pp. 93132). Berlin: Springer-Verlag.Google Scholar
Jung, R. E., Brooks, W. M., Yeo, R. A., Chiulli, S. J., Weers, D. C., & Sibbitt, W. L. Jr (1999). Biochemical markers of intelligence: A proton MR spectroscopy study of normal human brain. Proceedings of the Royal Society B-Biological Sciences, 266(1426), 13751379.Google Scholar
Jung, R. E., Gasparovic, C., Robert, R. S., Chavez, S., Caprihan, A., Barrow, R., & Yeo, R. A. (2009). Imaging intelligence with proton magnetic resonance spectroscopy. Intelligence, 37(2), 192198.Google Scholar
Jung, R. E., & Haier, R. J. (2007). The Parieto-Frontal Integration Theory (P-FIT) of intelligence: Converging neuroimaging evidence. Behavioral and Brain Sciences, 30(2):135154.Google Scholar
Jung, R. E., Haier, R. J., Yeo, R. A., Rowland, L. M., Petropoulos, H., Levine, A. S., … Brooks, W. M. (2005). Sex differences in N-acetylaspartate correlates of general intelligence: An H-1-MRS study of normal human brain. Neuroimage, 26(3), 965972.Google Scholar
Jung, R. E., Yeo, R. A., Sibbitt, W. L. Jr., Ford, C. C., Hart, B. L., & Brooks, W. M. (2001). Gerstmann syndrome in systemic lupus erythematosus: Neuropsychological, neuroimaging and dpectroscopic findings. Neurocase, 7(6), 515521.Google Scholar
Kocevar, G., Suprano, I., Stamile, C., Hannoun, S., Fourneret, P., Revol, O., … Sappey-Marinier, D. (2019). Brain structural connectivity correlates with fluid intelligence in children: A DTI graph analysis. Intelligence, 72, 6775.Google Scholar
Kumar, V., Sharma, U., & Jagannathan, N. R. (2012). In vivo magnetic resonance spectroscopy of cancer. Biomedical Spectroscopy and Imaging, 1(1), 89100.Google Scholar
Lehmann, J. E. (1937). The effect of changes in PH on the action of mammalian A nerve fibres. American Journal of Physiology, 118(3), 600612.Google Scholar
Lezak, M. D., Howieson, D. B., & Loring, D. W. (2004). Neuropsychological assessment, 4th ed. New York: Oxford University Press.Google Scholar
López-Villegas, D., Lenkinski, R. E., & Frank, I. (1997). Biochemical changes in the frontal lobe of HIV-infected individuals detected by magnetic resonance spectroscopy. Proceedings of the National Academy of Sciences of the United States of America, 94(18), 98549859.Google Scholar
Martin, P. R., Gibbs, S. J., Nimmerrichter, A. A., Riddle, W. R., Welch, L. W., & Willcott, M. R. (1995). Brain proton magnetic resonance spectroscopy studies in recently abstinent alcoholics. Alcoholism: Clinical and Experimental Research, 19(4), 10781082.Google Scholar
Moffett, J. R., Ross, B. D., Arun, P., Madhavarao, C. N., & Namboodiri, Aryan. (2007). N-Acetylaspartate in the CNS: From neurodiagnostics to neurobiology. Progress in Neurobiology, 81(2), 89131.Google Scholar
Navas-Sánchez, F. J., Alemán-Gómez, Y., Sánchez-Gonzalez, J., Guzmán-De-Villoria, J. A., Franco, C., Robles, O., … Desco, M. (2014). White matter microstructure correlates of mathematical giftedness and intelligence quotient. Human Brain Mapping, 35(6), 26192631.Google Scholar
Nikolaidis, A., Baniqued, P. L., Kranz, M. B., Scavuzzo, C. J., Barbey, A. K., Kramer, A. F., & Larsen, R. J. (2017). Multivariate associations of fluid intelligence and NAA. Cerebral Cortex, 27(4), 26072616.Google Scholar
Nordengen, K., Heuser, C., Rinholm, J. E., Matalon, R., & Gundersen, V. (2015). Localisation of N-acetylaspartate in oligodendrocytes/myelin. Brain Structure and Function, 220(2), 899917.Google Scholar
Nusbaum, F., Hannoun, S., Kocevar, G., Stamile, C., Fourneret, P., Revol, O., & Sappey-Marinier, D. (2017). Hemispheric differences in white matter microstructure between two profiles of children with high intelligence quotient vs. controls: A tract-based spatial statistics study. Frontiers in Neuroscience, 11, 173. doi: 10.3389/fnins.2017.00173.eCollection2017.Google Scholar
Parrish, R. G., Kurland, R. J., Janese, W. W., & Bakay, L. (1974). Proton relaxation rates of water in brain and brain tumors. Science, 183(4123), 438439.Google Scholar
Patel, T., & Talcott, J. B. (2014). Moderate relationships between NAA and cognitive ability in healthy adults: Implications for cognitive spectroscopy. Frontiers in Human Neuroscience, 8, 39. doi: 10.3389/fnhum.2014.00039.eCollection2014.Google Scholar
Paul, E. J., Larsen, R. J., Nikolaidis, A., Ward, N., Hillman, C. H., Cohen, N. J., … Barbey, A. K. (2016). Dissociable brain biomarkers of fluid intelligence. NeuroImage, 137, 201211.Google Scholar
Pfleiderer, B., Ohrmann, P., Suslow, T., Wolgast, M., Gerlach, A. L., Heindel, W., & Michael, N. (2004). N-Acetylaspartate levels of left frontal cortex are associated with verbal intelligence in women but not in men: A proton magnetic resonance spectroscopy study. Neuroscience, 123(4), 10531058.Google Scholar
Pietschnig, J., Penke, L., Wicherts, J. M., Zeiler, M., & Voracek, M. (2015). Meta-analysis of associations between human brain volume and intelligence differences: How strong are they and what do they mean? Neuroscience and Biobehavioral Reviews, 57, 411432.Google Scholar
Posner, M. I., & Raichle, M. E. (1998). The neuroimaging of human brain function. Proceedings of the National Academy of Sciences of the United States of America, 95(3), 763764.Google Scholar
Rae, C., Scott, R. B., Thompson, C. H., Kemp, G. J., Dumughn, I., Styles, P., … Radda, G. K. (1996). Is PH a biochemical marker of IQ? Proceedings of the Royal Society B: Biological Sciences, 263(1373), 10611064.Google Scholar
Rajanayagam, V., Balthazor, M., Shapiro, E. G., Krivit, W., Lockman, L., & Stillman, A. E. (1997). Proton MR spectroscopy and neuropsychological testing in adrenoleukodystrophy. American Journal of Neuroradiology, 18(10), 19091914.Google Scholar
Reiss, A. L., Abrams, M. T., Singer, H. S., Ross, J. L., & Denckla, M. B. (1996). Brain development, gender and IQ in children. A volumetric imaging study. Brain, 119(Pt 5), 17631774. doi: 10.1093/brain/119.5.1763.Google Scholar
Ross, A. J., & Sachdev, P. S. (2004). Magnetic resonance spectroscopy in cognitive research. Brain Research Reviews, 44(2–3), 83102.Google Scholar
Schuff, N., Ezekiel, F., Gamst, A. C., Amend, D. L., Capizzano, A. A., Maudsley, A. A., & Weiner, M. W. (2001). Region and tissue differences of metabolites in normally aged brain using multislice 1H magnetic resonance spectroscopic imaging. Magnetic Resonance in Medicine, 45(5), 899907.Google Scholar
Sibbitt, W. L. Jr., Sibbitt, R. R., & Brooks, W. M. (1999). Neuroimaging in neuropsychiatric systemic lupus erythematosus. Arthritis & Rheumatism, 42(10), 20262038.Google Scholar
Soher, B. J., van Zijl, P. C., Duyn, J. H., & Barker, P. B. (1996). Quantitative proton MR spectroscopic imaging of the human brain. Magnetic Resonance in Medicine: Official Journal of the Society of Magnetic Resonance in Medicine/Society of Magnetic Resonance in Medicine, 35(3), 356363.Google Scholar
Taylor, D. L., Davies, S. E. C., Obrenovitch, T. P., Doheny, M. H., Patsalos, P. N., Clark, J. B., & Symon, L. (2002). Investigation into the role of N-acetylaspartate in cerebral osmoregulation. Journal of Neurochemistry, 65(1), 275281.Google Scholar
Tedeschi, G., Bertolino, A., Righini, A., Campbell, G., Raman, R., Duyn, J. H., … Di Chiro, G. (1995). Brain regional distribution pattern of metabolite signal intensities in young adults by proton magnetic resonance spectroscopic imaging. Neurology, 45(7), 13841391.Google Scholar
Vilasboas, T., Herbet, G., & Duffau, H. (2017). Challenging the myth of right nondominant hemisphere: Lessons from corticosubcortical stimulation mapping in awake surgery and surgical implications. World Neurosurgery, 103, 449456.Google Scholar
Wickett, J. C., Vernon, P. A., & Lee, D. H. (1994). In vivo brain size, head perimeter, and intelligence in a sample of healthy adult females. Personality and Individual Differences, 16(6), 831838.Google Scholar
Willerman, L., Schultz, R., Rutledge, J. N., & Bigler, E. D. (1991). In vivo brain size and intelligence. Intelligence, 15(2), 223228.Google Scholar

References

An, L., Araneta, M. F., Johnson, C., & Shen, J. (2018). Simultaneous measurement of glutamate, glutamine, GABA, and glutathione by spectral editing without subtraction. Magnetic Resonance in Medicine, 80(5), 17761717. doi: 10.1002/mrm.27172.Google Scholar
Andres, R. H., Ducray, A. D., Schlattner, U., Wallimann, T., & Widmer, H. R. (2008). Functions and effects of creatine in the central nervous system. Brain Research Bulletin, 76(4), 329343. doi: 10.1016/j.brainresbull.2008.02.035Google Scholar
Apsvalka, D., Gadie, A., Clemence, M., & Mullins, P. G. (2015). Event-related dynamics of glutamate and BOLD effects measured using functional magnetic resonance spectroscopy (fMRS) at 3T in a repetition suppression paradigm. Neuroimage, 118, 292300.Google Scholar
Attwell, D., & Laughlin, S. B. (2001). An energy budget for signaling in the grey matter of the brain. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism, 21(10), 11331145. doi: 10.1097/00004647-200110000-00001.Google Scholar
Bäckman, L., Lindenberger, U., Li, S. C., & Nyberg, L. (2010). Linking cognitive aging to alterations in dopamine neurotransmitter functioning: Recent data and future avenues. Neuroscience and Biobehavioral Reviews, 34(5), 670677. doi: 10.1016/j.neubiorev.2009.12.008Google Scholar
Barreto, F. R., Costa, T. B., Landim, R. C., Castellano, G., & Salmon, C. E. (2014). (31)P-MRS using visual stimulation protocols with different durations in healthy young adult subjects. Neurochemical Research, 39(12), 23432350. doi: 10.1007/s11064-014-1433-9.Google Scholar
Bartha, R., Drost, D. J., Menon, R. S., & Williamson, P. C. (2000). Comparison of the quantification precision of human short echo time (1)H spectroscopy at 1.5 and 4.0 Tesla. Magnetic Resonance in Medicine, 44(2), 185192.Google Scholar
Baslow, M. H. (2000). Functions of N-acetyl-L-aspartate and N-acetyl-L-aspartylglutamate in the vertebrate brain: Role in glial cell-specific signaling. Journal of Neurochemistry, 75(2), 453459.Google Scholar
Baslow, M. H. (2003). Brain N-acetylaspartate as a molecular water pump and its role in the etiology of Canavan disease: A mechanistic explanation. Journal of Molecular Neuroscience, 21(3), 185190.Google Scholar
Bednarik, P., Tkac, I., Giove, F., DiNuzzo, M., Deelchand, D. K., Emir, U. E., … Mangia, S. (2015). Neurochemical and BOLD responses during neuronal activation measured in the human visual cortex at 7 T. Journal of Cerebral. Blood Flow and Metabolism, 35(4), 601610.Google Scholar
Bhakoo, K., & Pearce, D. (2000). In vitro expression of N-acetyl aspartate by oligodendrocytes: Implications for proton magnetic resonance spectroscopy signal in vivo. Journal of Neurochemistry, 74(1), 254262.Google Scholar
Chakraborty, G., Mekala, P., Yahya, D., Wu, G., & Ledeen, R. W. (2001). Intraneuronal N-acetylaspartate supplies acetyl groups for myelin lipid synthesis: Evidence for myelin-associated aspartoacylase. Journal of Neurochemistry, 78(4), 736745.Google Scholar
Chalavi, S., Pauwels, L., Heise, K.-F., Zivari Adab, H., Maes, C., Puts, N. A. J., … Swinnen, S. P. (2018). The neurochemical basis of the contextual interference effect. Neurobiology of Aging, 66, 8596. doi: 10.1016/j.neurobiolaging.2018.02.014.Google Scholar
Chen, W., Zhu, X. H., Adriany, G., & Uğurbil, K. (1997). Increase of creatine kinase activity in the visual cortex of human brain during visual stimulation: A 31P magnetization transfer study. Magnetic Resonance in Medicine, 38(4), 551557.Google Scholar
Coupland, N. J., Ogilvie, C. J., Hegadoren, K. M., Seres, P., Hanstock, C. C., & Allen, P. S. (2005). Decreased prefrontal myo-inositol in major depressive disorder. Biological Psychiatry, 57(12), 15261534.Google Scholar
Da Silva, T., Hafizi, S., Rusjan, P. M., Houle, S., Wilson, A. A., Price, I., … Mizrahi, R. (2019). GABA levels and TSPO expression in people at clinical high risk for psychosis and healthy volunteers: A PET-MRS study. Journal of Psychiatry and Neuroscience, 44(2), 111119. doi: 10.1503/jpn.170201.Google Scholar
Damoiseaux, J. S., Viviano, R. P., Yuan, P., & Raz, N. (2016). Differential effect of age on posterior and anterior hippocampal functional connectivity. NeuroImage, 133, 468476. doi: 10.1016/j.neuroimage.2016.03.047.Google Scholar
de Graaf, A. A., & Bovee, W. M. (1990). Improved quantification of in vivo 1H NMR spectra by optimization of signal acquisition and processing and by incorporation of prior knowledge into the spectral fitting. Magnetic Resonance in Medicine, 15(2), 305319.Google Scholar
De Stefano, N., Matthews, P. M., Fu, L., Narayanan, S., Stanley, J., Francis, G. S., … Arnold, D. L. (1998). Axonal damage correlates with disability in patients with relapsing-remitting multiple sclerosis. Results of a longitudinal magnetic resonance spectroscopy study. Brain, 121(Pt 8), 14691477.Google Scholar
Del Tufo, S. N., Frost, S. J., Hoeft, F., Cutting, L. E., Molfese, P. J., Mason, G. F., … Pugh, K. R. (2018). Neurochemistry predicts convergence of written and spoken language: A proton magnetic resonance spectroscopy study of cross-modal language integration. Frontiers in Psychology, 9, 1507. doi: 10.3389/fpsyg.2018.01507.eCollection2018.Google Scholar
Dienel, G. A. (2012). Fueling and imaging brain activation. ASN Neuro, 4(5), 267321. doi: 10.1042/AN20120021.Google Scholar
Du, F., Cooper, A., Lukas, S. E., Cohen, B. M., & Ongur, D. (2013). Creatine kinase and ATP synthase reaction rates in human frontal lobe measured by (31)P magnetization transfer spectroscopy at 4T. Magnetic Resonance Imaging, 31(1), 102108. doi: 10.1016/j.mri.2012.06.018Google Scholar
Duncan, J., Emslie, H., Williams, P., Johnson, R., & Freer, C. (1996). Intelligence and the frontal lobe: The organization of goal-directed behavior. Cognitive Psychology, 30(3), 257303.Google Scholar
Duncan, N. W., Wiebking, C., & Northoff, G. (2014). Associations of regional GABA and glutamate with intrinsic and extrinsic neural activity in humans – A review of multimodal imaging studies. Neuroscience and Biobehavioral Reviews, 47, 3652. doi: 10.1016/j.neubiorev.2014.07.016.Google Scholar
Fukushima, E., & Roeder, S. B. W. (1981). Experimental pulse NMR: A nuts and bolts approach. Reading, MA: Addison-Wesley.Google Scholar
Galton, F. (1883). Inquiries into human faculty. London: Macmillan.Google Scholar
Garwood, M., & DelaBarre, L. (2001). The return of the frequency sweep: Designing adiabatic pulses for contemporary NMR. Journal of Magnetic Resonance, 153(2), 155177. doi: 10.1006/jmre.2001.2340.Google Scholar
Geddes, J. W., Panchalingam, K., Keller, J. N., & Pettegrew, J. W. (1997). Elevated phosphocholine and phosphatidylcholine following rat entorhinal cortex lesions. Neurobiology of Aging, 18(3), 305308.Google Scholar
Goldstein, G., Panchalingam, K., McClure, R. J., Stanley, J. A.., Calhoun, V. D., Pearlson, G. D., & Pettegrew, J. W. (2009). Molecular neurodevelopment: An in vivo 31 P- 1 H MRSI study. Journal of the International Neuropsychological Society, 15(5), 671683.Google Scholar
Govindaraju, V., Young, K., & Maudsley, A. A. (2000). Proton NMR chemical shifts and coupling constants for brain metabolites. NMR in Biomedicine, 13(3), 129153.Google Scholar
Harper, D. G., Joe, E. B., Jensen, J. E., Ravichandran, C., & Forester, B. P. (2016). Brain levels of high-energy phosphate metabolites and executive function in geriatric depression. International Journal of Geriatric Psychiatry, 31(11), 12411249. doi: 10.1002/gps.4439.Google Scholar
Harris, A. D., Saleh, M. G., & Edden, R. A. E. (2017). Edited 1H magnetic resonance spectroscopy in vivo: Methods and metabolites. Magnetic Resonance in Medicine, 77(4), 13771389. doi: 10.1002/mrm.26619.Google Scholar
Howarth, C., Gleeson, P., & Attwell, D. (2012). Updated energy budgets for neural computation in the neocortex and cerebellum. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism, 32(7), 12221232. doi: 10.1038/jcbfm.2012.35.Google Scholar
Huang, Z., Davis, H. H., Yue, Q., Wiebking, C., Duncan, N. W., Zhang, J., … Northoff, G. (2015). Increase in glutamate/glutamine concentration in the medial prefrontal cortex during mental imagery: A combined functional MRS and fMRI study. Human Brain Mapping, 36(8), 32043212. doi: 10.1002/hbm.22841Google Scholar
Hunt, E. (1980). Intelligence as an information-processing concept. British Journal of Psychology, 71(4), 449474.Google Scholar
Ip, B., Berrington, A., Hess, A. T., Parker, A. J., Emir, U. E., & Bridge, H. (2017). Combined fMRI-MRS acquires simultaneous glutamate and BOLD-fMRI signals in the human brain. NeuroImage, 155, 113119.Google Scholar
Isaacson, J. S., & Scanziani, M. (2011). How inhibition shapes cortical activity. Neuron, 72(2), 231243. doi: 10.1016/j.neuron.2011.09.027.Google Scholar
Jahng, G. H., Oh, J., Lee, D. W., Kim, H. G., Rhee, H. Y., Shin, W., … Ryu, C. W. (2016). Glutamine and glutamate complex, as measured by functional magnetic resonance spectroscopy, alters during face-name association task in patients with mild cognitive impairment and Alzheimer’s disease. Journal of Alzheimers Disease, 53(2), 745. doi: 10.3233/JAD-169004.Google Scholar
Jung, R. E., Gasparovic, C., Chavez, R. S., Caprihan, A., Barrow, R., & Yeo, R. A. (2009). Imaging intelligence with proton magnetic resonance spectroscopy. Intelligence, 37(2), 192198. doi: 10.1016/j.intell.2008.10.009.Google Scholar
Jung, R. E., & Haier, R. J. (2007). The Parieto-Frontal Integration Theory (P-FIT) of intelligence: Converging neuroimaging evidence. Behavioral and Brain Science, 30(2), 135154; discussion 154–187.Google Scholar
Kapogiannis, D., Reiter, D. A., Willette, A. A., & Mattson, M. P. (2013). Posteromedial cortex glutamate and GABA predict intrinsic functional connectivity of the default mode network. Neuroimage, 64, 112119. doi: 10.1016/j.neuroimage.2012.09.029.Google Scholar
Kato, T., Murashita, J., Shioiri, T., Hamakawa, H., & Inubushi, T. (1996) Effect of photic stimulation on energy metabolism in the human brain measured by 31P-MR spectroscopy. Journal of Neuropsychiatry and Clinical Neuroscience, 8(4), 417422.Google Scholar
Keltner, J. R., Wald, L. L., Frederick, B. D., & Renshaw, P. F. (1997): In vivo detection of GABA in human brain using a localized double-quantum filter technique. Magnetic Resonance in Medicine, 37(3), 366371.Google Scholar
Kemp, G. J. (2000). Non-invasive methods for studying brain energy metabolism: What they show and what it means. Developmental Neuroscience, 22(56), 418428. doi: 10.1159/000017471.Google Scholar
Kim, H., McGrath, B. M., & Silverstone, P. H. (2005). A review of the possible relevance of inositol and the phosphatidylinositol second messenger system (PI-cycle) to psychiatric disorders – Focus on magnetic resonance spectroscopy (MRS) studies. Human Psychopharmacology, 20(5), 309326.Google Scholar
Kroll, J. L., Steele, A. M., Pinkham, A. E., Choi, C., Khan, D. A., Patel, S. V., … Ritz, T. (2018). Hippocampal metabolites in asthma and their implications for cognitive function. Neuroimage Clinical, 19, 213221. doi: 10.1016/j.nicl.2018.04.012.eCollection2018.Google Scholar
Kyllonen, P. C., & Christal, R. E. (1990). Reasoning ability is (little more than) working-memory capacity?! Intelligence, 14(4), 389433.Google Scholar
Lacreuse, A., Moore, C. M., LaClair, M., Payne, L., & King, J. A. (2018). Glutamine/glutamate (Glx) concentration in prefrontal cortex predicts reversal learning performance in the marmoset. Behavioral Brain Research, 346, 1115. doi: 10.1016/j.bbr.2018.01.025.Google Scholar
Lauritzen, M., Mathiesen, C., Schaefer, K., & Thomsen, K. J. (2012). Neuronal inhibition and excitation, and the dichotomic control of brain hemodynamic and oxygen responses. NeuroImage, 62(2), 10401050. doi: 10.1016/j.neuroimage.2012.01.040Google Scholar
Lin, Y., Stephenson, M. C., Xin, L., Napolitano, A., & Morris, P. G. (2012). Investigating the metabolic changes due to visual stimulation using functional proton magnetic resonance spectroscopy at 7 T. Journal of Cerebral Blood Flow and Metabolism, 32(8), 14841495. doi: 10.1038/jcbfm.2012.33.Google Scholar
Lindner, M., Bell, T., Iqbal, S., Mullins, P. G., & Christakou, A. (2017). In vivo functional neurochemistry of human cortical cholinergic function during visuospatial attention. PLoS One, 12(2), e0171338. doi: 10.1371/journal.pone.0171338.Google Scholar
Maffei, A. (2017). Fifty shades of inhibition. Current Opinion in Neurobiology, 43, 4347. doi: 10.1016/j.conb.2016.12.003Google Scholar
Mangia, S., Tkac, I., Gruetter, R., Van de Moortele, P. F., Maraviglia, B., & Uğurbil, K. (2007) Sustained neuronal activation raises oxidative metabolism to a new steady-state level: Evidence from 1H NMR spectroscopy in the human visual cortex. Journal of Cerebral Blood Flow and Metabolism, 27(5), 10551063. doi: 10.1038/sj.jcbfm.96004-01.Google Scholar
Mazoyer, B., Zago, L., Mellet, E., Bricogne, S., Etard, O., Houdé, O., … Tzourio-Mazoyer, N. (2001). Cortical networks for working memory and executive functions sustain the conscious resting state in man. Brain Research Bulletin, 54(3), 287298.Google Scholar
McEwen, B. S., & Morrison, J. H. (2013). The brain on stress: Vulnerability and plasticity of the prefrontal cortex over the life course. Neuron, 79(1), 1629. doi: 10.1016/j.neuron.2013.06.028.Google Scholar
McIlwain, H., & Bachelard, H. S. (1985). Biochemistry and the central nervous system, vol. 5. Edinburgh: Churchill Livingstone.Google Scholar
McRobbie, D., Moore, E., Graves, M., & Prince, M. (2006). MRI from picture to proton. Cambridge University Press. doi: 10.1017/CBO9780511545405.Google Scholar
Mergenthaler, P., Lindauer, U., Dienel, G. A., & Meisel, A. (2013). Sugar for the brain: The role of glucose in physiological and pathological brain function. Trends in Neurosciences, 36(10), 587597. doi: 10.1016/j.tins.2013.07.001.Google Scholar
Miller, B. L. (1991). A review of chemical issues in 1H NMR spectroscopy: N-Acetyl-L-aspartate, creatine and choline. NMR in Biomedicine, 4(2), 4752.Google Scholar
Mlynárik, V., Gambarota, G., Frenkel, H., & Gruetter, R. (2006). Localized short-echo-time proton MR spectroscopy with full signal-intensity acquisition. Magnetic Resonance in Medicine, 56(5), 965970. doi: 10.1002/mrm.21043.Google Scholar
Mochel, F., N’Guyen, T. M., Deelchand, D., Rinaldi, D., Valabregue, R., Wary, C., … Henry, P. G. (2012) Abnormal response to cortical activation in early stages of Huntington disease. Movement Disorders, 27(7), 907910. doi: 10.1002/mds.25009.Google Scholar
Murashita, J., Kato, T., Shioiri, T., Inubushi, T., & Kato, N. (1999). Age dependent alteration of metabolic response to photic stimulation in the human brain measured by 31P MR-spectroscopy. Brain Research, 818(1), 7276.Google Scholar
Neubauer, A. C., & Fink, A. (2009). Intelligence and neural efficiency. Neuroscience and Biobehavioral Reviews, 33(7), 10041023. doi: 10.1016/j.neubiorev.2009.04.001.Google Scholar
Patel, T., Blyth, J. C., Griffiths, G., Kelly, D., & Talcott, J. B. (2014). Moderate relationships between NAA and cognitive ability in healthy adults: Implications for cognitive spectroscopy. Frontiers in Human Neuroscience, 14(8), 39. doi: 10.3389/fnhum.2014.00039.eCollection2014.Google Scholar
Pettegrew, J. W., Klunk, W. E., Panchalingam, K., McClure, R. J., & Stanley, J. A. (2000). Molecular insights into neurodevelopmental and neurodegenerative diseases. Brain Research Bulletin, 53(4), 455469. doi: S0361-9230(00)00376-2[pii].Google Scholar
Pettegrew, J. W., Panchalingam, K., Withers, G., McKeag, D., & Strychor, S. (1990). Changes in brain energy and phospholipid metabolism during development and aging in the Fischer 344 rat. Journal of Neuropathology and Experimental Neurology, 49(3), 237249.Google Scholar
Pouwels, P. J., Brockmann, K., Kruse, B., Wilken, B., Wick, M., Hanefeld, F., & Frahm, J. (1999). Regional age dependence of human brain metabolites from infancy to adulthood as detected by quantitative localized proton MRS. Pediatric Research, 46(4), 474485.Google Scholar
Pradhan, S., Bonekamp, S., Gillen, J. S., Rowland, L. M., Wijtenburg, S. A., Edden, R. A. E., & Barker, P. B. (2015). Comparison of single voxel brain MRS AT 3T and 7T using 32-channel head coils. Magnetic Resonance Imaging, 33(8), 10131018. doi: 10.1016/j.mri.2015.06.003.Google Scholar
Provencher, S. W. (1993). Estimation of metabolite concentrations from localized in vivo proton NMR spectra. Magnetic Resonance in Medicine, 30(6), 672679.Google Scholar
Rango, M., Bonifati, C., & Bresolin, N. (2006). Parkinson’s disease and brain mitochondrial dysfunction: A functional phosphorus magnetic resonance spectroscopy study. Journal of Cerebral Blood Flow and Metabolism, 26(2), 283290. doi: 10.1038/sj.jcbfm.96001.-92.Google Scholar
Rango, M., Castelli, A., & Scarlato, G. (1997). Energetics of 3.5 s neural activation in humans: A 31P MR spectroscopy study. Magnetic Resonance in Medicine, 38(6), 878883.Google Scholar
Rijpma, A., van der Graaf, M., Meulenbroek, O., Olde Rikkert, M. G. M., & Heerschap, A. (2018). Altered brain high-energy phosphate metabolism in mild Alzheimer’s disease: A 3-dimensional ³¹P MR spectroscopic imaging study. Neuroimage: Clinical, 18, 254261. doi: 10.1016/j.nicl.2018.01.031.eCollection2018.Google Scholar
Ross, B., & Bluml, S. (2001). Magnetic resonance spectroscopy of the human brain. The Anatomical Record, 265(2), 5484.Google Scholar
Rothman, D. L., Petroff, O. A., Behar, K. L., & Mattson, R. H. (1993). Localized 1H NMR measurements of gamma-aminobutyric acid in human brain in vivo. Proceedings of the National Academy of Sciences USA, 90(12), 56625666.Google Scholar
Sappey-Marinier, D., Calabrese, G., Fein, G., Hugg, J. W., Biggins, C., & Weiner, M. W. (1992). Effect of photic stimulation on human visual cortex lactate and phosphates using 1H and 31P magnetic resonance spectroscopy. Journal of Cerebral Blood Flow and Metabolism, 12(4), 584592. doi: 10.1038/jcbfm.1992.82.Google Scholar
Schaller, B., Mekle, R., Xin, L., Kunz, N., & Gruetter, R. (2013). Net increase of lactate and glutamate concentration in activated human visual cortex detected with magnetic resonance spectroscopy at 7 tesla. Journal of Neuroscience Research, 91(8), 10761083. doi: 10.1002/jnr.23194.Google Scholar
Schaller, B., Xin, L., O’Brien, K., Magill, A. W., & Gruetter, R. (2014). Are glutamate and lactate increases ubiquitous to physiological activation? A (1)H functional MR spectroscopy study during motor activation in human brain at 7Tesla. NeuroImage, 93(Pt 1), 138145. doi: 10.1016/j.neuroimage.2014.02.016.Google Scholar
Scheenen, T. W. J., Klomp, D. W. J., Wijnen, J. P., & Heerschap, A. (2008). Short echo time 1H-MRSI of the human brain at 3T with minimal chemical shift displacement errors using adiabatic refocusing pulses. Magnetic Resonance in Medicine, 59(1), 16. doi: 10.1002/mrm.21302.Google Scholar
Schlattner, U., Tokarska-Schlattner, M., & Wallimann, T. (2006). Mitochondrial creatine kinase in human health and disease. Biophysica Biochimica Acta - Molecular Basis of Disease, 1 762(2), 164180. doi: 10.1016/j.bbadis.2005.09.004.Google Scholar
Shoubridge, E. A., Briggs, R. W., & Radda, G. K. (1982). 31p NMR saturation transfer measurements of the steady state rates of creatine kinase and ATP synthetase in the rat brain. FEBS Letters, 140(2), 289292. doi: 10.1016/0014-5793(82)80916-2.Google Scholar
Simmons, M. L., Frondoza, C. G., & Coyle, J. T. (1991). Immunocytochemical localization of N-acetyl-aspartate with monoclonal antibodies. Neuroscience, 45(1), 3745. doi: 10.1016/0306-4522(91)90101-s.Google Scholar
Sokoloff, L. (1991). Measurement of local cerebral glucose utilization and its relation to local functional activity in the brain. Advances in Experimental Medicine and Biology, 291, 2142. doi: 10.1007/978-1-4684-5931-5994.Google Scholar
Sokoloff, L. (1993). Function-related changes in energy metabolism in the nervous system: Localization and mechanisms. Keio Journal of Medicine, 42(3), 95-103.Google Scholar
Somogyi, P., Tamás, G., Lujan, R., & Buhl, E. H. (1998). Salient features of synaptic organisation in the cerebral cortex. Brain Research Brain Research Reviews, 26(2–3), 113135.Google Scholar
Stagg, C. J. (2014). Magnetic resonance spectroscopy as a tool to study the role of GABA in motor-cortical plasticity. Neuroimage, 86, 1927.Google Scholar
Stanley, J. A. (2002). In vivo magnetic resonance spectroscopy and its application to neuropsychiatric disorders. Canadian Journal of Psychiatry, 47(4), 315326.Google Scholar
Stanley, J., Burgess, A., Khatib, D., Ramaseshan, K., Arshad, M., Wu, H., & Diwadkar, V. (2017). Functional dynamics of hippocampal glutamate during associative learning assessed with in vivo 1H functional magnetic resonance spectroscopy. NeuroImage, 153, 189197. doi: 10.1016/j.neuroimage.2017.03.051.Google Scholar
Stanley, J. A., Kipp, H., Greisenegger, E., MacMaster, F. P., Panchalingam, K., Keshavan, M. S., … Pettegrew, J. W. (2008). Evidence of developmental alterations in cortical and subcortical regions of children with attention-deficit/hyperactivity disorder: A multivoxel in vivo phosphorus 31 spectroscopy study. Archives of General Psychiatry, 65(12), 14191428. doi: 65/12/1419[pii]10.1001/archgenpsychiatry.2008.503.Google Scholar
Stanley, J. A., & Pettegrew, J. W. (2001). A post-processing method to segregate and quantify the broad components underlying the phosphodiester spectral region of in vivo 31P brain spectra. Magnetic Resonance in Medicine, 45(3), 390396.Google Scholar
Stanley, J. A., Pettegrew, J. W., & Keshavan, M. S. (2000). Magnetic resonance spectroscopy in schizophrenia: Methodological issues and findings – Part I. Biological Psychiatry, 48(5), 357368. doi: S0006-3223(00)00949-5[pii].Google Scholar
Stanley, J. A., & Raz, N. (2018). Functional magnetic resonance spectroscopy: The “new” MRS for cognitive neuroscience and psychiatry research. Frontiers in Psychiatry – Neuroimaging and Stimulation, 9, 76. doi: 10.3389/fpsyt.2018.00076.Google Scholar
Sui, J., Huster, R., Yu, Q., Segall, J. M., & Calhoun, V. D. (2014). Function-structure associations of the brain: Evidence from multimodal connectivity and covariance studies. Neuroimage, 102(Pt 1), 1123. doi: 10.1016/j.neuroimage.2013.09.044.Google Scholar
Tallan, H. (1957). Studies on the distribution of N-acetyl-L-aspartic acid in brain. Journal of Biological Chemistry, 224(1), 4145.Google Scholar
Tatti, R., Haley, M. S., Swanson, O. K., Tselha, T., & Maffei, A. (2017). Neurophysiology and regulation of the balance between excitation and inhibition in neocortical circuits. Biological Psychiatry, 81(10), 821831. doi:10.1016/j.biopsych.2016.09.017.Google Scholar
Taylor, R., Schaefer, B., Densmore, M., Neufeld, R. W. J., Rajakumar, N., Williamson, P. C., & Théberge, J. (2015). Increased glutamate levels observed upon functional activation in the anterior cingulate cortex using the Stroop Task and functional spectroscopy. Neuroreport, 26(3), 107112. doi: 10.1097/WNR.0000000000000309.Google Scholar
Thielen, J. W., Hong, D., Rohani Rankouhi, S., Wiltfang, J., Fernández, G., Norris, D. G., & Tendolkar, I. (2018). The increase in medial prefrontal glutamate/glutamine concentration during memory encoding is associated with better memory performance and stronger functional connectivity in the human medial prefrontal-thalamus-hippocampus network. Human Brain Mapping, 39(6), 23812390. doi: 10.1002/hbm.24008.Google Scholar
Tkac, I., Andersen, P., Adriany, G., Merkle, H., Ugurbil, K., & Gruetter, R. (2001). In vivo 1H NMR spectroscopy of the human brain at 7 T. Magnetic Resonance in Medicine, 46(3), 451456.Google Scholar
Tkác, I., Starcuk, Z., Choi, I. Y., & Gruetter, R. (1999). In vivo 1H NMR spectroscopy of rat brain at 1 ms echo time. Magnetic Resonance in Medicine, 41(4), 649656.Google Scholar
Uğurbil, K., Adriany, G., Andersen, P., Chen, W., Garwood, M., Gruetter, R., … Zhu, X. H. (2003). Ultrahigh field magnetic resonance imaging and spectroscopy. Magnetic Resonance Imaging, 21(10), 12631281.Google Scholar
Urenjak, J., Williams, S. R., Gadian, D. G., & Noble, M. (1993). Proton nuclear magnetic resonance spectroscopy unambiguously identifies different neural cell types. Journal of Neuroscience, 13(3), 981989.Google Scholar
van de Bank, B. L., Maas, M. C., Bains, L. J., Heerschap, A., & Scheenen, T. W. J. (2018). Is visual activation associated with changes in cerebral high-energy phosphate levels? Brain Structure and Function, 223, 27212731.Google Scholar
van den Heuvel, M. P., Stam, C. J., Kahn, R. S., & Hulshoff Pol, H. E. (2009). Efficiency of functional brain networks and intellectual performance. Journal of Neuroscience, 29(23), 76197624. doi: 10.1523/JNEUROSCI.1443-09.2009.Google Scholar
van der Knaap, M. S., van der Grond, J., Luyten, P. R., den Hollander, J. A., Nauta, J. J., & Valk, J. (1992). 1H and 31P magnetic resonance spectroscopy of the brain in degenerative cerebral disorders. Annals of Neurology, 31(2), 202211.Google Scholar
Wallimann, T., Wyss, M., Brdiczka, D., Nicolay, K., & Eppenberger, H. M. (1992). Intracellular compartmentation, structure and function of creatine kinase isoenzymes in tissues with high and fluctuating energy demands: The “phosphocreatine circuit” for cellular energy homeostasis. Biochemical Journal, 281(Pt 1), 2140. doi: 10.1042/bj2810021.Google Scholar
Wijtenburg, S. A., McGuire, S. A., Rowland, L. M., Sherman, P. M., Lancaster, J. L., Tate, D. F., … Kochunov, P. (2013). Relationship between fractional anisotropy of cerebral white matter and metabolite concentrations measured using (1)H magnetic resonance spectroscopy in healthy adults. Neuroimage, 66, 161168. doi: 10.1016/j.neuroimage.2012.10.014.Google Scholar
Wilman, A. H., & Allen, P. S. (1995). Yield enhancement of a double-quantum filter sequence designed for the edited detection of GABA. Journal of Magnetic Resonance B, 109(2), 169174.Google Scholar
Woodcock, E. A., Anand, C., Khatib, D., Diwadkar, V. A., & Stanley, J. A. (2018). Working memory modulates glutamate levels in the dorsolateral prefrontal cortex during (1)H fMRS. Frontiers in Psychiatry, 9, 66. Epub 2018/03/22. doi: 10.3389/fpsyt.2018.00066.Google Scholar
Yang, S., Hu, J., Kou, Z., & Yang, Y. (2008). Spectral simplification for resolved glutamate and glutamine measurement using a standard STEAM sequence with optimized timing parameters at 3, 4, 4.7, 7, and 9.4T. Magnetic Resonance in Medicine, 59(2), 236244. doi: 10.1002/mrm.21463.Google Scholar
Yeo, R. A., Hill, D., Campbell, R., Vigil, J., & Brooks, W. M. (2000). Developmental instability and working memory ability in children: A magnetic resonance spectroscopy investigation. Developmental Neuropsychology, 17(2), 143159.Google Scholar
Yuksel, C., Du, F., Ravichandran, C., Goldbach, J. R., Thida, T., Lin, P., … Cohen, B. M. (2015). Abnormal high-energy phosphate molecule metabolism during regional brain activation in patients with bipolar disorder. Molecular Psychiatry, 20(9), 10791084. doi: 10.1038/mp.2015.13.Google Scholar
Zhu, X.-H., Qiao, H., Du, F., Xiong, Q., Liu, X., Zhang, X., … Chen, W. (2012). Quantitative imaging of energy expenditure in human brain. NeuroImage, 60(4), 21072117. doi: 10.1016/j.neuroimage.2012.02.013.Google 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
×