Sunteți pe pagina 1din 29

Cursul 1 – Introducere în neuropsihologie

Bibliografie video

https://www.youtube.com/watch?v=HKJlSY0DBBA – Înțelegerea lui Derrida, deconstrucție și gramatologie

https://www.youtube.com/watch?v=1Yxg2_6_YLs – Then & Now – Introducere în Baudrillard

https://www.youtube.com/watch?v=bf9J35yzM3E – Cuck Philosophy – Ce credea Baudrillard despre Matrix?

https://www.youtube.com/watch?v=pv6QHxkBFzY – Ted Aschulter – Marea dezbatere neurologică

https://www.youtube.com/watch?v=EFNpcIfvOnA&t - Daniel David - Despre Liberul Arbitru. Decizii Libere:


Există?

https://www.youtube.com/watch?v=5Yj3nGv0kn8 - Nancy Kanwisher – Un portret neuronal al minții umane

https://www.youtube.com/watch?v=uhRhtFFhNzQ – David Chalmers – Cum explici conștiința?

https://www.youtube.com/watch?v=PeZ-U0pj9LI – Thomas Insel – Către o nouă înțelegere a psihopatologiei


Bibliografie 1.
Mihai Ioan Botez – Neuropsihologie clinică și neurologia comportamentului, p.3-
5;
Introducere istorică și critică în problema
localizărilor cerebrale
Bruno Cardu

Generalități și scurt istoric


Neuropsihologia, în sens larg, reprezintă examenul legăturii între
activitatea psihologică şi afecţiunea cerebrală corespunzătoare.
Ea studiază cunt modificările de la nivelul creierului afectează
comportamentul: cum, de exemplu, ablaţia regiunilor prefrontale
la om îi va modifica acestuia capacităţile intelectuale sau
caracterul: cum ablaţia regiunilor infero-temporale la maimuţe va
influenţa percepţia şi memoria vizuală, cum ablaţia extensivă a
cortexului cerebral la sobolani va modifica invățarea unui labirint
cu mai multe locuri înfundate.
Răspunsul la aceste întrebări implică o dublă abordare cantitativă.
Pe de o parte, există abordarea ştiinţifică a psihologiei, care
vizează descrierea obiectivă şi înţelegerea funcţiilor ca percepţia,
memoria, inteligenţa, limbajul, şi, pe de altă parte, este vorba de
ansamblul de cunoştinţe ştiinţifice despre creier, descrise de
neuroanatomia macroscopică şi microscopică (citoarhitectonică,
traiecte nervoase), neurofiziologie, neurochimie, neurologie
clinică. În general, în neuropsibologie se diferenţiază:
neuropsihologia clinică, neuropsihologia experimentală, neurologia comportamentului şi neuropsihologia cognitivă.
Neuropsihologia clinică are ca obiect măsurarea şi analizarea, la om, a modificărilor capacităţilor intelectuale,
perceptuale, mnezice şi a modificărilor de personalitate apărute în urma unei leziuni cerebrale ca un accident vascular,
o intoxicaţie, o ablaţie sau propagarea intracerebrală a unei tumori. Neuropsihologul clinician stabileşte diagnosticul
leziunilor cerebrale, ţinând cont de dublul aspect psihologic şi neurologic. Pe plan psihologic, el va utiliza teste
standardizate care permit evaluarea şi situarea performanţelor psihologice ale unui pacient în interiorul scalei
cantitative a testului. Pe plan neurologic, el va avea sarcina de a stabili comportamentul pacientului, şi, în colaborare
cu neurologul şi neuroradiologul, de a stabili localizarea leziunii cerebrale, întinderea sa şi repercusiunile asupra
creierului în ansamblu. Cercetarea în neuropsihologia clinică va utiliza deseori comparaţii statistice între loturi mari
de pacienţi şi martori normali. Neuropsihologia experimentală răspunde cerinţei de rigoare şi de cercetare
fundamentală în studiul relaţiei creier-comportament. Cercetarea în neuropsihologia experimentală are loc deseori (dar
nu totdeauna) fără precauțiuni de aplicare, şi îi datorim conntribuții importante în cunoașterea mecanisemelor cerebrale
ale comportamentului, cum ar fi motivația, fenomenul de autostimulare, rolul corpului calos, etc. Este vorba, de
exemplu, de experiențe de secționare a corpului calos la pisici și la maimuțe care au modificat problematica de corp
calos la om. Neurologia comportamentului este o disciplină particulară, care se concentrează asupra analizei
aprofundate a cazurilor individuale, ca sursă de date generalizabile. În loc să evalueze un comportament raportându-l
la o scală cantitativă prealabilă, ea studiază cazurile individuale organizând situații test care permit diferențierea
deviațiilor anormale de funcționarea normală.

Neuropsihologia cognitivă, foarte răspândită și influentă la ora actuală, are ca scop al cercetării înțelegerea
mecanismelor psihologiei normale, plecând de la modificările de comportament induse de leziunile cerebrale la
oameni. Ea nu se interesează de localizarea leziunii ca atare, nici de localizarea cerebrală a funcţiilor psihologice; ea
încearcă mai ales să înţeleagă cum, plecând de la dezintegrarea psihologică produsă de leziune, se poate defini
organizarea psihoiogică normală. Cititorul interesat va putea găsi o ilustrare a acestui curent de gândire în lucrările lui
McCarthy şi Warrington şi ale lui Seron. Concepţia modernă a creierului, ca motor al activităţii psihologice, este o
cucerire relativ recentă, care datează, în forma actuală, numai de la sfârşitul secolului al XVIII-lea. Frenologia lui
Franz Joseph Gall ocupă aici un loc important. Gall era un anatomist valoros, adept al psihologiei "facultăţilor”. După
el, în spirit ar fi existat facultăţi separate ca inteligenţa, memoria, percepţia etc. A fost primul care a afirmat că aceste
facultăţi ar avea localizări precise în anumite regiuni ale creierului. Aceste regiuni, de importanţă diferită de la un
individ la altul, ar putea exercita o presiune asupra porţiunii de craniu corespunzătoare, şi da naştere unor proeminenţe
craniene numite «bose». Ar fi fost suficientă, în opinia lui, reperarea lor pe craniul unei persoane, pentru a face
inventarul psihologic al facultăţilor sale. Această concepţie, chiar fantezistă, reprezintă prima tentativă importantă de
a lega o «facultate» psihologică de o anumită structură cerebrală; era suficient ca aceasta din urmă să fie cunoscută,
pentru a o deduce pe cealaltă. Astfel a luat naştere teoria localizărilor cerebrale. Doctrina lui Gall a fost intens
combătută de contemporanii săi. Unul din cei mai mari adversari ai săi a fost fiziologul francez Pierre Flourens primul
care a studiat sistematic efectul leziunilor cerebrale asupra comportamentului la animale. În urma unor studii
experimentale extensive, el a diferenţiat şase unităţi ale sistemului nervos - emisferele cerebrale, cerebelul, corpii
cvadrigemeni, bulbul, măduva spinării şi nervii. Emisferelor cerebrale le corespundeau voinţa, judecata, amintirile,
percepţia; dar toate aceste funcţii constituiau pentu el o facultate esenţialmente unică. Pierre Flourens a fost precursorul
antilocalizaţioniştilor moderni.

De la polernica Gall-Flourens, evoluţia teoriei localizărilor cerebrale seamănă cu o spirală cu mai multe nivele. La
fiecare nivel, dialectica localizaţionistă şi antilocalizaţionistă este reluată, o generaţie sau două mai târziu, cu noi
argumente provenite din noi descoperiri. De exemplu, teza avansată de Paul Broca în 1861, după care limbajul articulat
este localizat în piciorul celei de-a treia circumvoluţiuni frontale din emisfera stângă, constituia prima localizare
serioasă a unei funcţii psihologice într-o parte determinată a creierului. În această introducere, nu vom expune istoria
descoperirilor care ar putea duce la confirmarea tezelor localizaţioniste sau antilocalizaţioniste. Aceasta poate fi găsită
într-un mare număr de lucrări specializate. Ceea ce ne propunem noi este de a descrie funcţionarea a două concepţii
opuse şi de a analiza ponderea argumentelor în favoarea fiecăreia. Nu există o corespondenţă univocă între o concepţie
teoretică în psihologie şi o concepţie cerebrală localizaţionistă sau holistică (non localizaţionistă). Până la sfarşitul
secolului al XIX-lea, doctrina asociaţionistă era dominată atât la localizaţionişti (Broca, Wernicke, Dejerine etc.)
ca şi la holiști sau non-localizaţionişti (Hughlings Jackson, Henry Head, von Monakow şi Mourgue). Mai târziu,
o concepţie holistică, inspirată din psihologia Gestalt-ului, a fost apărată în neuropsihologie de către Kurt Goldstein
şi Karl Lashley.

Bibliografie 2.
Bryan Kolb; Ian Q. Whishaw –
Fundamentals of Human Neuropsychology
p.1-3
THE DEVELOPMENT OF
NEUROPSYCHOLOGY
The term neuropsychology in its English version originated
quite recently, in part because it represented a new approach
to studying the brain. According to Daryl Bruce, it was first
used by Canadian physician William Osler in his early-
twentieth-century textbook, which was a standard medical
reference of the time. It later appeared as a subtitle to Canadian
psychologist Donald O. Hebb’s 1949 treatise on brain
function, The Organization of Behavior: A
Neuropsychological Theory. Although Hebb neither defined
nor used the word in the text itself, he probably intended it to
represent a multidisciplinary focus of scientists who believed
that an understanding of human brain function was central to understanding human behavior. By 1957, the term had
become a recognized designation for a subfield of the neurosciences. Heinrich Kluver, an American investigator into
the neural basis of vision, wrote in the preface to his Behavior Mechanism in Monkeys that the book would be of
interest to neuropsychologists and others. (Kluver had not used the term in the 1933 preface to the same book.) In
1960, it appeared in the title of a widely read collection of writings by American psychologist Karl S. Lashley—The
Neuropsychology of Lashley—most of which described rat and monkey studies directed toward understanding
memory, perception, and motor behavior. Again, neuropsychology was neither used nor defined in the text. To the
extent that they did use the term, however, these writers, who specialized in the study of basic brain function in
animals, were recognizing the emergence of a subdiscipline of investigators who specialized in human research and
would find the animal research relevant to understanding human brain function. Today, we define neuropsychology
as the study of the relation between human brain function and behavior. Although neuropsychology draws information
from many disciplines—for example, anatomy, biology, biophysics, ethology, pharmacology, physiology,
physiological psychology, and philosophy—its central focus is the development of a science of human behavior based
on the function of the human brain. As such, it is distinct from neurology, which is the diagnosis of nervous system
injury by physicians who are specialists in nervous system diseases, from neuroscience, which is the study of the
molecular basis of nervous system function by scientists who mainly use nonhuman animals, and from psychology,
which is the study of behavior more generally. Neuropsychology is strongly influenced by two traditional foci of
experimental and theoretical investigations into brain function: the brain hypothesis, the idea that the brain is the
source of behavior; and the neuron hypothesis, the idea that the unit of brain structure and function is the neuron.
This chapter traces the development of these two ideas. We will see that, although the science is new, its major ideas
are not.

THE BRAIN HYPOTHESIS

People knew what the brain looked like long before they had any idea of what it did. Very early in human history,
hunters must have noticed that all animals have a brain and that the brains of different animals, including humans,
although varying greatly in size, look quite similar. Within the past 2000 years, anatomists began producing drawings
of the brain and naming some of its distinctive parts without knowing what function the brain or its parts performed.
We will begin this chapter with a description of the brain and some of its major parts and will then consider some
major insights into the functions of the brain.

What Is the Brain?

Brain is an Old English word for the tissue that is found within the skull. Figure 1.1 shows a typical human brain as
oriented in the skull of an upright human. The brain has two relatively symmetrical halves called hemispheres, one
on the left side of the body and one on the right. Just as your body is symmetrical, having two arms and two legs, so
is the brain. If you make your right hand into a fist and hold it up with the thumb pointing toward the front, the fist
can represent the position of the brain’s left hemisphere within the skull.
Taken as a whole, the basic plan of the brain is that of a tube filled with
fluid, called cerebrospinal fluid (CSF). Parts of the covering of the tube
have bulged outward and folded, forming the more complicated looking
surface structures that initially catch the eye.

The most conspicuous outer feature of the brain consists of a crinkled


tissue that has expanded from the front of the tube to such an extent that
it folds over and covers much of the rest of the brain. This outer layer is
known as the cerebral cortex (usually referred to as just the cortex). The
word cortex, which means “bark” in Latin, is aptly chosen both because
the cortex’s folded appearance resembles the bark of a tree and because
its tissue covers most of the rest of the brain, just as bark covers a tree.
The folds of the cortex are called gyri, and the creases between them are
called sulci (gyrus is Greek for “circle” and sulcus is Greek for “trench”).
Some large sulci are called fissures, such as the longitudinal fissure that
divides the two hemispheres and the lateral fissure that divides each
hemisphere into halves (in our fist analogy, the lateral fissure is the crease
separating the thumb from the other fingers). The cortex of each hemisphere is divided into four lobes, named after
the skull bones beneath which they lie. The temporal lobe is located at approximately the same place as the thumb
on your upraised fist. The lobe lying immediately above the temporal lobe is called the frontal lobe because it is
located at the front of the brain. The parietal lobe is located behind the frontal lobe, and the occipital lobe constitutes
the area at the back of each hemisphere.

The cerebral cortex comprises most of the forebrain, so named because it develops from the front part of the tube that
makes up the embryo’s primitive brain. The remaining “tube” underlying the cortex is referred to as the brainstem.

The brainstem is in turn connected to the spinal


cord, which descends down the back in the
vertebral column. To visualize the relations
between these parts of the brain, again imagine
your upraised fist: the folded fingers represent the
cortex, the hand represents the brainstem, and the
arm represents the spinal cord. This three-part
division of the brain is conceptually useful
evolutionarily, anatomically, and functionally.
Evolutionarily, animals with only spinal cords
preceded those with brainstems, which preceded
those with forebrains. Likewise, in prenatal
development, the spinal cord forms before the
brainstem, which forms before the forebrain.
Functionally, the forebrain mediates cognitive
functions; the brainstem mediates regulatory
functions such as eating, drinking, and moving;

THE NEURON HYPOTHESIS


After the development of the brain hypothesis, that
the brain is responsible for all behavior, the second
major influence on modern neuropsychology was
the development of the neuron hypothesis, the idea
that the nervous system is composed of discrete,
autonomous units, or neurons, that can interact but
are not physically connected. In this section, we
will first provide a brief description of the cells of
the nervous system, and then we will describe how
the neuron hypothesis led to a number of ideas that
are central to neuropsychology.

Nervous System Cells


The nervous system is composed of two basic
kinds of cells, neurons and glia (a name that
comes from the Greek word for “glue”). The
neurons are the functional units that enable us to
receive information, process it, and produce
actions. The glia help the neurons out, holding
them together (some do act as glue) and providing
other supporting functions. In the human nervous
system, there are about 100 billion neurons and
perhaps 10 times as many glial cells. (No, no one
has counted them all. Scientists have estimated the
total number by counting the cells in a small sample of brain tissue and then multiplying by the brain’s volume.)
Figure 1.8 shows the three basic parts of a neuron. The neuron’s core region is called the cell body. Most of a neuron’s
branching extensions are called dendrites (Latin for “branch”), but the main “root” is called the axon (Greek for
“axle”). Neurons have only one axon, but most have many dendrites. Some small neurons have so many dendrites that
they look like garden hedges. The dendrites and axon of the neuron are extensions of the cell body, and their main
purpose is to extend the surface area of the cell. The dendrites of a cell can be a number of millimeters long, but the
axon can extend as long as a meter, as do those in the pyramidal tract that extend from the cortex to the spinal cord.
In the giraffe, these same axons are a number of meters long. Understanding how billions of cells, many with long,
complex extensions, produce behavior is a formidable task, even with the use of the powerful instrumentation available
today. Just imagine what the first anatomists with their crude microscopes thought when they first began to make out
some of the brain’s structural details. But insights into the cellular organization did follow. Through the development
of new, more powerful microscopes and techniques for selectively staining tissue, good descriptions of neurons
emerged. By applying new electronic inventions to the study of neurons, researchers began to understand how axons
conduct information. By studying how neurons interact and by applying a growing body of knowledge from chemistry,
they discovered how neurons communicate and how learning takes place.

The Neuron
The earliest anatomists who tried to examine the substructure of the nervous system found a gelatinous white
substance, almost a goo.

Eventually it was discovered that, if brain tissue were placed in alcohol or


formaldehyde, water would be drawn out of the tissue, making it firm. Then, if the
tissue were cut into thin sections, many different structures could be seen. Early
theories described nerves as hollow, fluid containing tubes; however, when the first
cellular anatomist, Anton van Leeuwenhoek (1632–1723), examined nerves with a
primitive microscope, he found no such thing. He did mention the presence of
“globules,” which may have been cell bodies. As microscopes improved, the various
parts of the nerve came into ever sharper focus, eventually leading Theodor
Schwann, in 1839, to enunciate the theory that cells are the basic structural units of
the nervous system, just as they are for the rest of the body.

An exciting development in visualizing cells was the introduction of staining, which


allows different parts of the nervous system to be distinguished. Various dyes used
for staining cloth in the German clothing industry were applied to thinly cut tissue
with various results: some selectively stained the cell body, some stained the nucleus,
and some stained the axons. The most amazing cell stain came from the application
of photographic chemicals to nervous system tissue. Italian anatomist Camillo Golgi
(1843–1926) in 1875 impregnated tissue with silver nitrate (one of the substances
responsible for forming the images in black-and-white photographs) and found that
a few cells in their entirety—cell body, dendrites, and axons—became encrusted with
silver. This technique allowed the entire neuron and all its processes to be visualized
for the first time. Golgi never described how he had been led to his remarkable
discovery. Microscopic examination revealed that the brain was nothing like an
amorphous jelly; rather, it had an enormously intricate substructure with components
arranged in complex clusters, each interconnected with many other clusters. How did
this complex organ work? Was it a net of physically interconnected fibers or a
collection of discrete and separate units? If it were an interconnected net, then
changes in one part should, by diffusion, produce changes in every other part.
Because it would be difficult for a structure thus organized to localize function, a
netlike structure would favor a holistic, or “mind,” type of brain function and
psychology.

Alternatively, a structure of discrete units functioning autonomously would favor a psychology characterized by
localization of function. In 1883, Golgi suggested that axons, the longest fibers coming out of the cell body, are
interconnected, forming an axonic net. Golgi claimed to have seen connections between cells, and so he did not think
that brain functions were localized. This position was opposed by Spanish anatomist Santiago Ramón y Cajal (1852–
1934), on the basis of the results of studies in which he used Golgi’s own silver-staining technique. Cajal examined
the brains of chicks at various ages and produced beautiful drawings of neurons at different stages of growth. He was
able to see a neuron develop from a simple cell body with few extensions to a highly complex cell with many
extensions. He never saw connections from cell to cell. Golgi and Cajal jointly received the Nobel Prize in 1906; each
in his acceptance speech argued his position on the organization of neurons, Golgi supporting the nerve net and Cajal
supporting the idea of separate cells.

Information conduction
We have mentioned early views that suggested a hydraulic flow of liquid through nerves into muscles (rminiscent of
the way that filling and emptying changes the shape and hardness of a balloon). Such theories have been called
balloonist theories. Descartes espoused the balloonist hypothesis, arguing that a fluid from the ventricles flows
through nerves into muscles to make them move English physician Francis Glisson (1597–1677) in 1677 made a direct
test of the balloon hypothesis by immersing a man’s arm in water and measuring the change in the water level when
the muscles of the arm were contracted. Because the water level did not rise, Glisson concluded that no fluid entered
the muscle (bringing no concomitant change in density). Johan Swammerdam (1637–1680) in Holland reached the
same conclusion from similar experiments on frogs, but his manuscript lay unpublished for 100 years. (We have asked
students in our classes if the water will rise when an immersed muscle is contracted. Many predict that it will.)

The impetus to adopt a theory of electrical conduction in neurons came from an English scientist, Stephen Gray (1666–
1736), who in 1731 attracted considerable attention by demonstrating that the human body could conduct electricity.
He showed that, when a rod containing static electricity was brought close to the feet of a boy suspended by a rope, a
grass-leaf electroscope (a thin strip of conducting material) placed near the boy’s nose would be attracted to the boy’s
nose. Shortly after, Italian physicist Luigi Galvani (1737–1798) demonstrated that electrical stimulation of a frog’s
nerve could cause muscle contraction. The idea for this experiment came from his observation that frogs’ legs hanging
on a metal wire in a market twitched during an electrical storm. In 1886, Joseph Bernstein (1839–1917) developed the
theory that the membrane of a nerve is polarized (has a positive charge on one side and a negative charge on the other)
and that an electric potential can be propagated along the membrane by the movements of ions across the membrane.
Many of the details of this ionic conduction were worked out by English physiologists Alan Hodgkin (1914–1988)
and Andrew Huxley (1917– 2012), who received the Nobel Prize in physiology in 1963.

As successive findings refuted the hydraulic models of conduction andbrought more dynamic electrical models into
favor, hydraulic theories of behavior were also critically reassessed. For example, Viennese psychiatrist Sigmund
Freud (1856–1939) had originally envisioned the biological basis of his theory of behavior, with its three levels of id,
ego, and superego, as being a hydraulic mechanism of some sort. Although conceptually useful for a time, it had no
effect on concepts of brain function, because there was no evidence of the brain functioning as a hydraulic system.

Connections Between Neurons as the Basis of Learning


Even though neurons are independent structures, they must influence one another. Charles Scott Sherrington (1857–
1952), an English physiologist, examined how nerves connect to muscles and first suggested how the connection is
made. He applied an unpleasant stimulation to a dog’s paw, measured how long it took the dog to withdraw its foot,
and compared that rate with the speed at which messages were known to travel along axons. According to

Sherrington’s calculations, the dog took 5 milliseconds too long to respond. Sherrington theorized that neurons are
connected by junctions, which he called synapses (from the Greek word for “clasp”), and that additional time is
required for the message to get across the junction. The results of later electron microscopic studies were to confirm
that synapses do not quite touch the cells with which they synapse. The general assumption that developed in response
to this discovery was that a synapse releases chemicals to influence the adjacent cell. In 1949, on the basis of this
principle, Donald Hebb proposed a learning theory stating that, when individual cells are activated at the same time,
they grow connecting synapses or strengthen existing ones and thus become a functional unit. He proposed that new
or strengthened connections, sometimes called Hebb or plastic synapses, are the structural bases of memory. Just how
synapses are formed and change is a vibrant area of research today.
Bibliografie 3.
Mark Solms; Michael Saling – A Moment of Transition. Two Neuroscientific
Articles by Sigmund Freud – p.vii-xv;
Foreword
Mortimer Ostow
World War II, and especially the Holocaust that accompanied it,
drove the centre of ferment of psychoanalysis from Central Europe,
where it had originated, to England and the United States—the
English-speaking world. As a result it bcame necessary to provide
accurate English translations of the foundation works of
psychoanalysis, Freud's essays and books, as well as the
contributions of his disciples. These translations have been
undertaken by a series of dedicated and gifted individuals, for whose
efforts we are grateful. Of course the papers first translated were
those that bore directly upon clinical practice and its supporting
theory. Only latterly have the works of primarily historical interest
been presented to the English-speaking public.

Nevertheless, as Solms and Sating note, many of Freud's


neurological works remain untranslated, so that English readers
are deprived of the opportunity to see Freud functioning as a
neurologist and then turning his interest to psychopathology, on
the way to the evolution of psychoanalysis. Solms has undertaken
to provide a series of English translations of Freud's hitherto
untranslated neurological papers, the first two of which are presented
in this volume. As the reader will observe, the work of translation has been carried out with an unusual degree of
meticulousness, faithfully rendering even complex thoughts into clear, lucid, unambiguous, and easily readable
English. The translator and his associate, Michael Saling, have provided also a thoroughgoing discussion of the origin
of these pre-analytical articles, what they reveal to us about the scientific climate in which Freud worked, how his
work and ideas related to those of his contemporaries, and how they can be seen as steps on the road to the discipline
of psychoanalysis. They also confront various challenges that have been posed to psychoanalysis in recent
decades, that are based upon claims that Freud was limited by the archaic and dated views of central nervous
system function that prevailed in his time. The authors demonstrate decisively Freud's departure from the
conventional wisdom and the development of his original ideas. We are treated to a view of the thinking of the most
distinguished neuropsychiatrist of history as he disappointingly comes to grips with the fact that neurological
studies offer no window onto the discipline of psychopathology. On the contrary, Freud observed, psychological
considerations must be taken into consideration in anatomical and physiological theorizing. For example, in the
'Aphasie' article, on the basis of clinical observation of aphasic speech disorders, he rejects the localization theories
of aphasia, the concept that various 'speech centres' each controls a specific aspect of speech, proposing instead a
'speech field', a broad area in which the operations subserving speech are carried on. Lesions in specific regions
impair those operations in some-what specific ways because they impinge upon fibres passing through these regions,
but one is not justified to infer that these regions are discrete speech centres in the sense that they are concerned with
the psychological formulation or control of speech. In this respect, Freud anticipated important developments in
twentieth-century aphasiology and is considered by the editors of this volume to be one of the founding fathers of
modern neuropsychology.

The 'Gehirn' (Brain) article demonstrates cogently Freud's dynamic mode of thinking, so that even his description of
the gross neuroanatomy of the brain takes the form of an exploration rather than a colourless exposition. I refer to
Freud as a neuropsychiatrist deliberately, because until recent decades, neurology and psychiatry were practised
by the same individual, the assumption being that knowledge of the one would facilitate the practice of the
other, and that the knowledge of brain anatomy and physiology would facilitate the practice of both. For similar
reasons, obstetrics and gynaecology are frequently practised together to this day.

Freud was a neuropsychiatrist. So was Meynert, and so was Charcot. The psychoanalytical reader will recognize the
name of Paul Flechsig, whose neuroanatomical investigations Freud quotes admiringly, as the psychiatrist who cared
for Daniel Paul Schreber. In practice, the two disciplines are approached independently. In the second half of the
nineteenth century, their interests coincided primarily in the case of dementia paralytics and of focal lesions of the
brain, traumatic, vascular, or neoplastic. In our day, psychopharmacological theory has been promising to promote a
fruitful exchange between neurophysiology and neurochemistry on the one hand, and psychiatric theory on the other,
but at this time, psycho-pharmacology remains an empirical discipline.

Sigmund Freud, like the other scholarly neuropsychiatrists of his time, seems to have been beguiled by the hope
that neuroanatomy could illuminate psychical function. He was disappointed. In the two essays in this book I
was able to find only two examples of a claimed relation between these areas. Freud, following Meynert, calls the
fibres linking parts of the cerebral cortex to each other, association fibres, 'because they serve the association of ideas'
(ibid.). Despite these exceptions. Freud seems to be giving up the hope of correlating brain structure with mental
life. He says clearly that he does not know how to relate brain function to psychical function. We see him here
cutting his psychology loose from neuroanatomy and neurophysiology. Nevertheless, in his 1891 essay on aphasia,
he comes out clearly for psychophysical parallelism (see Jones, 1963, p. 403). Evidently the transition from the search
for a material basis for psychological and psychopathological phenomena to the determination to formulate a theory
of mental events in their own terms was not a smooth one. In the Cocaine Papers (Freud, 1884-87) which were written
just before these two pre-analytical articles, we find splendid clinical descriptions of the effects of cocaine and
suggestions about its clinical utility, but almost no attempt to contrive any kind of psychological description or
theory.

In the same encyclopedia in which these two pre-analytical papers were published, Freud also published an article on
hysteria (1888b3), in which he makes no attempt to relate its symptoms to brain structure or function, although
he writes that:

“Hysteria is based wholly and entirely on physiological modifications of the nervous system and its essence should
be expressed in a formula which took account of the conditions of excitability in the different parts of the nervous
system”

In 1886, Freud announced that he was preparing a paper on 'Some points for a comparative study of organic and
hysterical motor paralyses' (1893c) (Quelques considerations pour une etude comparative des paralysies motrices
organiques et hysteriques). Presumably it was written soon after the cocaine papers and almost simultaneously with
the two papers in this volume. Here he proposes perhaps the first description of what has come to be called 'structure'
in psychoanalytical theory—namely, an enduring complex of psychical function, in this case, the conception of a
thing.

Meanwhile, in 1891, Waldeyer gave final form to the neurone theory. The word 'neurone' appears in Freud's 1893
paper, followed by a parenthetical definition, 'cellulo-fibrillary neural unit'. With the concept of the neurone, Freud
saw an opportunity to revive his quest to base his psychology on knowledge of the structure of the nervous
system. He forsook gross anatomy for a feature of microscopic anatomy. But he realized intermittently that he
was not building upon true anatomy and physiology, but only on a fairly gross and simplistic schematic
hypothesis.

“Anyone, however, who is engaged scientifically in the construction of hypotheses will only begin to take his theory
seriously if they can be fitted into our knowledge from more than one direction and if the arbitrariness of a constructio
ad hoc can be mitigated in relation to them. (1950a (1895), p. 302)

Freud was evidently trying to overcome his own doubts. It is not surprising therefore to learn that during the
year 1895 he alternately worked feverishly on the Project and abandoned it. Focusing on the usefulness of the
'energy' concept, Freud, in the 'Project', ignored almost completely the complex gross structure of the nervous system,
the knowledge of which he demonstrates and communicates so skilfully in Gehirn. He differentiates between sensory
and motor nerves, between the superficial and deeper layers of the cortex, and among the various sensory reception
areas (ibid., p. 315). Otherwise he treats the brain as a homogeneous structure, composed of a collection of small
objects, which he calls neurones, all of which have the same properties. He does not allow for variation among
them in these properties, nor for spontaneous activity internally generated. By discarding the manuscript, he tried
to terminate this attempt to relate psychology to brain function permanently.

The 'energy' factor has run like a red thread from Freud's earliest thoughts about psychology and psychopathology
until the last. In 'Gehirn' he speaks of the 'specific energy' of sensory nerves (p. 65, this volume). In his final and
uncompleted work on psychoanalytical theory, An Outline of Psycho-analysis (1940a), Freud wrote:

“The future may teach us to exercise a direct influence by means of particular chemical substances on the amounts of
energy and their distribution in the mental apparatus”

I believe that in that statement he was thinking back to his experiences with cocaine, restating his persistent belief
in the importance of an energy factor in psychology, and anticipating the mode of action of other
psychopharmacological agents. It is of no importance that his earliest thoughts about energy dealt with measurable
conduction processes in nerves and his later thoughts with a non-material, hypothetical parameter of psychical
function. In both instances he was concerned about motivation and its disturbances.

Meanwhile, psychoanalysis has flourished. Its theories, the fundamentals of which were established by Freud, have
been elaborated, altered and simplified, challenged and defended. Practice has changed less than theory and is probably
more consistent among different psychoanalytical communities. While in some locations the demand for
psychoanalytical treatment has diminished and authentic analysis has been displaced to some extent by inauthentic
analysis and by other therapies, psychoanalysis has contributed to a host of psychodynamic therapies, so that its
influence has grown greatly.

The dynamic that Freud saw one hundred years ago was that psychology had to be disengaged from the
anatomy and physiology that was then available and permitted to develop on its own, and then in developed
form slowly recruited to re-engage a far more subtle and sophisticated anatomy, physiology, pharmacology,
and chemistry. As he has taught us in other contexts, needs can often best be satisfied if they are temporarily
renounced. Psychoanalysis, Freud said twenty years after these papers were written,

hopes to discover the common ground on the basis of which the convergence of physical and mental disorder will
become intelligible. With this aim in view, psychoanalysis must keep itself free from any hypothesis that is alien to
it, whether of an anatomical, chemical or physiological kind, and must operate entirely with purely psychological
auxiliary ideas (Freud, 1916-17, p. 211)

In these papers we see the beginning of that renunciation and disengagement; they give us a glimpse of this
crucial moment, the preparation for the application of the psycho-analytical method to the resolution of
problems of human psychology. We are impatient to initiate the process of reengagement, but it continues to
elude us.
Bibliografie 4.
Michael Gazzaniga – The Mind’s Past – p.11-13
Over a hundred years ago William James lamented, “I wished by
treating Psychology like a natural science, to help her to become
one.”
Well, it never occurred. Psychology, which for many was the study
of mental life, gave way during the past century to other disciplines.
Today the mind sciences are the province of evolutionary
biologists, cognitive scientists, neuroscientists, psychophysicists,
linguists, computer scientists—you name it. This book is about
special truths that these new practitioners of the study of mind have
unearthed. Psychology itself is dead. Or, to put it another way,
psychology is in a funny situation. My college, Dartmouth, is
constructing a magnificent new building for psychology. Yet its
four stories go like this: The basement is all neuroscience. The first
floor is devoted to classxi rooms and administration. The second
floor houses social psychology, the third floor, cognitive science,
and the fourth, cognitive neuroscience. Why is it called the
psychology building?

Traditions are long lasting and hard to give up. The odd thing is that
everyone but its practitioners knows about the death of psychology.
A dean asked the development office why money could not be
raised to reimburse the college for the new psychology building. “Oh, the alumni think it’s a dead topic, you know,
sort of just counseling. If those guys would call themselves the Department of Brain and Cognitive Science, I could
raise $25 million in a week.” The grand questions originally asked by those trained in classical psychology have
evolved into matters other scientists can address. My dear friend the late Stanley Schachter of Columbia University
told me just before his death that his beloved field of social psychology was not, after all, a cumulative science. Yes,
scientists keep asking questions and using the scientific method to answer them, but the answers don’t point to a body
of knowledge where one result leads to another. It was a strong statement—one that he would be the first to qualify.
But he was on to something. The field of psychology is not the field of molecular biology, where new discoveries
building on old ones are made every day. This is not to say that psychological processes and psychological states are
uninteresting, even boring, subjects.
On the contrary, they are fascinating pieces of the mysterious unknown that many curious minds struggle to
understand. How the brain enables mind is the question to be answered in the twenty-first century—no doubt about it.
The next question is how to think about this question. That is the business of this little book. I think the message here
is significant, one important enough to be held up for examination if it is to take hold. My view of how the brain works
is rooted in an evolutionary perspective that moves from the fact that our mental life reflects the actions of many,
perhaps dozens to thousands, of neural devices that are built into our brains at the factory. These devices do crucial
things for us, from managing our walking and breathing to helping us with syllogisms.

There are all kinds and shapes of neural devices, and they are all clever.

At first it is hard to believe that most of these devices do their jobs before we are aware of their actions. We human
beings have a centric view of the world. We think our personal selves are directing the show most of the time. I argue
that recent research shows this is not true but simply appears to be true because of a special device in our left brain
called the interpreter. This one device creates the illusion that we are in charge of our actions, and it does so by
interpreting our past—the prior actions of our nervous system.
Bibliografie 5.
J. Graham Beaumont – Introduction to Neuropsychology, Second Edition, p.3-8;
WHAT IS NEUROPSYCHOLOGY?

The human brain is a fascinating and engimatic machine. Weighing only


about 3 pounds (1.36 kilograms) and with a volume of about 1,250 cubic
centimeters, it has the ability to monitor and control our basic life support
systems, to maintain our posture and direct our movements, to receive
and interpret information about the world around us, and to store
information in a readily accessible form throughout our lives. It allows us
to solve problems that range from the strictly practical to the highly
abstract, to communicate with our fellow human beings through
language, to create new ideas and imagine things that have never existed,
to feel love and happiness and disappointment, and to experience an
awareness of ourselves as individuals. Not only can the brain undertake
such a variety of different functions, but it can do more or less all of them
simultaneously. How this is achieved is one of the most challenging and
exciting problems faced by contemporary science. It has to be said at the
outset that we are completely ignorant of many of the things that the brain
does, and of how they are done. Nevertheless, very considerable advances
have been made in the neurosciences over the last decade or two, and
there is growing confidence among neuroscientists that a real
understanding is beginning to emerge. This feeling is encouraged by the
increasing integration of the various disciplines involved in neuroscience,
and a convergence of both experimental findings and theoretical models.

Neuropsychology, as one of the neurosciences, has grown to be a separate field of specialization within psychology
over about the last 40 years, although there has always been an interest in it throughout the 120-year history of modern
scientific psychology. Neuropsychology seeks to understand the relationship between the brain and behavior, that is,
it attempts to explain the way in which the activity of the brain is expressed in observable behavior. What mechanisms
are responsible for human thinking, learning, and emotion, how do these mechanisms operate, and what are the effects
of changes in brain states upon human behavior? There are a variety of ways in which neuropsychologists conduct
their investigations into such questions, but the central theme of each is that to understand human behavior we need
to understand the human brain. A psychology without any reference to physiology can hardly be complete. The
operation of the brain is relevant to human conduct, and the understanding of how the brain relates to behavior may
make a significant contribution to understanding how other, more purely psychological, factors operate in directing
behavior. Just how the brain deals with intelligent and complex human functions is, in any case, an important subject
of investigation in its own right, and one that has an immediate relevance for those with brain injuries and diseases,
as well as a wider relevance for medical practice.

BRANCHES OF NEUROPSYCHOLOGY

Neuropsychology is often divided into two main areas: clinical neuropsychology and experimental neuropsychology.
The distinction is principally between clinical studies, on brain-injured subjects, and experimental studies, on normal
subjects, although the methods of investigation also differ. The division between the two is not absolutely clear-cut
but it helps to form an initial classification of the kinds of work in which neuropsychologists are involved.

Clinical neuropsychology deals with patients who have lesions of the brain. These lesions may be the effects of
disease or tumors, may result from physical damage or trauma to the brain, or be the result of other biochemical
changes, perhaps caused by toxic substances. Trauma may be accidental, caused by wounds or collisions; it may result
from some failure in the vascular system supplying blood to the brain; or it may be the intended result of neurosurgical
intervention to correct some neurological problem. The clinical neuropsychologist measures deficits in intelligence,
personality, and sensory–motor functions by specializedtesting procedures, and relates the results to the particular
areas of the brain that have been affected. The damaged areas may be clearly circumscribed and limited in extent,
particularly in the case of surgical lesions (when an accurate description of the parts of the brain that have been
removed can be obtained), or may be diffuse, affecting cells throughout much of the brain, as is the case with certain
cerebral diseases. Clinical neuropsychologists employ these measurements not only in the scientific investigation of
brain–behavior relationships, but also in the practical clinical work of aiding diagnosis of brain lesions and
rehabilitating brain-injured patients.

Behavioral neurology, as a form of clinical neuropsychology, also deals with clinical patients, but the emphasis is
upon conceptual rather than operational definitions of behavior. The individual case rather than group statistics is the
focus of attention, and this approach usually involves less formal tests to establish qualitative deviations from “normal”
functioning. Studies in behavioral neurology may often sample broader aspects of behavior than is usual in clinical
neuropsychology. The distinction between clinical neuropsychology and behavioral neurology is not entirely clear,
and it is further blurred by the historical traditions of investigation in different countries, particularly in the United
States, the former Soviet Union, and Great Britain. Examples of clinical work in these countries are discussed below
and in Chapter 15. By contrast, experimental neuropsychologists work with normal subjects with intact brains. This
is the most recent area of neuropsychology to develop and has grown rapidly since the 1960s, with the invention of a
variety of techniques that can be employed in the laboratory to study higher functions in the brain. There are close
links between experimental neuropsychology and general experimental and cognitive psychology, and the laboratory
methods employed in these three areas have strong similarities. Subjects are generally required to undertake
performance tasks while their accuracy or speed of response is recorded, from which inferences about brain
organization can be made. Associated variables, including psychophysiological or electrophysiological variables, may
also be recorded.

COMPARATIVE NEUROPSYCHOLOGY

Although the subject of this book is human neuropsychology, it should not be forgotten that much experimental
neuropsychology has been conducted with animals, although this form of research is now in decline. At one time, the
term neuropsychology was in fact taken to refer to this area, but it is now used more generally and the relative
importance of the animal studies of comparative neuropsychology has decreased. The obvious advantage of working
with animals, ethical issues apart, is that precise lesions can be introduced into the brain and later confirmed by
histology. Changes in the animal’s behavior are observed and can be correlated with the experimental lesions. The
disadvantages are the problems of investigating high-level functions using animals as subjects (the study of language
is ruled out, to take the most obvious example) and the difficulty of generalizing from the animal brain to the human
brain. Although it may be possible to discover in great detail how some perceptual function is undertaken in the brain
of the rat, the cat, or the monkey, it may not necessarily be undertaken in the same way in the human brain. There are
also basic differences in the amount and distribution of different types of cortical tissue in the brains of the various
animals and of humans, which add to the difficulties of generalization.

Nevertheless, animal studies continue to be important, particularly with regard to the functions of subcortical
systems—those functions located in the structures below the surface mantle of the brain that deal with relatively basic
aspects of sensation, perception, learning, memory, and emotion. These systems are harder to study in humans,
because damage to these regions may interfere much more radically with a whole range of behaviors, and may often
result in death. One of the problems facing contemporary neuropsychology is to integrate the study of cortical
functions and higher-level behaviors, which have generally been studied in humans, with the study of subcortical
structures and more basic behavioral systems, which have been studied in animals. These have tended to be separate
areas of research, although there are now signs of integration between the two. For example, intelligence is now being
discussed not just in terms of human performance on intelligence tests, but also in terms of underlying basic processes
of learning, attention, and motivation that are only understood, in neuropsychological terms, from animal studies.
Sexual behavior is another area where the basic systems are only open to experimental study in animals, yet must be
viewed within the context of socialized and cognitively controlled behavior in humans.

CONCEPTUAL ISSUES
Neuropsychology suffers philosophical and conceptual difficulties no less than other areas of psychology, and perhaps
more than many. There are two problems in particular of which every student of the subject should be aware. The first
of these springs from the nature of the methods that must be used in neuropsychological investigation. Descriptions
of brain organization can only be relatively distant inferences from the human performance that is actually observed.
The real states of the brain are not observed. Behavioral measures are taken, and by a line of reasoning that is based
on background information about either the general arrangement of the brain (in the case of experimental
neuropsychology) or about the gross changes in the brain of a particular type of patient (in the case of clinical
neuropsychology), conclusions are drawn about what the correlation must be between brain states and behavior. The
one exception to this general rule is in electrophysiological studies and studies of cerebral blood flow and metabolism
through advanced scanning techniques, where actual brain states can be observed, albeit rather crudely, in “real time”
alongside the human performance being measured. This makes these studies of special importance in
neuropsychology.

However, in general, neuropsychological study proceeds only by inference. It is important to remember this in
assessing the validity of many of the findings claimed by neuropsychologists, and also to be particularly vigilant that
the reasoning used in drawing inferences is soundly based and the data not open to alternative explanations.

The second problem is even more fundamental, and is that usually referred to as the mind–body problem. It is a subject
far too complex to receive satisfactory treatment here, but in brief it is concerned with the philosophical difficulties
that arise when we talk about mental events or “mind,” and physiological events or “body,” and try to relate the two.
We first have to decide whether mind and body are, or are not, fundamentally different kinds of things. If they are,
then there are problems in giving explanations that correlate the two. If they are not, then we have to be careful not to
be misled by our everyday language and concepts, which tend to treat mind and body as if they were different kinds
of things. The debate has gone on for some centuries, and is far from being resolved, but there is a general position
accepted by most if not all neuropsychologists. This position is known as “emergent materialism” or “emergent
psychoneural monism.” It rejects the idea that mind and body are fundamentally different (hence it is “monist” rather
than “dualist”) and proposes that all mental states are states of the brain. Mental events therefore exist but are not
separate entities. However, mental states cannot be reduced to a set of physical states because the brain is not a physical
machine but a biosystem, and so possesses properties peculiar to living things. The brain is seen as not simply a
complex composition of cells, but as having a structure and an environment. The result is that there are “emergent”
properties that include being able to think and feel and perceive. These properties are emergent just as the sweetness
of an apple is an emergent property. There is nothing in the chemistry or physical structure of the apple that possesses
sweetness. It is the whole object, in interaction with the eater, that produces the quality of sweetness. Mind is therefore
seen as a collection of emergent bioactivities, and this has implications for both theories and methods in
neuropsychology. It means that it is sometimes quite proper and sensible to reduce explanations to lower levels of
description, purely in terms of the physiology or the biochemistry involved. However, it also means that integration
among these lower processes and their description in terms of higher level concepts (concerning the emergent
properties) are both feasible and valuable.
The student first taking an interest in neuropsychology should not be overly concerned about these philosophical
issues; much, if not most, of neuropsychological work is conducted while ignoring them altogether. However, some
position is always implied in any investigation or theoretical model, and it is wise not to lose sight of the implications
of holding a particular position for a satisfactory understanding of how the brain works.
Bibliografie 6.

Leon Dănăilă; Mihai Golu – Tratat de neuropsihologie, volumul 1, p.17;


Capitolul I
RAPORTUL PSIHIC-CREIER

Indiferent dacă definirea neuropsihologiei are un caracter mai


restrâns sau mai extins, domeniul ei specific de studiu rămâne
tot raportul psihic-creier. Principala sa finalitate
epistemologică rezidă tocmai în elucidarea, pe baze
experimentale şi clinice, a naturii şi esenţei acestui raport în jurul
căruia, în istoria gândirii filosofice şi ştiinţifice, au existat aprinse
dispute şi controverse. Trebuie subliniat că dacă problema originii
şi naturii psihicului s-a conştientizat şi a devenit obiect de
preocupare intelectuală din cele mai vechi timpuri, de când
omul a dobândit conştiinţă de sine, studiul organizării
structural-funcţionale a creierului a intrat mult mai târziu în
circuitul epistemologic. Evidenţierea şi afirmarea legăturii
celor două entităţi - psihicul şi creierul - se realizează abia în
antichitatea târzie, doar cu câteva secole înaintea erei noastre.
Până atunci, cea mai inrădăcinată era convingerea că sufletul
este un atribut al întregului corp, mecanismul „dinamizării"
şi „primenirii" lui fiind considerat actul respiraţiei sau circulaţia
sângelui. Chiar în secolul V a.e.n. , Hippocrate şi Kroton
implicau creierul numai în realizarea gândirii, a procesele şi
stările afective fiind puse pe seama aparatului cardiovascular.

De-abia în secolul II î.e.n., Galen a făcut un pas mai serios înainte, afirmând într-o formă mai explicită şi mai
completă existenţa unei legături permanente între viaţa psihică internă şi creier. El formulează pentru prima dată
ipoteza localizării directe a funcţiilor şi proceselor psihice în structurile cerebrale.

Astfel, considerând că impresiile din lumea externă pătrund în forma fluidelor, prin ochi, în ventriculii
cerebrali, acest gânditor conchidea că talamusul optic reprezintă acel mecanism în care fluidele respective se
asociază cu fluidele vitale sosite din ficat, transformându-se, la nivelul sistemului vascular, în fluide psihice
(pneuma psihikon sau pneuma loghistikon). Pe cât de naivă şi puerilă ne pare astăzi această explicaţie, pe
atât de avansată şi revoluţionară a fost ea pentru timpul acela. De altfel, ideea că ventriculii cerebrali
(mai exact, lichidul care-i irigă) constituie substratul material nemijlocit al psihicului s-a perpetuat mai bine de
un mileniu şi jumătate. Modul de abordare şi soluţionare ulterioară a problemei raportului dintre psihic şi
creier a fost condiţionat atât de evoluţia reprezentărilor şi testărilor psihologice, cat şi de perfecţionarea
metodelor şi tehnicilor de investigare şi descriere anatomofiziologică a sistemului nervos. În general,
reprezentările şi teoriile psihologice, evoluand în cadrul diferitelor sisteme filosofice s-au detașat tot mai mult de
suportul lor intuitiv-concret iniţial, dobândind un caracter speculativ-abstract. Chiar în aceste condiţii, viaţa
psihică a omului nu va mai fi însă abordată global, ca ceva amorf, ci va fi supusă operaţiei de analiză şi
clasificare, care va duce la desprinderea şi descrierea unor funcţii precise şi capacităţi particulare distincte.
Cât priveşte cunoaşterea sistemului nervos, ea va continua încă multă vreme să se menţină la un nivel vag,
ipotetic, lipsită de un material faptic dobândit prin metode ştiinţifice riguroase. Aceasta a făcut ca raportarea
psihicului la substratul neuronal să fie concepută tot într-o manieră globală, fenomenologică.
Bibliografie 7.

Todd E. Feinberg; Martha J. Farah – Behavioral Neurology and


Neuropsychology, p.14-16;
THE RISE OF EXPERIMENTAL
NEUROPSYCHOLOGY

Most of the advances described so far in this chapter were made


by studying individual patients, or at most a small series of
patients with similar disorders. In many instances, particularly
before the middle of this century, patients' behavior was studied
relatively naturalistically, without planned protocols or
quantitative measurements. In the nineteen sixties and
seventies, a different approach to the study of brain-behavior
relations took hold. Neurologists and neuropsychologists began
to design experiments patterned on research methods in
experimental psychology.

Typical research designs in experimental psychology involved


groups of normal subjects given different experimental
treatments (for example, different training or different stimulus
materials), and the effects of the treatments were measured in
standardized protocols and compared using statistical methods
such as analysis of variance. In neuropsychology, the
"treatments" were, as a rule, naturally occurring brain lesions.
Groups of patients with different lesion sites or behavioral syndromes were tested with standard protocols, yielding
quantitative measures of performance, and these performances were compared across patient groups and with non-
brain-damaged control groups. Unlike the impairments studied previously in single-case designs, which were so striking
that control subjects would generally have been superfluous, experimental neuropsychology often focused on group
differences of a rather subtle nature, which required statistical analysis to substantiate.

The most common question addressed by these studies concerned localization of function. Often the localization
sought was no more precise than left versus right hemisphere or one quadrant of the brain (which, in the days before
computed tomography, often amounted to left versus right hemisphere with presence or absence of visual field
defects and/or hemiplegia). Given the huge amount of research done during this period on language, memory,
perception, attention, emotion, praxis, and so-called executive functions, it would be hopeless even to attempt a
summary. For those interested in some examples of this approach, we cite here some classic papers from a variety of
the active laboratories of the period, addressing the question "Is the right hemisphere specialized for spatial
perception of properties such as location,78-8° orientation,81-82 and large-scale topography? 83,84, ,

The influential research program of the Montreal Neurological Institute also began during this period. In the wake of
William Scoville's discovery that the bilateral medial temporal resection he performed on epileptic patient H.M.
resulted in permanent and dense amnesia, Brenda Milner and her colleagues investigated this patient and groups of
other operated epileptic patients. This enabled them to address questions of functional localization with the anatomic
precision of known surgical lesions (e.g., see Refs. 85 and 86 for reviews of research from that period on frontal lobe
function and temporal lobe function, respectively). At the same time, another surgical intervention for epilepsy,
callosotomy, also spawned a productive and influential research program. Roger Sperry and his students and
collaborators were able to address a wide variety of questions about hemispheric specialization by studying the isolated
functioning of the human cerebral hemispheres.'

In addition to answering questions about localization, the experimental neuropsychology of the sixties and seventies
also uncovered aspects of the functional organization of behavior. By examining patterns of association and
dissociation among abilities over groups of subjects, researchers tried to determine which abilities depend on the
same underlying functional systems and which are functionally independent. For example, the frequent association
of aphasia and apraxia had been taken by some to support the notion that aphasia was not language-specific but
was just one manifestation of a more pervasive loss of the ability to symbolize or represent ("asymbolia"). A
classic group study by Goodglass and Kaplan undermined this position by showing that severity of apraxia
and aphasia were uncorrelated in a large sample of left-hemisphere-damaged subjects. A second example of the
use of dissociations between groups of patients from this period is the demonstration of the functional
distinction, by Newcombe and Russell, within vision between pattern recognition and spatial orientation. 89

By the end of the seventies, experimental neuropsychology had matured to the point where many perceptual,
cognitive, and motor abilities bad been associated with particular brain regions, mod certain features of the
functional organization of these abilities had been delineated. Accordingly, it was at this time that first editions
of some at the best-known neuropsychology texts appeared, such as those by Hecaen and Albert, Heilman
and Valenstein, Kolb and Whishaw, Springer and Deutsch, and Walsh.

Despite the tremendous progress of this period, experimental neuropsychology remained distinct from and
relatively unknown within academic psychology. Particularly in the United States, but also to a large extent in
Canada and Europe (the three largest contributors to the world's psychology literature), experimental
neuropsychologists tended to work in medical centers rather than university psychology departments and to
publish their work in journals separate from mainstream experimental psychology. An important turning point
in the histories of both neuropsychology and the psychology of normal human function came when researchers
in each area became aware of the other.

THE MARRIAGE OF EXPERIMENTAL NEUROPSYCHOLOGY AND COGNITIVE PSYCHOLOGY

The predominant approach to human experimental psychology in the 1970s was cognitive psychology. The
hallmark of this approach was the assumption that all of cognition (broadly construed to include perception and
motor control) could be viewed as information processing. Although the effects of damage to an information-
processing mechanism might seem to be a good source of clues as to its normal operation, cognitive psychol-
ogists of the seventies were generally quite ignorant of contemporary neuropsychology.

The reason that most cognitive psychologists of the 1970s ignored neuropsychology stemmed from an overly
narrow conception of information processing, based on the digital computer. A basic tenet of cognitive psychology
was the computer analogy for the mind: the mind is to the brain as software is to hardware in a computer. Given
that the same computer can run different programs and the same program can be run on different
computers, this analogy suggests that hardware and software are independent and that the brain is therefore
irrelevant to cognitive psychology. If you want to understand the nature of the program that is the human mind,
studying neuropsychology is as pointless as trying to understand how a computer is programmed by looking at
the circuit boards.

The problem with the computer analogy is that hardware and software are independent only for very special
types of computational systems: those systems that have been engineered, through great effort and ingenuity, to
make the hardware and software independent, enabling one computer to run many programs and enabling
those programs to be portable to other computers. The brain was "designed" by very different pressures, and
there is no reason to believe that, in general, information-processing functions and the physical subtrate of those
functions will be independent. In fact, as cognitive psychologists finally began to learn about
neuropsychology, it became apparent that cognitive functions break down in characteristic and highly
informative ways after brain damage. By the early 1980s, cognitive psychology and neuropsychology were finally
in communication with one another. Since then, we have seen an explosion of meetings, books, and new
journals devoted to so-called cognitive neuropsychology. Perhaps more important, existing cognitive psy-
chology journals have begun to publish neuropsychological studies, and articles in existing neuropsychology and
neurology journals frequently include discussions of the cognitive psychology literature.

Let us take a closer look at the scientific forces that drove this change in disciplinary boundaries. By 1980, both
cognitive psychology and neuropsychology had reached stages of development that were, if not exactly impasses, points
of diminishing returns for the concepts and methods of their own isolated disciplines. In cognitive psychology, the
problem concerned methodologic limitations. By varying stimuli and instructions and measuring responses and response
latencies, cognitive psychologists made inferences about the information processing that intervened between stimulus
and response. But such inferences were indirect, and in some cases they were incapable of distinguishing between
rival theories. In 1978 the cognitive psychologist John Anderson published an influential paper' in which he called this
the "identifiability" problem and took as his example the debate over whether mental images were more like perceptual
representations or linguistic representations. He argued that the field's inability to resolve this issue, despite many
years of research, was due to the impossibility of uniquely identifying internal cognitive processes from stimulus-
response relations. He suggested that the direct study of brain function could, in principle, make a unique
identification possible, but he indicated that such a solution probably lay in the distant future.

That distant future came to pass within the next 10 years, as cognitive psychologists working on a variety of different
topics found that the study of neurologic patients provided a powerful new source of evidence for testing their
theories. In the case of mental imagery, taken by Anderson to be emblematic of the identifiability problem, the
finding that perceptual impairments after brain damage were frequently accompanied by parallel imagery
impairments strongly favored the perceptual hypothesis." The study of learning and memory within cognitive
psychology was revolutionized by the influx of ideas and findings on preserved learning in amnesia, leading
to the hypothesis of multiple memory systems.97-99 In the study of attention, cognitive psychologists had for years
focused on the issue of early versus late attentional selection without achieving a resolution, and here too
neurologic disorders were crucial in moving the field forward. The phenomena of neglect provided dramatic evidence
of selection from spatially formatted perceptual representations, and the variability in neglect's manifestations
from case to case helped to establish the possibility of multiple loci for attentional selection as opposed to a single
early or late locus. The idea of separate visual feature maps, supported by cases of acquired color, motion, and depth
blindness, provided the inspiration for the most novel development in recent cognitive theories of attention—namely,
feature integration theory.loo

What did neuropsychology gain from the rapprochement with cognitive psychology? The main benefits were
theoretical rather than methodologic. Traditionally, neuropsychologists studied the localization and functional
organization of abilities, such as speech, reading, memory, object recognition, and so forth. But few would doubt that
each of these abilities depends upon an orchestrated set of component cognitive processes, and it seems far more
likely that the underlying cognitive components, rather than the task-defined abilities, are what is implemented in
localized neural tissue. The theories of cognitive psychology therefore allowed neuropsychologists to pose questions
about the localization and functional organization of the components of the cognitive architecture, a level of theoretical
analysis that was more likely to yield clear and generalizable findings.

Among patients with reading disorders, for example, some are impaired at reading nonwords (e.g., plif) while others
are impaired at reading irregular words (e.g., yacht). Rather than attempt to localize nonword reading or irregular
word reading per se and delineate them as independent abilities, neuropsychologists have been able to use a theory
of reading developed in cognitive psychology to interpret these disorders in terms of damage to a whole-word
recognition system and a grapheme-to-phoneme translation system, respectively. This interpretation has the
advan¬twe of correctly predicting additional features of patient behavior, such as the tendency to misread
nonwords as words of overall similar appearance when operating with only the whole-word system.
In recent years the neurology and neuropsychology of every major cognitive system has adopted the theoretical
framework of cognitive psychology in a general way, and in some cases specific theories have been
incorporated. This is reflected in the content and organization of the present book. For the most intensively
studied areas of behavioral neurology and neuropsychol¬oev—namely, visual attention, memory, language,
frontal lobe function, and Alzheimer disease—integrated pairs of chapters review the clinical and anatomic
aspects of the relevant disorders and their cognitive theoretical interpretations. Chap¬ters on other topics will
cover both the clinical and theoretical aspects together.

Bibliografie 8.
Leon Dănăilă; Mihai Golu – Tratat de neuropsihologie, volumul 2, p.431-452;
Capitolul XI
PRINCIPII DE ORGANIZARE FUNCȚIONALĂ A
CREIERULUI UMAN

La baza organizării funcţionale a creierului stau anumite


principii, care-i anferă şi-i asigură statutul său de sistem
informaţional logico-computaţional, capabil să efectueze
complexa gamă a operaţiilor pe care le implică discrirninarea
recunoaşterea stimulilor, conceptualizarea (categorizarea),
rezolvarea problemeL elaborarea şi adoptarea deciziilor,
corelarea semnificaţiei factorilor externi cu irrie proprii de
necesitate ale individului, planificarea şi programarea
acţiunilor, reglarea optimă a comportamentului. Esenţiale, din
acest punct de vedere, trebuie considerate următoarele:

1.principiul adaptării reciproce între funcţie şi structură;


2.principiul diferenţierii şi specializării;
3.principiul ierarhizării și integrării sistemice;
Fiecare din aceste principii ne ajută să dezvăluim şi să
înţelegem o anumită a modului de funcţionare a creierului în
calitatea lui de organ (mecanism) al pshicului.

1.PRINCIPIUL ADAPTĂRII RECIPROCE ÎNTRE FUNCȚIE ȘI STRUCTURĂ

Pricipiul adaptării reciproce între funcţie şi structură înscrie creierul în logica evoluţiei biologice generale,
evidenţiind faptul că apariţia sa ca structură materiei vii s-a datorat necesităţii realizării unui nou mod de relaţionare
a organismlui animal cu mediul, pe care-1 numim psihic.

Specificul modului psihic de relaţionare cu mediul rezidă în aceea că actele comportamentale sunt mediate şi
reglate de modele informaţionale interne atât ale organismului însuşi, cât şi ale stimulilor şi condiţiilor externe.
Realizarea unor asemenea modele reclamă structuri adecvate care se vor constitui în cursul evoluţiei organismelor
animale şi vor fi reprezentate de sistemul nervos. Între funcţionalitatea psihică şi structura neuronală se stabileşte
astfel o relaţie congenetică, de interdependenţă şi intercondiţionare reciprocă: forma psihică de relaţionare cu
mediul, potrivit principiului general al biologiei care postulează că funcţia creează organul, a impus apariţia
structurii neuronale ca mecanism adecvat pentru îndeplinirea unei asemenea forme de relaţionare, iar constituirea
structurii neuronale a făcut posibilă realizarea efectivă a psihicului ca modalitate şi capacitate superioară de
adaptare. În filogeneză se constată existenţa unui paralelism legic între nivelul de dezvoltare-organizare a
sistemului nervos şi gradul de complexitate şi eficienţă adaptativă al funcţiilor psihice şi comportamentelor
mediate de ele. Dezvoltarea sistemului nervos s-a produs filogenetic în însuşi cursul funcţionării sale sub influenţa
permanentă a stimulilor şi solicitărilor externe de complexitate crescândă. În ontogeneză, creierul uman îşi
desăvârşeşte organizarea internă tot în cursul funcţionării, prin procesarea fluxurilor informaţionale generate de
modificările de stare ale propriului organişm şi de situaţiile externe, reprezentate nu numai de simplii stimuli fizici
(sunete, lumini, obiecte, plante, animale, persoane umane etc.), ci şi de sarcini cu caracter practic concret şi
cognitiv. În interacţiunea cu obiectele concrete se elaborează schemele funcţionale ale percepţiei, în contact şi
interacţiune cu sarcini de descompunere, asamblare, comparaţie, clasificare, substituţie, calcul, inferenţă etc. se
elaborează schemele operatorii şi reţelele logice care reprezintă mecanismele gândirii.
Organizarea funcţională actuală a creierului uman este cea pe care o reclamă şi presupune nivelul actual al
dezvoltării psihicului uman, în condiţiile mediului socio-cultural actual.

În principiu, din punct de vedere anatomic şi histologic, evoluţia creierului se consideră încheiată odată cu apariţia
lui homo sapiens, dar din punct de vedere funcţional el a rămas deschis la perfecţionare, la creşterea capacităţii
de procesare a informaţiilor, de rezolvare a problemelor şi de creaţie. Se poate presupune că, deşi structurile
anatomice ale creierului rămân aceleaşi în decursul istoriei omenirii, proprietăţile şi capacităţile lor funcţionale se
dezvoltă continuu sub impactul factorilor şi solicitărilor socio-culturale care s-au diversificat şi au sporit în
permanenţă în complexitate. Astfel, creierul omului contemporan se situează, din punct de vedere funcţional, la
un nivel mai înalt decât creierul omului primitiv. Aceasta îşi găseşte expresia şi în faptul că aproape 80% din
numărul total al savanţilor din toate timpurile se concentrează în secolul al XX-lea. De asemenea, se constată şi o
curbă ascendentă a Q.I. (quotientul inteligenţei generale) în succesiunea generaţiilor, ceea ce reflectă creşterea
potenţialului funcţional bazal înnăscut al creierului.

2. PRINCIPIUL DIFERENŢIERII ŞI SPECIALIZĂRII

Principiul diferenţierii şi specializării evidenţiază faptul că creierul nu este o masă omogenă din punct de vedere
structural şi echipotenţială sub raport funcţional. Dimpotrivă, în cursul evoluţiei filogenetice, s-a impus şi s-a
afirmat ca o legitate obiectivă diferenţierea structurală şi specializarea funcţională a elementelor componente.
Sistemul nervos şi reţeaua blocurilor receptoare conexe prezintă o alcătuire celulară eterogenă, diversificată: tipuri
diferite de neuroni şi de elemente sensibile (neuroni unipolari, neuroni bipolari, neuroni multipolari, după
configuraţia terminaţiilor; apoi, neuroni granulari, neuroni stelaţi, neuroni piramidali, după forma somei; la rândul
lor, celulele sensibile-receptoare diferă din punot de vedere morfoanatomic de la un analizator la altul: receptorii
vizuali sunt diferiţi de cei auditivi; cei olfactivi diferă de cei gustativi etc.).

Diferenţierea structurală în interiorul sistemului nervos a fost determinată de necesitatea peifecţionării funcţionale
prin specializare. Astfel, dacă ne referim la scoarţa cerebrală, formaţiunea cea mai complexă a creierului uman,
se poate observa că neuronii granulari (rotunzi şi ovali) s-au specializat în procesarea primară (senzorială) a
informaţiei, ei constituind substratul material al codurilor secvenţial-unidimensionale (corespunzătoare
senzaţiilor); neuronii stelaţi şi piramidali de talie mijlocie s-au specializat în procesarea secundară a informaţiei,
constituind substratul material al codurilor complexe de tip irriagistic şi conceptual-abstract; în fine, neuronii
piramidali, de talie mare, giganţi, s-au specializat în integrarea schemelor şi patternurilor motorii.

În ceea ce priveşte statutul specializării, constatăm că, cea a formaţiunilor subcorticale este integral determinată
genetic, fiind înnăscută: funcţiile biologice, ca, de pildă, circulaţia, respiraţia, digestia, excreţia etc. - se leagă
constant şi invariabil de anumiţi centri nervoşi; de asemenea, o mare parte a componentelor vieţii psihice
inconştiente (instinctele, trebuinţele primare etc.) se leagă de formaţiuni subcorticale bine stabilite; în schimb,
specializarea scoarţei cerebrale este în parte înnăscută, în parte dobândită în ontogeneză. Cea înnăscută se referă
la funcţiile senzoriale primare şi la cele motorii, ambele categorii având prin însuşi programul genetic al speciei
o „localizare" „în zone precis circumscrise topografic", care sunt aceleaşi la toţi oamenii. (De exemplu, la toţi
oamenii funcţia văzului se leagă de aria 17 Brodmann din lobii occipitali.) Specializarea dobândită se referă la
funcţiile psihice superioare, complexe - gândirea, memoria semantică, planificarea, limbajul, autoreglarea
comportamentelor profesionale şi sociale etc. Mecanismele acestor „funcţii" nu sunt date de la naştere, ci se
elaborează treptat în cursul vieţii individului prin învăţare. Configuraţia acestor mecanisme şi gradul lor de
elaborare diferă semnificativ de la un individ la altul. Datele experimentale arată că registrul diferenţelor
interindividuale este mult mai întins în cazul specialiării dobândite, decăt în cazul specializării înnăscute.

Specializarea în interiorul sistemului nervos, al creierului a permis: a) diversificarea şi individualizarea funcţiilor


psihice şi, implicit, a comportamentelor şi ac-tivităţilor adaptative; b) creşterea fineţei şi preciziei procesării
informaţiei, c) sporirea capacităţii combinatorice şi a flexibilităţii în abordarea şi rezolvarea problemelor.

Potrivit principiului specializării, fiecare funcţie psihică presupune şi dispune de un mecanism propriu, cu
individualitate anatomică şi configuraţional-conexională distinctă, a cărui identificare este un obiectiv central al
neuropsihologiei.

Spre deosebire de mecanismul unei funcţii biologice care are o organizare mononivelară, fiind reprezentat printr-
un singur centru nervos (exemplu, centrul respiraţiei situat în bulbul rahidian), mecanismul funcţiei psihice are o
organizare multinivelară, ierarhică, în alcătuirea lui intrând structuri situate la diferite niveluri ale nevraxului
(exemplu, mecanismul senzaţiilor vizuale - de lumină sau de culoare - include: structuri neuronale periferice -
celulele bipolare şi celulele multipolare, structuri neuronale subcorticale, situate în corpii geniculaţi externi din
talamus, şi structuri corticale, reprezentate de aria 17 Brodmann din lobii occipitali). O altă particularitate
importantă a mecanisrnelor proprii funcţiilor psihice constă în combinarea, în cadrul lor, a procesărilor seriale cu
cele paralele, a codificării analogice cu cea discretă, ceea ce le conferă o mare rezistenţă la perturbaţii şi o mare
fiabilitate. În funcţionarea lor ca „sisteme computaţionale" se aplică integral legea lui von Neumann, potrivit
căreia, cu cât numărul transformărilor (operaţiilor) implicate într-un proces este mai mare, cu atât ponderea şi
importanţa unei operaţii singulare oarecare în deterrninarea rezultatului final sunt mai insignifiante. Ca urmare,
în pofida producerii anumitor erori secvenţiale, funcţia psihică integrală continuă să se menţină în limitele
normalităţii, datorită capacităţii de corecţie a mecanismului său neuronal, favorizată de procesarea paralelă a
fluxurilor de informaţie.

În interiorul mecanismelor funcţiilor psihice are loc, de asemenea, o specializare, diferite componente (structuri)
care intră în alcătuirea lor realizând verigi şi operaţii diferite. Astfel, o funcţie psihică pune în evidenţă întotdeauna
o arhitectură internă complexă, ea fiind, în ultimă instanţă, rezultatul unei integrfui emergente a unor operaţii
particulare de procesare sau calcul informaţional efectuate de structuri situate la niveluri diferite, pe verticală, şi
în zone diferite, pe orizontală.

3. PRINCIPIUL IERARHIZĂRII ȘI INTEGRĂRII SISTEMICE

Principiul ierarhizării şi integrării sistemice exprimă o legitate opusă celei preconizate de principiul diferenţierii
şi specializării. Astfel, dacă diferenţierea specializarea au acţionat în direcţia delimitării, individualizării şi
autonomizării structurilor şi formaţiunilor neuronale, ierarhizarea şi integrarea sistemică au acţionat în
direcţia stabilirii de legături şi interacţiuni intre ele, prevenind absolutizarea fărâmiţării şi izolării, transfor-
marea creierului şi, implicit, a vieţii psihice, într-un simplu conglomerat de „elemente" în sine independente.

Ierarhizarea reflectă modul de organizare pe verticală a SNC. Ea presupune stabilirea unor raporturi de
subordonare a segmentelor inferioare celor superioare şi creşterea succesivă a complexităţii operaţiilor de
prelucrare a informaţiei, a actelor reflexe (psihocomportamentale) de la un nivel inferior la altul superior.

Ca principale niveluri ale organizării structural-funcţionale ale SNC pot fi menţionate următoarele:

1. nivelul medular;

2. nivelul trunchiului cerebral;

3. nivelul cerebelos (creierul mic);


4. nivelul diencefalic (formaţiunile talamice şi hipotalamice);

5. nivelul nucleilor bazali;

6. nivelul limbic (formaţiunile sistemului limbic);

7. nivelul cortical (neocortical).

Între toate aceste niveluri ierarhice se realizează o legătură în dublu sens: ascendent; de la nivelurile
inferioare spre cele superioare, şi descendent, de la nivelurile superioare spre cele inferioare. in plan funcţional,
nivelurile superioare coordonează şi controlează activitatea celor inferioare. Nivelul ierarhic cel mai înalt -
neocortexul - devine mecanism central (supraordonat) de comandă-control în raport cu toate celelalte
niveluri.
În interiorul acestei scheme ierarhice, circulaţia informaţiei se realizează după două reguli principale:
regula convergenţei şi regula divergenţei. Prima postulează că, transmiţându-se de jos în sus, fluxul
informaţional suferă succesiv un proces de filtrare şi comprimare, reţinându-se la nivelul inferior dat,
elementele nesemnificative sau de importanţă locală şi expediindu-se nivelului superior cele esenţiale cu
semnificaţie adaptativ-reglatorie generală. Astfel, din volumul uriaş de informaţie captat de receptorii periferici,
la nivelul scoarţei cerebrale, la un moment dat, ajunge doar o mică parte, restul fiind reţinut la nivelurile
subiacente.

Dar nici semnalele care ajung la scoarţă nu au acelaşi „statut": multe nu se integrează în structuri informaţionale
specifice (respectiv în „produse" psihice propriu-zise"), ci acţionează ca factor de activare energetică
nespecifică. Chiar de la nivelul trunchiului cerebral, căile senzoriale ascendente se divid în două grupe
principale: a) căi specifice, care intră în alcătuirea sistemelor analizatorilor, transmiţând informaţia în zonele de
proiecţie corespunzătoare, unde se transformă în conţinuturi ale senzaţiilor şi percepţiilor; b) căi nespecifice,
alcătuind sistemul reticulat activator ascendent (SRAA), care nu transmit semnalele la o adresă anume, ca
cele specifice, ci le distribuie difuz pe suprafaţa întregii scoarţe cerebrale. Efectul acestora se concretizează
într-o stare de activare globală şi apoi locală a scoarţei, optimizând condiţiile fiziologice de fond, necesare
integrării informaţiei în entităţi psihice specifice - cognitive, motivaţionale, afective.

Potrivit regulii divergenţei, transmiţându-se de sus în jos, de la nivelul scoarţei cerebrale spre segmentele
subcorticale apropiate şi îndepărtate, fluxul informaţional este supus succesiv unor operaţii de detaliere,
imbogăţindu-se cu elemente şi date noi la fiecare nivel inferior la care ajunge, asigurându-se corelarea co-
menzilor „centrale" cu cele „locale". Fiecare „mecanism local" se integrează astfel în contextul me-canismului
„reglării globale" a sistemului organismului şi perso-nalităţii: rezolvarea unei sarcini „locale" de reglare se
înscrie ca moment sau verigă în procesul specific de rezolvare a sarcinii de reglare generală, supraordonată
Ierarhizarea este în mod logic dublată de integrare. Prin integrare se înţelege procesul de articulare şi
asimilare a unor entităţi (elemente) iniţial distincte într-un întreg (sistem) supraordonat cu identitate şi
proprietăţi calitative noi, ireductibile la simpla sumă a proprietăţilor părţilor componente.

În funcţionarea creierului, principiul integrării vine să limiteze şi să contrabalanseze acţiunea principiului


diferenţierii şi specializării, prevenind astfet fărâmiţarea excesivă şi izolarea componentelor în interiorul vieţii
psihice şi în sfera comportamentului extern. Ar fi fost, desigur, slab eficientă, sub aspect adaptativ, o organizare
psiho-comportamentală dezarticulată, în care fiecare entitate şi-ar menţine în toate situaţiile autonomia şi
independenţa, neinteracţionând şi nearticulându-se în nici un moment şi în nici un fel cu celelalte. Nu ar fi fost
suficientă nici gruparea simplu sumativă după legile asemănării, contrastului şi contiguităţii spaţio-temporale
formulate de asociaţionism.

Trecerea de la rezolvarea unor sarcini mai simple (exemplu - diferenţierea sunetelor după frecvenţă sau
înălţime) la rezolvarea altora complexe (exemplu. confecţionarea unui obiect, ierarhizarea noţiunilor,
elaborarea raţionamentelor. demonstrarea unor teoreme, descoperirea şi fon-nularea legilor etc.) reclamă nu
doar creşterea numărului „elementelor" psihice interne (al senzaţiilor), dar și integrarea lor în structuri noi,
calitativ superioare, ireductibile la o simplă sumă aritmetică sau un conglomerat.
Bibliografie 9 - Christian Jarett – Great Myths of the Brain, p.1-15;
Introduction
“As humans, we can identify galaxies light years away, we can
study Particles smaller than an atom. But we still haven’t unlocked
the mystery of the three pounds of matter that sits between our
ears.” That was US President Barack Obama speaking in April
2013 at the launch of the multimillion dollar BRAIN Initiative. It
stands for “Brain Research through Advancing Innovative
Neurotechnologies” and the idea is to develop new ways to
visualize the brain in action. The same year the EU announced its
own €1 billion Human Brain Project to create a computer model of
the brain (see p. 105).

This focus on neuroscience isn’t new – back in 1990, US President


George W. Bush designated the 1990s the “Decade of the Brain”
with a series of public awareness publications and events. Since
then interest and investment in neuroscience has only grown more
intense; some have even spoken of the twenty-first century as the
“Century of the Brain.” Despite our passion for all things neuro,
Obama’s assessment of our current knowledge was accurate.
We’ve made great strides in our understanding of the brain, yet
huge mysteries remain. They say a little knowledge can be a
dangerous thing and it is in the context of this excitement and
ignorance that brain myths have thrived. By brain myths I mean
stories and misconceptions about the brain and brain-related
illness, some so entrenched in everyday talk that large sections of
the population see them as taken-for-granted facts.

With so many misconceptions swirling around, it’s increasingly difficult to tell proper neuroscience from brain
mythology or what one science blogger calls neurobollocks (see neurobollocks.wordpress.com), otherwise known as
neurohype, neurobunk, neurotrash, or neurononsense. Daily newspaper headlines tell us the “brain spot” for this or
that emotion has been identified. Salesmen are capitalizing on the fashion for brain science by placing the neuro prefix
in front of any activity you can think of, from neuroleadership to neuromarketing (see p. 188). Fringe therapists and
selfhelp gurus borrow freely from neuroscience jargon, spreading a confusing mix of brain myths and self-
improvement propaganda. In 2014, a journalist and over-enthusiastic neuroscientist even attempted to explain the
Iranian nuclear negotiations (occurring at that time) in terms of basic brain science. Writing in The Atlantic, the authors
actually made some excellent points, especially in terms of historical events and people’s perceptions of fairness. But
they undermined their own credibility by labeling these psychological and historical insights as neuroscience, or by
gratuitously referencing the brain. It’s as if the authors drank brain soup before writing their article, and just as they
were making an interesting historical or political point, they hiccupped out another nonsense neuro reference. This
book takes you on a tour of the most popular, enduring and Dangerous of brain myths and misconceptions, from the
widely accepted notion that we use just 10 percent of our brains, to more specific and harmful misunderstandings
about brain illnesses, such as the mistaken idea that you should place an object in the mouth of a person having an
epileptic fit to stop them from swallowing their tongue. I’ll show you examples of writers, filmmakers, and charlatans
spreading brain myths in newspaper headlines and the latest movies. I’ll investigate the myths’ origins and do my best
to use the latest scientific consensus to explain the truth about how the brain really works.

The Urgent Need for Neuro Myth-Busting

When Sanne Dekker at the Vrije Universiteit in Amsterdam and her colleagues surveyed hundreds of British and
Dutch teachers recently about common brain myths pertaining to education, their results were alarming. The teachers
endorsed around half of 15 neuromyths embedded among 32 statements about the brain.2 What’s more, these weren’t
just any teachers. They were teachers recruited to the survey because they had a particular interest in using
neuroscience to improve teaching. Among the myths the teachers endorsed were the idea that there are left-brain and
right-brain learners and that physical coordination exercises can improve the integration of function between the brain
hemispheres. Worryingly, myths related to quack brain-based teaching were especially likely to be endorsed by the
teachers.Most disconcerting of all, greater general knowledge about the brain was associated with stronger belief in
educational neuromyths – another indication that a little brain knowledge can be a dangerous thing.

If the people educating the next generation are seduced by brain myths, it’s a sure sign that we need to do more to
improve the public’s understanding of the difference between neurobunk and real neuroscience. Still further reason to
tackle brain myths head on comes from research showing that presenting people, including psychology students, with
correct brain information is not enough – many still endorse the 10 percent myth and others. Instead what’s needed is
a “refutational approach” that first details brain myths and then debunks them, which is the format I’ll follow through
much of this book. Patricia Kowalski and Annette Taylor at the University of San Diego compared the two teaching
approaches in a 2009 study with 65 undergraduate psychology students.3 They found that directly refuting brain and
psychology myths, compared with simply presenting accurate facts, significantly improved the students’ performance
on a test of psychology facts and fiction at the end of the semester. Post-semester performance for all students had
improved by 34.3 percent, compared with 53.7 for those taught by the refutational approach.

Yet another reason it’s important we get myth-busting is the media’s treatment of neuroscience. When Cliodhna
O’Connor at UCL’s Division of Psychology and Language Sciences, and her colleagues analyzed UK press coverage
of brain research from 2000 to 2010, they found that newspapers frequently misappropriated new neuroscience
findings to bolster their own agenda, often perpetuating brain myths in the process (we’ll see through examples later
in this book that the US press is guilty of spreading neuromyths too). From analyzing thousands of news articles about
the brain, O’Connor found a frequent habit was for journalists to use a fresh neuroscience finding as the basis for
generating new brain myths – dubious self improvement or parenting advice, say, or an alarmist health warning.
Another theme was using neuroscience to bolster group differences, for example, by referring to “the female brain”
or “the gay brain,” as if all people fitting that identity all have the same kind of brain. “[Neuroscience] research was
being applied out of context to create dramatic headlines, push thinly disguised ideological arguments, or support
particular policy agendas,” O’Connor and her colleagues concluded.

This introductory section ends with a primer on basic brain anatomy, techniques, and terminology. Chapter 1 then
kicks off the myth-busting by providing some historical context, including showing how our understanding of the
brain has evolved since Ancient times, and detailing outdated myths that are no longer widely believed, but which
linger in our proverbs and sayings. This includes the centuries’ long belief that the mind and emotions are located in
the heart – an idea betrayed through contemporary phrases like “heart break” and “learn by heart.” Chapter 2 continues
the historical theme, looking at brain techniques that have entered psychiatric or neurological folklore, such as the
brutal frontal lobotomy. Chapter 3 examines the lives and brains of some of neurosciences mythical figures – including
the nineteenth century rail worker Phineas Gage, who survived an iron rod passing straight through his brain, and
Henry Molaison, the amnesiac who was examined by an estimated 100 psychologists and neuroscientists. Chapter 4
moves on to the classic brain myths that refuse to die away. Many of these will likely be familiar to you – in fact,
maybe you thought they were true. This includes the idea that right-brained people are more creative; that we use just
10 percent of our brains; that women lose their minds when they are pregnant; and that neuroscience is changing
human self-understanding. We’ll see that there is a grain of truth to many of these myths, but that the reality is more
nuanced, and often more fascinating, than the myths suggest. Chapter 5 deals with myths about the physical structure
of the brain, including the idea that bigger means better. And we’ll look at mythology surrounding certain types of
brain cells – the suggestion that mirror neurons are what makes us human and that you have in your brain a cell that
responds only to the thought of your grandmother. Next we turn to technology-related myths about the brain. These
relate to the kind of topical claims that make frequent appearances in the press, including the ubiquitous suggestion
that brain scans can now read your mind, that the Internet is making us stupid, and that computerized brain training
games are making you smart. The penultimate chapter deals with the way the brain relates to the world and the body.
We’ll debunk the popular misconception that there are only five senses, and we’ll also challenge the idea that we
really see the world exactly how it is. The book concludes in Chapter 8 by dealing with the many misconceptions that
exist about brain injury and neurological illness. We’ll see how conditions like epilepsy and amnesia are presented in
Hollywood films and tackle the widespread belief that mood disorders somehow arise from a chemical imbalance in
the brain. The Need for Humility To debunk misconceptions about the brain and present the truth about how the brain
really works, I’ve pored over hundreds of journal articles, consulted the latest reference books and in some cases made
direct contact with the world’s leading experts. I have strived to be as objective as possible, to review the evidence
without a pre-existing agenda. However, anyone who spends time researching brain myths soon discovers that many
of today’s myths were yesterday’s facts. I am presenting you with an account based on the latest contemporary
evidence, but I do so with humility, aware that the facts may change and that people make mistakes. While the
scientific consensus may evolve, what is timeless is to have a skeptical, open-minded approach, to judge claims on
the balance of evidence, and to seek out the truth for its own sake, not in the service of some other agenda. I’ve written
the book in this spirit and in the accompanying box on p. 7 I present you with six tips for applying this skeptical,
empirical approach, to help you spot brain myths for yourself.

Before finishing this Introduction with a primer on basic brain anatomy, I’d like to share with you a contemporary
example of the need for caution and humility in the field of brain mythology. Often myths arise because a single claim
or research finding has particular intuitive appeal. The claim makes sense, it supports a popular argument, and soon it
is cemented as taken-for-granted fact even though its evidence base is weak. This is exactly what happened in recent
years with the popular idea, accepted and spread by many leading neuroscientists, that colorful images from brain
scans are unusually persuasive and beguiling. Yet new evidence suggests this is a modern brain myth. Two researchers
in this area, Martha Farah and Cayce Hook, call this irony the “seductive allure of ‘seductive allure.’” Brain scan
images have been described as seductive since at least the 1990s and today virtually every cultural commentary on
neuroscience mentions the idea that they paralyze our usual powers of rational scrutiny. Consider an otherwise brilliant
essay that psychologist Gary Marcus wrote for the New Yorker late in 2012 about the rise of neuroimaging: “Fancy
color pictures of brains in action became a fixture in media accounts of the human mind and lulled people into
a false sense of comprehension,” he said (emphasis added). Earlier in the year, Steven Poole writing for the New
Statesman put it this way: “the [fMRI] pictures, like religious icons, inspire uncritical devotion.” What’s the
evidence for the seductive power of brain images? It mostly
hinges on two key studies. In 2008, David McCabe and Alan Castel showed that undergraduate participants found the
conclusions of a study (watching TV boosts maths ability) more convincing when accompanied by an fMRI brain
scan image than by a bar chart or an EEG scan.8 The same year, Deena Weisberg and her colleagues published
evidence that naïve adults and neuroscience students found bad psychological explanations more satisfying when they
contained gratuitous neuroscience information (their paper was titled “The Seductive Allure of Neuroscience
Explanations”).

What’s the evidence against the seductive power of brain images? First off, Farah and Hook criticize the 2008 McCabe
study. McCabe’s group claimed that the different image types were “informationally equivalent,” but Farah and Hook
point out this isn’t true – the fMRI brain scan images are unique in providing the specific shape and location of
activation in the temporal lobe, which was relevant information for judging the study. Next came a study published in
2012 by David Gruber and Jacob Dickerson, who found that the presence of brain images did not affect students’
ratings of the credibility of science news stories. Was this failure to replicate the seductive allure of brain scans an
anomaly? Far from it. Through 2013 no fewer than three further investigations found the same or a similar null result.
This included a paper by Hook and Farah themselves,11 involving 988 participants across three experiments; and
another led by Robert Michael involving 10 separate replication attempts and nearly 2000 participants. Overall,
Michael’s team found that the presence of a brain scan had only a tiny effect on people’s belief in an accompanying
story. The result shows “the ‘amazingly persistent meme of the overly influential image’ has been wildly overstated,”
they concluded. So why have so many of us been seduced by the idea that brain scan images are powerfully seductive?
Farah and Hook say the idea supports non-scanning psychologists’ anxieties about brain scan research stealing all the
funding. Perhaps above all, it just seems so plausible. Brain scan images really are rather pretty, and the story that
they have a powerful persuasive effect is very believable. Believable, but quite possibly wrong.
Brain scans may be beautiful but the latest evidence suggests they aren’t as beguiling as we once assumed. It’s a
reminder that in being skeptical about neuroscience we must be careful not to create new brain myths of our own.

Arm Yourself against Neurobunk

This book will guide you through many of the most popular and pervasive neuromyths but more are appearing every
day. To help you tell fact from fiction when encountering brain stories in the news or on TV, here are six simple tips
to follow:
1.Look out for gratuitous neuro references. Just because someone mentions the brain it doesn’t necessarily make
their argument more valid. Writing in The Observer in 2013, clinical neuropsychologist Vaughan Bell called out a
politician who claimed recently that unemployment is a problem because it has “physical effects on the brain,” as if it
isn’t an important enough issue already for social and practical reasons. This is an example of the mistaken idea that
a neurological reference somehow lends greater authority to an argument, or makes a societal or behavioral problem
somehow more real. You’re also likely to encounter newspaper stories that claim a particular product or activity really
is enjoyable or addictive or harmful because of a brain scan study showing the activation of reward pathways or some
other brain change. Anytime someone is trying to convince you of something, ask yourself – does the brain reference
add anything to what we already knew? Does it really make the argument more truthful?

2.Look for conflicts of interest. Many of the most outrageous and farfetched brain stories are spread by people with
an agenda. Perhaps they have a book to sell or they’re marketing a new form of training or therapy. A common tactic
used by these people is to invoke the brain to shore up their claims. Popular themes include the idea that technology
or other aspects of modern life are changing the brain in a harmful way, or the opposite – that some new form of
training or therapy leads to real, permanent beneficial brain changes. Often these kinds of brain claims are mere
conjecture, sometimes even from the mouths of neuroscientists or psychologists speaking outside their own area of
specialism. Look for independent opinion from experts who don’t have a vested interest. And check whether
brain claims are backed by quality peer-reviewed evidence (see point 5). Most science journals require authors to
declare conflicts of interest so check for this at the end of relevant published papers.

3.Watch out for grandiose claims. No Lie MRI is a US company that offers brain scan-based lie detection services.
Its home page states, “The technology used by No Lie MRI represents the first and only direct measure of truth
verification and lie detection in human history!” Sound too good to be true? If it does, it probably is.

Words like “revolutionary,” “permanent,” “first ever,” “unlock,” “hidden,” “within seconds,” should all set
alarm bells ringing when uttered in relation to the brain. One check you can perform is to look at the career of the
person making the claims. If they say they’ve developed a revolutionary new brain technique that will for the first
time unlock your hidden potential within seconds, ask yourself why they haven’t applied it to themselves and become
a best-selling artist, Nobel winning scientist, or Olympic athlete.

4.Beware of seductive metaphors. We’d all like to have balance and calm in our lives but this abstract sense of
balance has nothing to do with the literal balance of activity across the two brain hemispheres (see also p. 196) or
other levels of neural function. This doesn’t stop some self-help gurus invoking concepts like “hemispheric balance”
so as to lend a scientific sheen to their lifestyle tips – as if the route to balanced work schedules is having a balanced
brain. Any time that someone attempts to link a metaphorical concept (e.g. deep thinking) with actual brain activity
(e.g. in deep brain areas), it’s highly likely they’re talking rubbish. Also, beware references to completely made up
brain areas. In February 2013, for instance, the Daily Mail reported on research by a German neurologist who they
said had discovered a tell-tale “dark patch” in the “central lobe” of the brains of killers and rapists. The thing is, there
is no such thing as a central lobe!

5.Learn to recognize quality research. Ignore spin and take first-hand testimonials with a pinch of salt. When it
comes to testing the efficacy of brain-based interventions, the gold standard is the randomized, double-blind,
placebo-controlled trial. This means the recipients of the intervention don’t know whether they’ve received the target
intervention or a placebo (a form of inert treatment such as a sugar pill), and the researchers also don’t know who’s
been allocated to which condition. This helps stop motivation, expectation, and bias from creeping into the results.
Related to this, it’s important for the control group to do something that appears like a real intervention, even though
it isn’t. Many trials fail to ensure this is the case. The most robust evidence to look for in relation to brain claims is
the meta-analysis, so try to search for these if you can. They weigh up all the evidence from existing trials in a given
area and help provide an accurate picture of whether a treatment really works or whether a stated difference really
exists.

6.Recognize the difference between causation and correlation (a point I’ll come back to in relation to mirror
neurons in Chapter 5). Many newspaper stories about brain findings actually refer to correlational studies that only
show a single snapshot in time. “People who do more of activity X have a larger brain area Y,” the story might say.
But if the study was correlational we don’t know that the activity caused the larger brain area. The causal direction
could run the other way (people with a larger Y like to do activity X), or some other factor might influence both X
and Y. Trustworthy scientific articles or news stories should draw attention to this limitation and any others. Indeed,
authors who only focus on the evidence that supports their initial hypotheses or beliefs are falling prey to what’s
known as “confirmation bias.” This is a very human tendency, but it’s one that scrupulous scientists and journalists
should deliberately work against in the pursuit of the truth.

Arming yourself with these six tips will help you tell the difference between a genuine neuroscientist and a charlatan,
and between a considered brain-based news story and hype. If you’re still unsure about a recent development, you
could always look to see if any of the following entertaining expert skeptical bloggers have shared their views: www.
mindhacks.com; http://blogs.discovermagazine.com/neuroskeptic/; http:// neurocritic.blogspot.co.uk;
http://neurobollocks.wordpress.com; http:// neurobonkers.com. And check out my own WIRED neuroscience blog
www.wired.com/wiredscience/brainwatch/

A Primer on Basic Brain Anatomy, Techniques, and Terminology

Hold a human brain in your hands and the first thing you notice is its impressive heaviness. Weighing about three
pounds, the brain feels dense. You also see immediately that there is a distinct groove – the longitudinal fissure –
running front to back and dividing the brain into two halves known as hemispheres. Deep within the brain, the two
hemispheres are joined by the corpus callosum, a thick bundle of connective fibers. The spongy, visible outer layer
of the hemispheres – the cerebral cortex (meaning literally rind or bark) – has a crinkled appearance: a swathe of
swirling hills and valleys, referred to anatomically as gyri and sulci, respectively. The cortex is divided into five
distinct lobes: the frontal lobe, the parietal lobe near the crown of the head, the two temporal lobes at each side near
the ears, and the occipital lobe at the rear. Each lobe is associated with particular domains of mental function. For
instance, the frontal lobe is known to be important for self-control and movement; the parietal lobe for processing
touch and controlling attention; and the occipital lobe is involved in early visual processing. The extent to which
mental functions are localized to specific brain regions has been a matter of debate throughout neurological history
and continues to this day.

Hanging off the back of the brain is the cauliflower-like cerebellum, which almost looks like another mini-brain
(in fact cerebellum means“little brain”). It too is made up of two distinct hemispheres, and remarkably it contains
around half of the neurons in the central nervous system despite constituting just 10 percent of the brain’s volume.
Traditionally the cerebellum was associated only with learning and motor control (i.e. control of the body’s
movements), but today it is known to be involved in many functions, including emotion, language, pain, and memory.
Holding the brain aloft to study its underside, you see the brain stem sprouting downwards, which would normally
be connected to the spinal cord. The brain stem also projects upwards into the interior of the brain to a point
approximately level with the eyes. Containing distinct regions such as the medulla and pons, the brain stem is
associated with basic life support functions, including control of breathing and heart rate. Reflexes like sneezing and
vomiting are also controlled here. Some commentators refer to the brain stem as “the lizard brain” but this is a
misnomer. Slice the brain into two to study the inner anatomy and you discover that there are a series of fluid-filled
hollows, known as ventricles, which act as a shock-absorption system. You can also see the midbrain that sits atop
the brainstem and plays a part in functions such as eye movements. Above and anterior to the midbrain is the thalamus
– a vital relay station that receives connections from, and connects to, many other brain areas. Underneath the thalamus
is the hypothalamus and pituitary gland, which are involved in the release of hormones and the regulation of basic
needs such as hunger and sexual desire.

Also buried deep in the brain and connected to the thalamus are the hornlike basal ganglia, which are involved
in learning, emotions, and the control of movement. Nearby we also find, one on each side of the brain, the
hippocampi (singular hippocampus) – the Greek name for “sea-horse” for that is what early anatomists believed it
resembled. Here too are the almond shaped amygdala, again one on each side. The hippocampus plays a vital role
in memory and the amygdala is important for memory and learning, especially when emotions are involved. The
collective name for the hippocampus, amygdala, and related parts of the cortex is the limbic system, which is an
important functional network for the emotions.

The brain’s awesome complexity is largely invisible to the naked eye.Within its spongy bulk are approximately 86
billion neurons forming a staggering 100 trillion plus connections (see Plate 4). There are also a similar number of
glial cells, which recent research suggests are more than housekeepers, as used to be believed, but also involved in
information processing. However, we should be careful not to get too reverential about the brain’s construction – it’s
not a perfect design by any means (more about this on p. 135). In the cortex, neurons are arranged into layers, each
containing different types and density of neuron. The popular term for brains – “gray matter” – comes from the
anatomical name for tissue that is mostly made up of neuronal cell bodies. The cerebral cortex is entirely made up of
gray matter, although it looks more pinkish than gray, at least when fresh. This is in contrast to “white matter” –
found in abundance beneath the cortex – which is tissue made up mostly of fat-covered neuronal axons (axons are a
tendrillike part of the neuron that is important for communicating with other neurons). It is the fat-covered axons that
give rise to the whitish appearance of white matter. Neurons communicate with each other across small gaps called
synapses. This is where a chemical messenger (a “neurotransmitter”) is released at the end of the axon of one neuron,
and then absorbed into the dendrite (a branch-like structure) of a receiving neuron. Neurons release neurotransmitters
in this way when they are sufficiently excited by other neurons. Enough excitation causes an “action potential,”
which is when a spike of electrical activity passes the length of the neuron, eventually leading it to release
neurotransmitters. In turn these neurotransmitters can excite or inhibit receiving neurons. They can also cause slower,
longer-lasting changes, for example by altering gene function in the receiving neuron. Traditionally, insight into the
function of different neural areas was derived from research on brain damaged patients. Significant advances were
made in this way in the nineteenth century, such as the observation that, in most people, language function is
dominated by the left hemisphere. Some patients, such as the railway worker Phineas Gage, have had a particularly
influential effect on the field. The study of particular associations of impairment and brain damage also remains an
important line of brain research to this day. A major difference between modern and historic research of this kind is
that today we can use medical scanning to identify where the brain has been damaged. Before such technology was
available, researchers had to wait until a person had died to perform an autopsy.

Modern brain imaging methods are used not only to examine the structure of the brain, but also to watch how it
functions. It is in our understanding of brain function that the most exciting findings and controversies are emerging
in modern neuroscience. Today the method used most widely in research of this kind, involving patients and healthy
people, is called functional magnetic resonance imaging (fMRI). The technique exploits the fact that blood is more
oxygenated in highly active parts of the brain. By comparing changes to the oxygenation of the blood throughout the
brain, fMRI can be used to visualize which brain areas are working harder than others. Furthermore, by carefully
monitoring such changes while participants perform controlled tasks in the brain scanner, fMRI can help build a map
of what parts of the brain are involved in different mental functions. Other forms of brain scanning include Positron
Emission Tomography (PET) and Single-Photon Computed Tomography, both of which involve injecting the
patient or research participant with a radioactive isotope. Yet another form of imaging called Diffusion Tensor
Imaging (DTI) is based on the passage of water molecules through neural tissue and is used to map the brain’s
connective pathways. DTI produces beautifully complex, colorful wiring diagrams.

The Human Connectome Project, launched in 2009, aims to map all 600 trillion wires in the human brain. An older
brain imaging technique, first used with humans in the 1920s, is electroencephalography (EEG), which involves
monitoring waves of electrical activity via electrodes placed on the scalp. The technique is still used widely in hospitals
and research labs today. The spatial resolution is poor compared with more modern methods such as fMRI, but an
advantage is that fluctuations in activity can be detected at the level of milliseconds (versus seconds for fMRI). A
more recently developed technique that shares the high temporal resolution of EEG is known as
magnetoencephalography (MEG), but it too suffers from a lack of spatial resolution. Brain imaging is not the only
way that contemporary researchers investigate the human brain. Another approach that’s increased hugely in
popularity in recent years is known as transcranial magnetic stimulation (TMS). It involves placing a magnetic coil
over a region of the head, which has the effect of temporarily disrupting neural activity in brain areas beneath that
spot. This method can be used to create what’s called a “virtual lesion” in the brain. This way, researchers can
temporarily knock out functioning in a specific brain area and then look to see what effect this has on mental
functioning. Whereas fMRI shows where brain activity correlates with mental function, TMS has the advantage of
being able to show whether activity in a particular area is necessary for that mental functioning to occur. The
techniques I’ve mentioned so far can all be used in humans and animals. There is also a great deal of brain research
that is only (or most often) conducted in animals. This research involves techniques that are usually deemed too
invasive for humans. For example, a significant amount of research with monkeys and other nonhuman primates
involves inserting electrodes into the brain and recording the activity directly from specific neurons (called single-cell
recording). Only rarely is this approach used with humans, for example, during neurosurgery for severe epilepsy. The
direct insertion of electrodes and cannulas into animal brains can also be used to monitor and alter levels of brain
chemicals at highly localized sites. Another ground-breaking technique that’s currently used in animal research is
known as optogenetics. Named 2010 “method of the year” by the journal Nature Methods, optogenetics involves
inserting light-sensitive genes into neurons. These individual neurons can then be switched on and off by exposing
them to different colors of light. New methods for investigating the brain are being developed all the time, and
innovations in the field will accelerate in the next few years thanks to the launch of the US BRAIN Initiative and the
EU Human Brain Project. As I was putting the finishing touches to this book, the White House announced a proposal
to double its investment in the BRAIN Initiative “from about $100 million in FY [financial year] 2014 to
approximately $200 million in FY 2015.”

S-ar putea să vă placă și