alteraciones neuropsicolÓgicas en consumidores de cocaÍna …hera.ugr.es/tesisugr/21168581.pdf ·...

221
Laura Moreno López ALTERACIONES NEUROPSICOLÓGICAS EN CONSUMIDORES DE COCAÍNA: CORRELATOS NEUROANATÓMICOS

Upload: others

Post on 13-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Laura Moreno López

ALTERACIONES NEUROPSICOLÓGICAS EN CONSUMIDORES DE COCAÍNA:

CORRELATOS NEUROANATÓMICOS

UNIVERSIDAD DE GRANADA

Departamento de Personalidad, Evaluación y Tratamiento Psicológico

TESIS DOCTORAL

ALTERACIONES NEUROPSICOLÓGICAS EN CONSUMIDORES DE COCAÍNA:

CORRELATOS NEUROANATÓMICOS

Doctoranda: Laura Moreno López

Directores: Dr. Antonio Javier Verdejo García y Dr. Miguel Pérez García

Editor: Editorial de la Universidad de GranadaAutor: Laura Moreno LópezD.L.: GR 1203-2013ISBN: 978-84-9028-529-9

Los directores Dr. Antonio Javier Verdejo García y Dr. Miguel Pérez García

autorizan la presentación de la tesis doctoral titulada: “Alteraciones Neuropsicológicas en

consumidores de cocaína: Correlatos neuroanatómicos” presentada por Dña. Laura

Moreno López.

Fdo. Dr. Antonio Javier Verdejo García Fdo. Dr. Miguel Pérez García

Fdo. Laura Moreno López

A mis padres

he future belongs to those who believe in the beauty of their dreams

(Eleanor Roosevelt)

T

Agradecimientos

Quiero agradecer la realización de este trabajo a todos aquellos que lo han

hecho posible y muy especialmente a mis tutores Antonio Verdejo García y Miguel

Pérez García, por su apoyo y animo constantes y su ayuda incondicional.

A Emmanuel Stamatakis y Maki Kasahara por estar conmigo en los buenos y

malos momentos de mi primera estancia y por ayudarme en todo lo que necesité y he

necesitado hasta el día de hoy. Siempre seréis lo mejor de Cambridge para mí.

A Rita Goldstein, Nelly Alia-Klein, Tom Maloney, Patricia Woicik, Anna Konova

y Muhammad Parvaz del Departamento de Medicina del laboratorio Brookhaven de

Nueva York y a Martin Paulus y Scott Mackey del Departamento de Psiquiatría de la

Universidad de California San Diego, por darme la oportunidad de trabajar con ellos y

aprender un poquito más de este apasionante campo.

Asimismo, no quiero olvidarme de Carles Soriano Más por su apoyo y su ayuda

constantes, de todos mis compañeros del grupo de Neuropsicología y

Neuroinmunología clínica y de todos los participantes voluntarios de mis estudios.

Y por último, y en mayúsculas, a aquellos sin los que nada de esto tendría

sentido, mis padres, mi hermano y Jorge.

[ÍNDICE]

Presentación 1 I. INTRODUCCIÓN 5 1. Introducción 6 2. Exploración de los factores neurobiológicos y neuropsicológicos implicados en la 8 adicción: Neuroimagen y neuropsicología de la impulsividad y las funciones ejecutivas

2.1. Neuroimagen y adicciones 8

2.1.1. Clasificación y características de las técnicas de neuroimagen 9

2.1.1.1. Resonancia magnética estructural 9 2.1.1.2. Tomografía por emisión de positrones 10

2.1.2. Hallazgos de neuroimagen en consumidores de distintas drogas 11

2.1.2.1. Cocaína 12 2.1.2.2. Heroína 13 2.1.2.3. Alcohol 14 2.1.2.4. Éxtasis (MDMA) 16 2.1.2.5. Cannabis 17

2.2. Impulsividad y adicciones 18

2.2.1. Instrumentos de evaluación de la impulsividad 19 2.2.1.1. Inventarios de personalidad impulsiva 19 2.2.1.2. Medidas de control inhibitorio 20 2.2.2. Hallazgos relacionados con la impulsividad en consumidores 21 de distintas drogas 2.2.2.1. Cocaína 21 2.2.2.2. Heroína 23 2.2.2.3. Alcohol 23 2.2.2.4. Éxtasis (MDMA) 24 2.2.2.5. Cannabis 25 2.3. Funciones ejecutivas y adicciones 25 2.3.1. Evaluación neuropsicológica de las funciones ejecutivas 26 2.3.1.1. Actualización 27 2.3.1.2. Control inhibitorio 28 2.3.1.3. Flexibilidad cognitiva 28 2.3.1.4. Toma de decisiones 28 2.3.1.5. Procesos emocionales 29 2.3.2. Hallazgos neuropsicológicos relacionados con las funciones 30 ejecutivas en consumidores de distintas drogas 2.3.2.1. Cocaína 31 2.3.2.2. Heroína 32 2.3.2.3. Alcohol 32 2.3.2.4. Éxtasis (MDMA) 33 2.3.2.5. Cannabis 34

II. JUSTIFICACIÓN Y OBJETIVOS 35 III. MEMORIA DE TRABAJOS 41 Artículo 1. Trait impulsivity and prefrontal gray matter reductions in cocaine 43 dependent individuals 1. Introduction 44 2. Materials and methods 46

2.1. Participants 46 2.2. Instruments and assessment procedures 47 2.2.1. Patterns of drug use 47 2.2.2. Trait impulsivity 49 2.3. MRI adquisition 49 2.4. Image analysis 50 2.4.1. Global effects of patterns of drug use 50 2.4.2. Regional GM and WM differences between cocaine dependent 51 individuals and non-drug using controls 2.4.3. Regional relationships between GM and WM and measures of 51 impulsivity and estimates of drug use 3. Results 51

3.1. Participants’ characteristics 51 3.2. Trait impulsivity 52 3.3. Imaging analysis 52 3.3.1. Global effects 52 3.3.2. Regional GM and WM differences between cocaine dependent 53 individuals and non-drug using controls 3.3.3. Regional relationships between GM and WM and measures of 56 impulsivity and estimates of drug use 4. Discussion 58 References 62 Artículo 2: Neural Correlates of the Severity of Cocaine, Heroin, Alcohol, 67 MDMA and Cannabis Use in Polysubstance Abusers: A Resting-PET Brain Metabolism Study 1. Introduction 68 2. Methods 69

2.1 Participants 69

2.2 Testing protocols and procedures 70 2.3 Tools 70

2.3.1 Drug Use Information 70

2.3.2 PET Image Acquisition 72

2.4 Data analysis 72

2.4.1 Preprocessing of PET images 72

2.4.2 Statistical Analysis 72

3. Results 73

3.1 Amount of use 73

3.2 Duration of use 74

4. Discussion 77 References 81 Artículo 3: Neural correlates of hot and cold executive functions in 85 polysubstance addiction: Association between neuropsychological performance and resting-PET brain metabolism 1. Introduction 86 2. Methods 88

2.1. Participants 88

2.2. Testing protocols and procedures 90

2.3. Instruments 90

2.3.1. Patterns of drug use 90

2.3.2. Neuropsychological tests 91

2.3.3. PET image acquisition 92

2.4. Data analysis 92

2.4.1. Preprocessing of PET images 92

2.4.2. ROI analysis 92

2.4.3. Statistical analyses 93

2.4.3.1. Behavioral analysis 93

2.4.3.2. ROI analysis 93

2.4.3.3. Voxel based analysis 93

3. Results 94

3.1. Neuropsychological performance 94

3.2. ROI analyses 95

3.2.1. Association between neuropsychological performance and ROIs 95

3.2.2. Association between the ROIs associated with 95 neuropsychological performance and estimates of drug use

3.3. Voxel-based analyses 96

3.3.1. Cold executive functions –whole brain voxel-based correlates 98 in SDI

3.3.2. Hot executive functions –whole brain voxel-based correlates 100 in SDI

3.3.3. The relationship between voxel-based derived regions 102 associated with neuropsychological performance and estimates of drug use

4. Discussion 102 References 107 Supplementary material 116

IV. DISCUSIÓN 121 1. Discusión general 122 2. Conclusiones 128 3. Perspectivas de futuro 129 DOCTORADO EUROPEO 131 1. Summary 132 2. Conclusions 136 3. Future perspectives 137 REFERENCIAS 139 ANEXOS 171

P á g i n a | 1

PRESENTACIÓN [ ]

P á g i n a | 2

El informe mundial sobre drogas emitido por Naciones Unidas en 2011 (UNODC,

2011) estimó que en 2009, entre 149 y 272 millones de personas con edades

comprendidas entre los 15 y los 64 años habría consumido drogas ilegales en al menos

una ocasión durante el último año, al menos la mitad lo habría hecho durante el último

mes y entre 15 y 39 millones de personas tendrían problemas con ese consumo.

A nivel mundial, el cannabis es la droga ilegal más utilizada seguida por las

anfetaminas, los opiáceos y la cocaína. Sin embargo, el patrón de consumo es diferente en

Europa, donde la cocaína es la segunda droga ilegal más consumida seguida por las

anfetaminas, el éxtasis y los opiáceos. La prevalencia del consumo de cocaína difiere entre

los países que conforman la Unión Europea pero es especialmente problemática entre

varones jóvenes de entre 15 y 34 años y en países como Dinamarca, Irlanda, Italia, España

y Reino Unido. Y es que, aunque la prevalencia del consumo es menor en la Unión Europea

(2,1%) que en el mismo colectivo en Australia (3,4%), Canadá (3,3%) o Estados Unidos

(4,1%), en España (4,4%) y Reino Unido (4,8%) ha presentado cifras más elevadas

(EMCDDA, 2011).

Aunque el consumo de drogas ilegales constituye uno de los mayores problemas de

nuestra sociedad, no debemos olvidar el efecto que el abuso o co-abuso de otras sustancias

legales como el alcohol tiene en la población. Según la organización mundial de la salud, el

alcoholismo es una de las drogodependencias más extendidas a nivel mundial y constituye

el tercer lugar entre los principales factores de riesgo de muerte prematura y discapacidad

en el mundo. En nuestra sociedad, el consumo de alcohol es una práctica profundamente

arraigada y se estima que la mayoría de la población consume bebidas alcohólicas

esporádica o habitualmente (OED, 2009). Según el último informe del Observatorio

Español sobre Drogas, el consumo de alcohol parece haberse establecido e incluso ha

descendido en algunos sectores de la población, pero en los últimos años, el consumo

entre los jóvenes, un consumo caracterizado por tomar grandes cantidades de alcohol

P á g i n a | 3

durante breves periodos de tiempo (el denominado binge drinking), que ha empezado a

asociarse con la presencia de alteraciones neuropsicológicas en estas poblaciones (Parada

et al., 2012; Squeglia et al., 2012), está alcanzando cifras alarmantes.

Además de sus evidentes repercusiones sobre la salud física y psicológica de los

consumidores, el consumo de drogas tiene importantes implicaciones familiares, sociales,

culturales, políticas y económicas, por lo que constituye un fenómeno de enorme

complejidad, pero también de relevancia equiparable. Como consecuencia de su

complejidad, este fenómeno puede ser abordado desde múltiples perspectivas científicas.

En este sentido, diversas fuentes de evidencia, incluyendo estudios preclínicos en

animales, estudios farmacológicos, neuropsicológicos y de neuroimagen cerebral, han

destacado la relevancia de las alteraciones cognitivas y emocionales encontradas en

individuos drogodependientes o con alto riesgo de iniciarse en el consumo. Especialmente,

ha aumento el interés por conocer los correlatos neuropsicológicos y neuroanatómicos del

consumo de distintas drogas, así como la importancia que ciertos rasgos de personalidad

como la impulsividad pueden tener en el inicio, la progresión del consumo recreativo a la

dependencia y la recaída en estos pacientes (Verdejo-García, Lawrence, & Clark, 2008).

En los últimos años, el extraordinario desarrollo de las técnicas de neuroimagen

cerebral ha permitido investigar con mayor precisión la naturaleza, localización y

extensión de dichas alteraciones y en muchas ocasiones, se ha mostrado como la

herramienta más eficaz para detectar alteraciones a nivel cerebral en consumidores de

ciertas sustancias psicoactivas (p.e. Barrós-Loscertales et al., 2011a; Tapert et al., 2007).

Sin embargo, la presencia de ciertas limitaciones inherentes al estudio del fenómeno de la

drogadicción ha llevado a resultados inconsistentes y en muchas ocasiones

contradictorios.

En el contexto de aportar conocimientos sobre esta temática y superar las

limitaciones existentes en estudios previos de adicción, la presente Tesis Doctoral tuvo

P á g i n a | 4

como objetivo general el estudio de las alteraciones cerebrales estructurales y funcionales

presentes en individuos policonsumidores de drogas y (i) las variables de personalidad

que incrementan la predisposición al consumo y la dependencia, (ii) las estimaciones de

cantidad y duración de consumo de diferentes drogas, y (iii) el funcionamiento

neuropsicológico de los procesos ejecutivos típicamente afectados por el consumo de

drogas.

P á g i n a | 5

INTRODUCCIÓN [ ]

P á g i n a | 6

1. Introducción

Durante la primera mitad del siglo XX, la adicción fue despreciada en su carácter de

enfermedad, siendo considerada fundamentalmente un problema de índole moral o

antisocial. Los modelos clásicos de la adicción habían enfatizado el papel del llamado

“circuito de la recompensa” o “del placer” (el circuito mesolímbico dopaminérgico) y se

defendía que las drogas se consumían porque eran reforzadores potentes y, por tanto,

placenteras. Los trastornos adictivos eran clasificados en las primeras ediciones del

Manual Diagnóstico y Estadístico de los Trastornos Mentales (DSM) y la Clasificación

Internacional de Enfermedades (CIE) como trastornos de la personalidad o “trastornos de

carácter, de conducta y de inteligencia”.

Efectivamente, los primeros contactos con las drogas producen efectos placenteros

en las personas que las consumen y durante mucho tiempo, estos efectos han sido

asociados con la activación del circuito mesolímbico dopaminérgico. Utilizando los

mismos mecanismos fisiológicos que los reforzadores naturales, las drogas actúan sobre

este sistema pero a distintos niveles (Nestler, 2005). Así, la cocaína, las anfetaminas y la

nicotina producirían la liberación de dopamina en el núcleo accumbens, los opiáceos

producirían la activación de los receptores de péptidos opioides en el área tegmental

ventral y el núcleo accumbens y el alcohol produciría la activación de sistemas GABA en el

núcleo accumbens y la amígdala.

Sin embargo, a medida que avanza el consumo de la sustancia, se producen un

conjunto de neuroadaptaciones en las regiones asociadas con los efectos reforzantes de las

drogas y en aquellas conectas neuroanatomicamente con estas que conllevan el paso del

consumo de drogas por sus efectos reforzantes, al consumo de estas de forma impulsiva y

finalmente compulsiva, el consumo crónico y la recaída. Esta transición implica la

reprogramación de circuitos neuronales implicados entre otros en el refuerzo y la

motivación, la memoria, el funcionamiento ejecutivo y la regulación emocional. De entre

P á g i n a | 7

los diferentes circuitos implicados, los más estudiados en los últimos años se localizan en

las áreas prefrontales y se han propuesto como bases anatómicas del control de impulsos,

la toma de decisiones y la regulación emocional (Goldstein & Volkow, 2002, 2011).

En la actualidad, la adicción se define como un trastorno crónico y recidivante

caracterizado por un consumo de drogas abusivo y persistente a pesar de sus crecientes

consecuencias negativas para la vida de la persona (DSM-IV). La clasificación CIE-10

introduce el matiz de que la adicción tiene un carácter compulsivo y se caracteriza por la

falta de control por parte del individuo.

Definidas de esta manera, las alteraciones que caracterizan la adicción tienen

importantes correlatos neuropsicológicos al nivel de una presumible disfunción de las

habilidades encargadas de organizar y programar conductas dirigidas a objetivos y tomar

decisiones adaptativas. En este sentido, las alteraciones neuropsicológicas pueden

contribuir significativamente al consumo y la adicción a través de al menos dos

mecanismos (Rogers & Robbins, 2001). En primer lugar, la presencia de alteraciones

neuropsicológicas puede incrementar la probabilidad de conductas de búsqueda y

consumo de drogas tanto en las fases iniciales del consumo como en las recaídas y facilitar

el paso de un consumo recreativo a la dependencia (George & Koob, 2010; Verdejo-García

et al., 2008). En segundo lugar, la existencia de déficits neuropsicológicos puede limitar o

interferir la capacidad de los individuos drogodependientes para asimilar los contenidos y

las actividades de los programas de rehabilitación (Aharonovich, Amrhein, Bisaga, Nunes,

& Hasin, 2008; Streeter et al., 2008; Turner, LaRowe, Horner, Herron, & Malcolm, 2009).

De acuerdo con su relevancia clínica y teórica, el estudio de las alteraciones

neuropsicológicas asociadas al consumo de drogas y su relación con los mecanismos

cerebrales implicados en la adicción ha experimentado un considerable avance en los

últimos años, con importantes aportaciones empíricas derivadas de modelos animales,

estudios farmacológicos y estudios de neuroimagen cerebral. El desarrollo y la aplicación

P á g i n a | 8

de las técnicas de neuroimagen se ha revelado como una potente herramienta de carácter

transversal (que puede emplearse en estudios con animales, y en estudios de farmacología

clínica, neuropsicología y tratamiento en humanos) para mejorar nuestra comprensión de

los procesos adictivos.

2. Exploración de los factores neurobiológicos y neuropsicológicos implicados en la

adicción: Neuroimagen y neuropsicología de la impulsividad y las funciones

ejecutivas

A continuación se presenta una breve revisión de las principales modalidades de

exploración neurobiológica y neurocognitiva en el estudio de los procesos de impulsividad

y funcionamiento ejecutivo en el campo de las adicciones. En cada apartado se presentará,

en primer lugar, un breve resumen de las principales técnicas utilizadas y, en segundo

lugar, un resumen de los principales hallazgos resultantes de la aplicación de las mismas

en consumidores de cocaína, heroína, alcohol, MDMA y cannabis (las drogas que

presentaron una mayor prevalencia de consumo entre los participantes de nuestros

estudios). Nos centraremos en los resultados del consumo crónico de las sustancias,

evitando los estudios de sus efectos agudos o de aquellos estudios que se han llevado a

cabo tras breves periodos de abstinencia (<48h), pues son las alteraciones a largo plazo las

que tienen un impacto más directo y significativo sobre la rehabilitación y el

funcionamiento diario de los individuos drogodependientes, incluso una vez abandonado

el consumo de drogas (Verdejo-García, López-Torrecillas, Giménez, & Pérez-García, 2004).

2.1. Neuroimagen y adicciones

Las técnicas de imagen cerebral se han convertido en un pilar fundamental para el

desarrollo de las neurociencias, permitido en los últimos años examinar “in vivo” los

efectos agudos de la administración de sustancias psicoactivas, los correlatos cerebrales

del deseo intenso por consumir las drogas (craving), o las alteraciones a largo plazo en

regiones cerebrales y sistemas neuroquímicos implicados en el consumo crónico y la

P á g i n a | 9

dependencia. Desde los inicios de su aplicación, y en un breve periodo de tiempo, los

hallazgos de neuroimagen cerebral han contribuido de manera significativa a comprender

los sustratos cerebrales de las drogodependencias y sus repercusiones sobre el

funcionamiento neuropsicológico de los consumidores.

2.1.1. Clasificación y características de las técnicas de neuroimagen

En función de su aplicación, las técnicas de neuroimagen pueden ser agrupadas en

dos grandes categorías: técnicas estructurales y técnicas funcionales. Las técnicas

estructurales informan sobre la localización, forma y tamaño de algunas regiones

cerebrales y permiten cuantificar los cambios volumétricos o de densidad de la sustancia

gris y la sustancia blanca cerebral. Las técnicas funcionales miden los cambios en la

actividad, el metabolismo cerebral o en ciertos parámetros neurofarmacológicos como la

densidad de receptores o los niveles de neurotransmisores y metabolitos.

Las técnicas estructurales más utilizadas en la investigación en drogodependencias

son la resonancia magnética estructural y las imágenes por tensor de difusión. Las técnicas

funcionales más utilizadas son la resonancia magnética funcional, la tomografía por

emisión de positrones (PET) y la tomografía por emisión de fotón único (SPECT). Aunque

el coste del SPECT es menor, el uso de la tomografía por emisión de positrones presenta

importantes ventajas con respecto a este, entre las que cabe destacar una localización más

precisa del trazador y una mejor resolución espacial.

2.1.1.1. Resonancia magnética estructural

Para que podamos obtener una imagen del interior de un objeto, sin que esto

conlleve la rotura o descomposición de su estructura interna, es necesario hacer llegar

hasta él ondas con forma conocida, ya sea por reflejo (ultrasonidos), por absorción y

transmisión o por emisión desde el interior, e intentar captar la modificación que se ha

producido en esa onda ya sea en su fase, en su frecuencia o en su amplitud. El tipo de

P á g i n a | 1 0

ondas que se emplean en resonancia son ondas electromagnéticas a frecuencias de radio

del orden de los megahertzios y los receptores y posteriormente los emisores del interior

del cuerpo humano son los protones de algunos núcleos atómicos que hacen de antena

emisora y receptora. El escáner de resonancia magnética consta de un campo magnético

muy intenso, homogéneo y uniforme. Cuando situamos a la persona bajo la influencia de

ese campo, los protones de sus tejidos se alinean con respecto a este y cuando dicha

energía cesa, el núcleo que ha captado esa energía la devuelve, y esta puede ser captada

desde el exterior mediante un receptor de campo magnético adecuado. Esta información

es empleada a continuación para construir una imagen con un alto nivel anatómico que

puede ser visualizada en dos o tres dimensiones y que tras ser analizada permitirá la

determinación del volumen, la densidad, la localización y la composición de la materia gris

y materia blanca cerebral.

2.1.1.2. Tomografía por emisión de positrones

La tomografía por emisión de positrones es una técnica no invasiva de diagnóstico

por imagen que permite determinar el nivel de actividad metabólica de los diferentes

tejidos del cuerpo humano, especialmente del sistema nervioso central, tras la

administración de una sustancia marcada radiactivamente. Como norma general se marca

alguna sustancia química como el oxígeno, el hidrógeno o la glucosa que será inyectada en

la sangre del paciente. Una vez fijada al tejido y transcurrido un tiempo (variable para cada

sustancia radioactiva), los átomos inestables del isótopo liberarán positrones que se

aniquilarán al contactar con los electrones de otros átomos circundantes. Dicho proceso de

aniquilación generará dos fotones que se desplazarán a la misma velocidad pero en

sentido opuesto y que serán detectados por un tomógrafo externo o sistema de detección

que mediante detectores de fotones situados en forma de cilindro alrededor de la cabeza,

será capaz de mapear el origen del proceso de aniquilación protón-electrón, y por tanto,

estimar la localización del proceso metabólico de interés (Mishina, 2008) (Figura 1).

P á g i n a | 1 1

Figura 1. Esquema del proceso de captura de la tomografía por emisión de positrones

De entre los diferentes radiofármacos emisores de positrones, el más utilizado es

el Flúor-18, un radiofármaco que al unirse a la 2-O-trifluorometilsulfonil manosa permite

la obtención del trazador 18-Flúor-Desoxi-Glucosa (18-FDG). La administración de este

radioisótopo permitirá identificar, localizar y cuantificar el consumo de glucosa de las

diferentes células del cerebro, permitiéndonos conocer que áreas se encuentran

hiperactivadas y cuales hipoactivadas. El uso de 18-FDG ha mostrado su validez en la

evaluación de los procesos adictivos, permitiéndonos conocer los efectos de la

administración aguda de una droga (Volkow et al., 2008), la respuesta a la presentación de

estímulos condicionados con esta (Volkow et al., 2011), el craving (Wang et al., 1999) o el

resultado de su consumo crónico (Galynker et al., 2007).

2.1.2. Hallazgos de neuroimagen en consumidores de distintas drogas

En esta sección presentamos los principales hallazgos derivados de la aplicación de

la resonancia magnética estructural y la tomografía por emisión de positrones, las técnicas

Unidad de Procesamiento

Aniquilación Reconstrucción de las imágenes

P á g i n a | 1 2

de neuroimagen cerebral utilizadas en nuestros estudios, a la investigación de los efectos

crónicos del consumo de cocaína, heroína, alcohol, MDMA y cannabis sobre la estructura o

morfología cerebral y sobre el metabolismo cerebral de la glucosa.

2.1.2.1. Cocaína

La cocaína produce sus efectos estimulantes a través de la inhibición de la

recaptación de los neurotransmisores dopamina, serotonina y norepinefrina. Se ha

propuesto que las alteraciones neuropsicológicas y de neuroimagen vinculadas al

consumo de cocaína están asociadas a neuroadaptaciones provocadas por la sobre-

estimulación de las vías dopaminérgicas y la consecuente hipo-activación o el agotamiento

de estas vías una vez abandonado el consumo (Gruber & Yergelun-Todd, 2001).

Alteraciones estructurales

La mayoría de los estudios que se han llevado a cabo con esta población han

encontrado reducciones del volumen de materia gris y materia blanca cerebral en las

cortezas prefrontal dorsolateral y orbitofrontal, frontal y temporal y en estructuras

subcorticales como la ínsula y en el cerebelo (Barrós-Loscertales et al., 2011b; Franklin et

al., 2002; Makris et al., 2004b; Matochik, London, Eldreth, Cadet, & Bolla, 2003; Sim et al.,

2007; Tanabe et al., 2009).

Alteraciones funcionales

La mayoría de los estudios de neuroimagen cerebral llevados a cabo en este campo

se han dirigido al estudio de la alteración de las vías dopaminérgicas (Martinez &

Narendran, 2010). Los pocos estudios que han evaluado el funcionamiento cerebral en

situación de reposo con PET han observado reducciones del metabolismo cerebral

preferentemente en áreas prefrontales (p.e. corteza prefrontal dorsolateral, orbitofrontal

y corteza anterior del cíngulo). En uno de los primeros estudios dirigidos a estudiar los

efectos a largo plazo del consumo de drogas en el metabolismo cerebral de consumidores

P á g i n a | 1 3

de cocaína en situación de abstinencia, Volkow y colaboradores (1992) encontraron

reducciones significativas de la actividad metabólica en la corteza frontal que persistieron

tras varios meses de abstinencia. Además, las alteraciones encontradas correlacionaron

con la dosis y la duración del consumo de cocaína. En otro estudio del mismo grupo, se

encontraron reducciones significativas del metabolismo regional del córtex prefrontal

(relacionadas con una menor concentración de receptores de dopamina en esta región) en

consumidores de cocaína (Volkow et al., 1993).

2.1.2.2. Heroína

La heroína y otros opiáceos ejercen sus efectos cerebrales a través de su acción

sobre los receptores específicos mu, delta y kappa, que se expresan en diversas áreas

relacionadas con los efectos reforzantes de las drogas. Estas sustancias también

incrementan la producción de dopamina, pero a través de una vía indirecta, al reducir la

actividad inhibitoria del GABA en el área tegmental ventral (Camí & Farré, 2003).

Muchos de los estudios encontrados en esta población presentan importantes

limitaciones metodológicas, entre las que cabe destacar el uso de pacientes en tratamiento

con metadona o de pacientes con importantes niveles de co-abuso de otras drogas (Lyoo

et al., 2004; Pezawas et al., 2002; Reid et al., 2008).

Alteraciones estructurales

Los estudios llevados a cabo en consumidores de heroína en situación de

abstinencia han mostrado discrepancias importantes. No obstante, en los últimos años, el

desarrollo de nuevas técnicas y análisis de neuroimagen cerebral ha permitido demostrar

la existencia de reducciones significativas de la materia gris cerebral en regiones

prefrontales y temporales en estas poblaciones (Pezawas et al., 1998; Yuan et al., 2009).

Pezawas y colaboradores (1998) encontraron ensanchamientos del espacio ventricular y

pericortical en consumidores de heroína en situación de abstinencia, revelando

P á g i n a | 1 4

reducciones del volumen cerebral en estos individuos. Asimismo, el volumen regional del

córtex frontal estaba correlacionado con la duración de la abstinencia en los consumidores

de heroína, de modo que la pérdida de volumen frontal era significativamente inferior en

aquellos individuos con periodos de abstinencia superiores a un año. En el estudio llevado

a cabo por Yuan y colaboradores (2009), se encontró una reducción significativa de la

densidad de materia gris cerebral en la corteza prefrontal, temporal y el cíngulo y análisis

posteriores controlando el efecto de la edad, el nivel educativo, el género, el uso de

nicotina y la duración de la abstinencia mostraron que la duración del consumo de heroína

correlacionaba negativamente con la densidad de la materia gris de estos pacientes.

Alteraciones funcionales

Los estudios del metabolismo cerebral llevados a cabo en estas poblaciones con

PET también son contradictorios. En este sentido, mientras que Galynker y colaboradores

(2000) encontraron aumentos significativos del metabolismo cerebral en la parte anterior

del cíngulo en consumidores de heroína en situación de abstinencia que habían recibido

tratamiento con metadona durante varios años, un estudio de este mismo grupo encontró

que los consumidores de opiáceos mostraban reducciones significativas del metabolismo

cerebral en las partes medial y anterior del cíngulo, ínsula y corteza frontal (Galynker et

al., 2007)

2.1.2.3. Alcohol

El alcohol (etanol) ejerce sus efectos psicoactivos en el sistema nervioso central a

través de su interacción con múltiples sistemas de neurotransmisores, incluyendo

receptores de GABA, serotonina, opioides y N-Metil-D-Aspartato (NMDA) (un subtipo de

receptor del glutamato). Adicionalmente, el alcohol incrementa los niveles de dopamina en

el estriado a través de un mecanismo indirecto de activación de los receptores GABA, o

inhibición de los receptores NMDA (Camí y Farre, 2003).

P á g i n a | 1 5

Los deterioros vinculados al consumo de alcohol han sido estudiados durante

décadas. Las investigaciones iniciales se centraron en el estudio de los deterioros

asociados con condiciones extremas de alcoholismo como el síndrome de Wernicke-

Korsakoff. Posteriormente, diversos estudios han contribuido a delimitar las alteraciones

neuropsicológicas y de imagen cerebral asociadas al consumo de alcohol en pacientes no

amnésicos.

Alteraciones estructurales

Múltiples estudios de neuroimagen estructural han documentado la existencia de

importantes alteraciones morfológicas en el cerebro de individuos consumidores de

alcohol. Se ha demostrado que estas alteraciones afectan de modo generalizado a diversos

aspectos de la sustancia gris y la sustancia blanca, produciendo atrofia cortical y

reducciones globales del volumen cerebral. Aunque las alteraciones son especialmente

pronunciadas en los lóbulos frontales, el sistema límbico y el cerebelo (Bühler & Mann,

2011; Chanraud et al., 2007; Makris et al., 2008a; Mechtcheriakow et al., 2007; Jang et al.,

2007), otras regiones como la corteza parietal y temporal también han sido encontradas

afectadas (Fein, Shimotsu, Chu, & Barakos, 2009; Gazdzinski et al., 2005).

Alteraciones funcionales

Con la excepción de un estudio (Eckardt, Rohrbaugh, Rio, & Martin, 1990), todos

los estudios llevados a cabo en pacientes dependientes de alcohol en situación de

abstinencia han demostrado reducciones significativas del metabolismo cerebral

fundamentalmente en estructuras frontales y prefrontales (Adams et al., 1993; Dao-

Castellana et al., 1998; Gilman et al., 1990). En un estudio en el que se combinaron técnicas

de PET y resonancia magnética con el objetivo de mejorar la precisión anatómica de las

regiones de interés, Dao-Castellana y colaboradores (1998) detectaron alteraciones del

metabolismo regional en el córtex frontal y prefrontal dorsolateral de consumidores de

alcohol.

P á g i n a | 1 6

2.1.2.4. Éxtasis (MDMA)

El éxtasis, cuyo componente principal es usualmente la 3-

4metilendioximetanfetamina (MDMA), ejerce sus efectos sobre el sistema nervioso central

a través de la inhibición de la recaptación de la serotonina.

Aunque se ha demostrado que la administración de MDMA produce efectos

neurotóxicos selectivos sobre la serotonina en animales (Baumann, Wang, & Rothman,

2007; Green, Mechan, Elliott, O’Shea, & Colado, 2003), los efectos de esta sustancia sobre el

sistema nervioso en humanos han sido ampliamente discutidos (p.e. Green, King, Shortall,

& Fone, 2011). En los últimos años, se ha demostrado la presencia de alteraciones tanto a

nivel estructural como a nivel funcional en consumidores de esta sustancia con o sin

consumo de otras sustancias. Alteraciones consistentes con reducciones significativas de

la disponibilidad de transportadores serotonérgicos en diversas regiones frontales,

temporales, mediales y basales (Parrott, 2012).

Alteraciones estructurales

El consumo crónico de éxtasis ha sido asociado con alteraciones morfológicas

incluso tras importantes períodos de abstinencia en regiones frontales, temporales,

occipitales, cíngulo anterior, cerebelo y ganglios basales. En un estudio llevado a cabo por

Cowan y colaboradores (2003) en el que se comparaban policonsumidores de drogas con

y sin co-abuso de MDMA, aquellos sujetos que consumían MDMA mostraron reducciones

significativas de la sustancia gris cerebral en áreas de la corteza frontal, temporal y

occipital, la corteza cingulada anterior, el tronco cerebral y el cerebelo. Asimismo, en un

reciente estudio en el que se comparó el volumen de materia gris cerebral entre

consumidores con alta y baja exposición de MDMA y anfetaminas y 16 controles sanos, se

encontró que, comparados con los consumidores con bajos niveles de exposición y los

controles sanos, aquellos sujetos que presentaban altos niveles de consumo de estas

P á g i n a | 1 7

sustancias mostraban reducciones significativas de materia gris cerebral en la corteza

orbitofrontal y la corteza anterior del cíngulo.

Alteraciones funcionales

Los estudios del metabolismo cerebral llevados a cabo con PET en consumidores

de éxtasis sugieren una reducción del metabolismo en el giro frontal inferior, los ganglios

basales, el hipocampo, la amígdala y el estriado (Buchert et al., 2001; Obrocki et al., 1999,

2002). Análisis posteriores llevados a cabo por estos autores encontraron asociaciones

significativas entre la duración de la abstinencia y el metabolismo del cíngulo y la amígdala

así como mayores reducciones en aquellos consumidores que habían empezado a utilizar

la sustancia antes de los 18 años de edad, indicando un efecto específico de vulnerabilidad

a la neurotoxicidad de la sustancia en edades tempranas. Sin embargo, no se encontraron

asociaciones significativas entre la cantidad de droga consumida y el metabolismo

cerebral (Buchert et al., 2001; Obrocki et al., 2002).

2.1.2.5. Cannabis

El cannabis, compuesto en su mayor parte por delta-9-tetrahidrocannabinol (THC)

produce sus efectos psicoactivos en el cerebro a través de su acción sobre receptores

cannabinoides CB1. El cannabis, al igual que la heroína, también estimula la producción de

dopamina de manera indirecta a través de la acción de los receptores CB1 sobre las

neuronas de los neurotransmisores de glutamato y GABA en el área tegmental ventral y el

estriatum (Camí & Farré, 2003).

Los estudios llevados a cabo con animales demuestran que la administración de

THC tiene efectos neurotóxicos sobre regiones cerebrales ricas en receptores

canabinoides como el hipocampo, la amígdala o el cerebelo (Pertwee & Ross, 2002). Sin

embargo, en contraste con la literatura animal, los estudios llevados a cabo en humanos

han reportado resultados contradictorios (Martín-Santos et al., 2010).

P á g i n a | 1 8

Alteraciones estructurales

En general no se han encontrado diferencias significativas en el volumen total de

materia gris cerebral entre consumidores de esta sustancia y controles sanos. Sin

embargo, estudios de ciertas regiones cerebrales han encontrado reducciones

significativas de materia gris en el hipocampo, la amígdala y regiones adyacentes en

consumidores abstinentes de esta sustancia (Lorenzetti, Lubman, Whittle, Solowij, &

Yücel, 2010). En general, parece que estas alteraciones se producen en consumidores

severos de esta sustancia (Matochik, Eldreth, Cadet, & Bolla, 2005; Yücel et al., 2008) o en

aquellos que inician su consumo a edades más tempranas (Wilson et al., 2000).

Alteraciones funcionales

Las alteraciones asociadas a los efectos residuales del cannabis consisten en

reducciones significativas del metabolismo cerebral localizadas preferentemente en

regiones frontales y el cerebelo (Sevy et al., 2008; Volkow et al., 1996). En un reciente

estudio llevado a cabo por Sevy y colaboradores (2008) en el que se comparó a un grupo

de 6 consumidores de cannabis en abstinencia con un grupo de 6 controles sanos, los

consumidores de cannabis presentaron reducción del metabolismo cerebral en la corteza

orbitofrontal, el putamen y el precuneus. En este estudio, sólo el metabolismo del córtex

parietal mostró una correlación negativa con la cantidad de uso de esta sustancia.

2.2. Impulsividad y adicciones

La impulsividad es un rasgo estable de la personalidad que varía normativamente

a través de la población normal (Barratt, 1959; Patton, Stanford, & Barratt, 1995) y que ha

sido definida desde la psicología como aquella conducta llevada a cabo con poca o

inadecuada planificación (Evenden, 1999). Este rasgo multifacético de la personalidad, que

en ocasiones puede ser adaptativo (Dickman, 1990), es considerado un rasgo disfuncional

P á g i n a | 1 9

asociado con actos violentos, peligrosos o inadecuados para las demandas de la situación

(Verdejo-García et al., 2008).

Desde de la neuropsicología la impulsividad es equiparada con la "desinhibición",

esto es, la alteración de los mecanismos de control “de arriba a abajo” encargados de

suprimir respuestas automatizadas o impulsadas por la obtención de refuerzos

inmediatos no apropiadas para la situación (Aron, 2007). Estos “mecanismos de control”

pueden verse afectados tanto por lesiones cerebrales como por trastornos mentales dando

lugar a conductas impulsivas. Definida de esta manera, la impulsividad juega un papel

fundamental en los trastornos adictivos. Y es que, las características de personalidad de

una persona podrían influir en las primeras etapas del consumo de la droga por ejemplo

determinando si la persona probará o no la droga o que cantidad tomará y a medida que

avanza el consumo, determinar que la persona persista en su conducta a pesar de las

consecuencias negativas que esta tiene para sí mismo y para los demás (y a pesar de sus

intentos por no hacerlo) o incluso que recaiga tras importantes periodos de abstinencia.

Dentro de la neuropsicología, la impulsividad o desinhibición ha sido evaluada a

través de dos tipos de instrumentos: los cuestionarios de impulsividad y las pruebas

neuropsicológicas que tratan de evaluar la conducta impulsiva o el control inhibitorio.

A continuación se presenta una breve revisión de las medidas que se han utilizado

con más frecuencia en el campo de las adicciones para evaluar la conducta impulsiva y los

principales hallazgos encontrados resultado del uso de las mismas en consumidores en

situación de abstinencia de cocaína, heroína, alcohol, MDMA y cannabis.

2.2.1 Instrumentos de evaluación de la impulsividad

2.2.1.1. Inventarios de personalidad impulsiva

Existe una amplia gama de cuestionarios destinados a medir la impulsividad y

ciertas conductas impulsivas como la búsqueda de sensaciones o la toma de riesgos. En

P á g i n a | 2 0

español contamos con la Escala de Impulsividad de Barratt (Barrat Impulsivity Scale -BIS)

(Oquendo et al., 2001; Patton et al., 1995), la UPPS (Verdejo-García, Lozano, Moya, Alcázar,

& Pérez-García, 2010; Whiteside & Lynam, 2001, 2003), la Escala de Búsqueda de

Sensaciones (Sensation Seeking Scale -SSS) del Cuestionario de Personalidad Zuckerman-

Kuhlman (Zuckerman et al., 1993), adaptada al castellano por Pérez y Torrubia (1986) o la

Escala de Sensibilidad al Castigo y Sensibilidad a la Recompensa (SCSR) (Torrubia, Ávila,

Moltó, & Caseras, 2001). El uso de estos cuestionarios ha mostrado que los consumidores

de diversas sustancias presentan mayores puntuaciones que los controles no

consumidores en diversos índices de impulsividad (ver revisión en Verdejo-García et al.,

2008).

2.2.1.2. Medidas de control inhibitorio

Este componente ha sido definido como la capacidad para inhibir o demorar

respuestas automatizadas, impulsivas o guiadas por el reforzamiento inmediato (Verdejo-

García et al., 2008). Para la evaluación de la impulsividad suelen emplearse tres tipos de

pruebas, aquellas que evalúan la inhibición de respuesta o la capacidad del individuo para

suprimir una respuesta automatizada, medidas de delay-discounting o “descuento

asociado a la demora” y aquellas que tratan de evaluar la impulsividad cognitiva en

términos de toma de decisiones desadaptativas. Dentro de las que permiten evaluar la

inhibición de respuesta o inhibición atencional destacan el test Stroop (Golden, 1978; TEA,

1993), el test de los Cinco Dígitos (Sedó, 2005), las tareas Go – No Go y el Stop Signal Task

(Logan, Cowan, & Davis, 1984). El test de Stroop evalúa la capacidad del individuo para

controlar una respuesta automatizada (leer una palabra) y dar la respuesta adecuada

(decir el color en el que está escrita la palabra) y ha mostrado consistentemente su

capacidad para detectar problemas de atención selectiva e inhibición en consumidores de

distintos tipos de drogas (Verdejo-García, Benbrook, Funderburk, David, Cadet, & Bolla,

2007a; Woicik et al., 2009). El test de los Cinco Dígitos ha demostrado ser una herramienta

P á g i n a | 2 1

útil en la discriminación de los perfiles neuropsicológicos de consumidores de cocaína vs.

opiáceos (Verdejo-García, Perales, & Pérez-García, 2007b) y por último, las tareas Go - No

Go y Stop-Signal han demostrado ser sensibles a la detección de alteraciones en este

componente en consumidores de distintos tipos de drogas (Fillmore & Rush, 2002;

Lawrence, Luty, Bogdan, Sahakian, & Clark, 2009; Monterosso, Aron, Cordova, Xu, &

London, 2005). Para evaluar el “descuento asociado a la demora” o la tendencia de un

individuo a reducir el valor de una recompensa en función del tiempo de espera (Bickel &

Marsch, 2001; Reynolds, 2006), se han utilizado cuestionarios en los que el individuo

selecciona sus preferencias a nivel hipotético (Kirby & Petry, 2004) o tareas en las que el

sujeto tiene que elegir entre pequeñas cantidades entregadas de inmediato o cantidades

mayores entregadas a medio y largo plazo (Reynolds & Schiffbauer, 2004). Por último,

para evaluar el continuo reflexividad-impulsividad se ha recurrido al test de

emparejamiento de figuras conocidas (MFFT-20) (Cairns & Cammock, 2002), la tarea de

recolección de información (Information Sampling Task) (Clark, Robbins, Ersche, &

Sahakian, 2006) y a pruebas de toma de decisiones como la Iowa Gambling Task (Bechara,

Damasio, Damasio, & Anderson, 1994), la Cambridge Gamble Task o la Risky Gains Task

(Rogers et al., 1999a; Rogers et al., 1999b).

2.2.2. Hallazgos relacionados con la impulsividad en consumidores de distintas drogas

2.2.2.1. Cocaína

Inventarios de personalidad impulsiva

El uso de cuestionarios de impulsividad en consumidores de cocaína en situación

de abstinencia ha mostrado que estos pacientes presentan altas puntuaciones de

impulsividad. Las investigaciones llevadas a cabo con estos cuestionarios ha demostrado

que al compararlos con individuos no consumidores, los individuos consumidores de

cocaína presentan altas puntuaciones de urgencia positiva y urgencia negativa (o

tendencia a actuar de forma impulsiva bajo condiciones de estado de ánimo positivo y

P á g i n a | 2 2

negativo respectivamente) (Albein-Urios, Martinez-González, Lozano, Clark, & Verdejo-

García, 2012), falta de premeditación (Fernández-Serrano, Perales, Moreno-López, Pérez-

García, & Verdejo-García, 2012) y búsqueda de sensaciones (Ersche, Turton, Pradhan,

Bullmore, & Robbins, 2010a). Asimismo, el uso de estos cuestionarios en consumidores de

cocaína se ha mostrado como una herramienta de enorme validez en la predicción de la

severidad del consumo, las alteraciones asociadas con el mismo o la predicción de recaídas

(Moeller et al., 2001; Verdejo-García, Bechara, Recknor, & Pérez-García, 2007c)

Medidas de control inhibitorio

Los estudios llevados a cabo en consumidores de cocaína han encontrado

alteraciones en las tareas de inhibición de conducta, descuento asociado a la demora e

impulsividad cognitiva. En este sentido, se ha encontrado que los consumidores de cocaína

en situación de abstinencia presentan más errores durante tareas Go – No Go (Fernández-

Serrano et al., 2012; Verdejo-García et al., 2007b), un mayor tiempo de reacción durante

los ensayos de la tarea Stop Signal (Li, Milivojevic, Kemp, Hong, & Sinha, 2006), mayores

puntuaciones de interferencia durante la realización de pruebas de Stroop (Albein-Urios et

al., 2012; Bolla, Funderburk, & Cadet, 2000) y alteraciones en tareas de delay discounting

(Coffey, Gudleski, Saladin, & Brady, 2003; Heil, Johnson, Higgins, & Bickel, 2006; Kirby &

Petry, 2004; Moeller et al., 2002) y toma de decisiones como el Risky Game o la Iowa

Gambling Task (Stout, Busemeyer, Lin, Grant, & Bonson, 2004; Verdejo-García et al.,

2007b).

De entre las diferentes versiones de la tarea Stroop utilizadas con esta población,

cobran especial relevancia aquellas que han utilizado estímulos emocionales asociados

con la droga (Goldstein et al., 2007). Utilizando estas tareas, diversos estudios han

demostrado que los consumidores de drogas suelen presentar dificultades para filtrar la

información irrelevante o inadecuada cuando ésta tiene que ver con estímulos asociados

con la droga (Nestor, Ghahremani, Monterosso, & London, 2011). Asimismo, se ha descrito

P á g i n a | 2 3

la hipoactivación de los circuitos asociados con el control atencional durante la realización

de esta tarea en estudios de resonancia magnética funcional (Barrós-Loscertales et al.,

2011a; Goldstein et al., 2007).

2.2.2.2. Heroína

Inventarios de personalidad impulsiva

El uso de diversos cuestionarios de impulsividad en consumidores de heroína y

otros opiáceos ha demostrado que estos pacientes presentan elevados niveles de

impulsividad y búsqueda de sensaciones (Cohen et al., 2005; Madden, Petry, Badger, &

Bickel, 1997).

Medidas de control inhibitorio

Se ha encontrado que los consumidores de heroína en situación de abstinencia

presentan altas puntuaciones de interferencia en pruebas de interferencia atencional

(Fernández-Serrano, Pérez-García, Perales, & Verdejo-García, 2010a; Verdejo, Toribio,

Orozco, Puente, & Pérez-García, 2005a), importantes descuentos asociados a la demora

tanto en tareas en las que la recompensa es hipotética como en aquellas en las que es real

(Kirby, Petry, & Bickel, 1999; Kirby & Petry, 2004) y mayor número de errores en tareas

de impulsividad cognitiva (Lee and Pau, 2002) y en tareas de toma de decisiones como la

Iowa Gambling Task (Mintzer, Copersino, & Stitzer, 2005; Verdejo-García et al., 2007b) o la

Cambridge Gamble Task (Fishbein et al., 2007).

2.2.2.3. Alcohol

Inventarios de personalidad impulsiva

Se ha encontrado que los alcohólicos o pacientes con consumo principal de alcohol

en situación de abstinencia muestran índices altos de impulsividad (Mitchell, Fields,

D’Esposito, & Boettiger, 2005), búsqueda de sensaciones (Bjork, Hommer, Grant, &

P á g i n a | 2 4

Danube, 2004) y urgencia negativa (Whiteside & Lynam, 2003). Asimismo, el uso de

cuestionarios de personalidad ha mostrado su capacidad predictiva tanto en la frecuencia

con la que se consume alcohol como en la cantidad consumida o las consecuencias

negativas de dicho consumo (Cyders, Flory, Rainer, & Smith, 2009).

Medidas de control inhibitorio

Las investigaciones llevadas a cabo en esta población han encontrado que los

consumidores de alcohol en situación de abstinencia al ser comparados con individuos no

consumidores cometen más errores en paradigmas Go – No Go (Bjork et al., 2004), tienen

un mayor tiempo de reacción durante tareas de Stop Signal (Goudriaan, Oosterlaan, de

Beurs, & van den Brink, 2006; Lawrence et al., 2009) y presentan mayores índices de

“descuento asociado a la demora” (Mitchell et al., 2005; Petry, 2001). Estos individuos

también presentan alteraciones en tareas de reflexión-impulsividad (Weijers, Wiesbeck, &

Böning, 2001) y toma de decisiones (Dom, De Wilde, Hulstijn, van den Brink, & Sabbe,

2006; Fein, Klein, & Finn, 2004; Fernández-Serrano, Pérez-García, Schmidt Río-Valle, &

Verdejo-García, 2010b).

2.2.2.4. Éxtasis (MDMA)

Inventarios de personalidad impulsiva

El uso de cuestionarios de impulsividad ha permitido encontrar altas puntuaciones

de impulsividad y búsqueda de sensaciones incluso tras 12 meses de abstinencia (Gerra et

al., 2000; Hanson, Luciana, & Sullwold, 2008; Parrott, Sisk, & Turner, 2000).

Medidas de control inhibitorio

Los estudios llevados a cabo en consumidores de éxtasis en situación de

abstinencia han mostrado también la presencia de alteraciones neuropsicológicas durante

la realización de diferentes pruebas de impulsividad (Morgan, Impallomeni, Pirona, &

P á g i n a | 2 5

Rogers, 2006; Quednow et al., 2007). En un estudio correlacional que exploró la asociación

entre la cantidad de MDMA consumida y el rendimiento de un grupo de consumidores

“puros” de MDMA en una amplia batería de pruebas neuropsicológicas, Halpern y

colaboradores (2004) demostraron que aquellos individuos con mayor consumo de esta

sustancia presentaban también mayores puntuaciones de interferencia.

2.2.2.5. Cannabis

Inventarios de personalidad impulsiva

La mayoría de los estudios que han evaluado la personalidad impulsiva en estas

poblaciones han tratado de estudiar los efectos predictivos de estas características sobre

el consumo de la sustancia. Un reciente estudio llevado a cabo por Prince van Leeuwen y

colaboradores (2011) demostró que el uso de la escala de impulsividad de Barratt

permitía predecir el uso y la frecuencia del consumo de esta sustancia. Específicamente, se

encontró que una puntuación alta en la escala de activación conductual se asociaba con

una mayor probabilidad de iniciar el consumo y una puntuación baja en la escala de

inhibición conductual con una mayor probabilidad de persistir en dicho consumo.

Medidas de control inhibitorio

La presencia de alteraciones neuropsicológicas a largo plazo como consecuencia

del consumo de cannabis es todavía hoy objeto de debate. Sin embargo, recientemente se

han descrito alteraciones en consumidores abstinentes de esta sustancia en tareas de

toma de decisiones e impulsividad (Bolla, Eldreth, Matochik, & Cadet, 2005; Pope et al.,

2003; Verdejo-Garcia et al., 2007a).

2.3. Funciones ejecutivas y adicciones

Desde una perspectiva neuropsicológica se considera que la adicción es el

resultado de un conjunto de alteraciones cerebrales que afectan a múltiples sistemas

P á g i n a | 2 6

neurobiológicos y que resultan en disfunciones en procesos motivacionales, emocionales,

cognitivos y conductuales. En los últimos años, la evaluación neuropsicológica de los

individuos drogodependientes ha centrado su atención en las llamadas funciones

ejecutivas, un conjunto integrado de habilidades implicadas en la producción, supervisión

y control de conductas dirigidas a objetivos (Stuss & Knight, 2002; Roberts, Robbins, &

Weiskrantz, 1998) y en la regulación de los estados emocionales que se consideran

adaptativos para la consecución de esos objetivos (Bechara, Damasio, & Damasio, 2000;

Davidson, 2002; Stuss & Alexander, 2000). En el contexto de las adicciones, el principal

objetivo de la neuropsicología es contribuir a dilucidar la naturaleza y extensión de las

alteraciones asociadas con el consumo de drogas mediante el uso de pruebas y tareas

conductuales que impliquen el funcionamiento selectivo de los distintos sistemas

cerebrales implicados en la adicción.

A continuación, se presentarán las principales pruebas utilizadas en la evaluación

neuropsicológica de individuos drogodependientes junto con los principales hallazgos

derivados de su utilización en consumidores de drogas en situación de abstinencia. Como

en apartados anteriores, nos centraremos en el resultado del consumo crónico de cocaína,

heroína, alcohol, MDMA y cannabis, las drogas más consumidas por los participantes de

nuestros estudios.

2.3.1. Evaluación neuropsicológica de las funciones ejecutivas

La aplicación de análisis factoriales y ecuaciones estructurales ha permitido

determinar la existencia de al menos tres componentes diferenciados aunque no

totalmente independientes dentro de las funciones ejecutivas (Fisk & Sharp, 2004; Miyake

et al., 2000). Estos componentes han sido definidos como actualización, inhibición de

respuestas y cambio. Además, recientemente se ha propuesto que la toma de decisiones

constituiría un cuarto componente independiente dentro de las funciones ejecutivas. Una

propuesta avalada por la evidencia que demuestra la existencia de importantes

P á g i n a | 2 7

alteraciones en los procesos de toma de decisiones en individuos drogodependientes y el

hecho de que estas no correlacionen con la ejecución de estos individuos en tareas que

implicadas componentes de actualización, inhibición y cambio (Bechara et al., 2001; Grant,

Contoreggi, & London, 2000; Verdejo-García & Pérez-García, 2007). Asimismo, en los

últimos años, los modelos neuropsicológicos de la adicción han enfatizado la importancia

de los factores emocionales en los procesos adictivos (Goldstein & Volkow, 2002; Redish,

Jensen, & Johnson, 2008; Verdejo-García & Bechara, 2009).

2.3.1.1. Actualización

El componente de actualización implica la monitorización, actualización y

manipulación de la información “on line” en la memoria operativa (Miyake et al., 2000).

Recientemente se ha propuesto que este componente incluye los procesos de memoria de

trabajo, fluidez y razonamiento (Verdejo-García & Pérez-García, 2007).

La memoria de trabajo es un sistema que permite el almacenamiento,

manipulación y actualización temporal de la información en el cerebro (M D’Esposito et al.,

1995). Tradicionalmente se ha evaluado utilizando algunos subtests de la escala WAIS

(Wechsler Adult Intelligence Scale, Wechsler, 1997) y otros instrumentos como la prueba

de Span visual de la Escala de Memoria Wechsler o las tareas n-back (Watter, Geffen, &

Geffen, 2001). Estas tareas han demostrado ser eficaces en la detección de alteraciones en

procesos de memoria de trabajo en consumidores de diversas drogas tanto en estudios

conductuales como en estudios de neuroimagen funcional (Bustamante, Barrós-

Loscertales, Ventura-Campos, Sanjuán, Llopis, Parcet, & Avila, 2011a; Tomasi, Goldstein,

Telang, Maloney, Alia-Klein, Caparelli, & Volkow, 2007). Podemos definir la fluidez como la

capacidad del individuo para iniciar su conducta de forma espontánea y creativa en

respuesta a una orden novedosa. Para evaluar este proceso se han empleado el Test de

fluidez verbal (FAS) (Lezak, 2004), Animales, Frutas y Herramientas y el Ruff Figural

Fluency Test (RFFT) (Ruff, 1996). Por último, el razonamiento analógico consiste en

P á g i n a | 2 8

obtener una conclusión a partir de premisas sobre las que se establece una comparación o

analogía entre elementos o conjuntos de elementos distintos. Para la evaluación de este

dominio tradicionalmente se ha recurrido al subtest de Semejanzas del WAIS.

2.3.1.2. Control inhibitorio

La evaluación del control inhibitorio ha sido ampliamente descrita en el apartado

de evaluación de la impulsividad en individuos drogodependientes. Brevemente, este

componente ha sido evaluado a través de tareas de inhibición de respuesta, “descuento

asociado a la demora” y tareas de toma de decisiones.

2.3.1.3. Flexibilidad cognitiva

La flexibilidad cognitiva es la capacidad del individuo para reestructurar su

conocimiento y adaptar su respuesta a las exigencias cambiantes del ambiente (Spiro &

Jehng, 1990). Se trata también de un componente multidimensional que ha sido estudiado

a través de distintos índices incluyendo pruebas que miden la respuesta del individuo ante

el cambio de las reglas de una tarea, el criterio de respuesta o el set atencional y tareas de

reversal learning que miden la capacidad del individuo para cambiar su respuesta en

función de cambios en los patrones de reforzamiento.

Para la evaluación de la respuesta del individuo ante el cambio de set se ha

recurrido a tareas como la Prueba de Categorías (DeFilippis, 2002) o el Test de

Clasificación de Tarjetas de Wisconsin (WCST) (Grant & Berg, 1948). La prueba de

Categorías es una tarea informatizada que presenta tamaños del efecto considerables en la

discriminación del rendimiento de consumidores y controles (Verdejo-García & Pérez-

García, 2007) y ha demostrado buenos valores de predicción del resultado del tratamiento

en consumidores de cocaína (Turner et al., 2009).

2.3.1.4. Toma de decisiones

P á g i n a | 2 9

La toma de decisiones es la habilidad para seleccionar de entre un conjunto de

posibles alternativas aquella que resulte más beneficiosa para el individuo. Una de las

pruebas más utilizadas dentro del campo de la neuropsicología es la Iowa Gambling Task

(Bechara et al., 1994). Esta prueba ha mostrado ser sensible a la detección de alteraciones

en la toma de decisiones en consumidores de diversas drogas. Otras tareas empleadas en

la evaluación de la toma de decisiones son el Juego de Datos (Game of Dice Task) o la

Cambridge Gamble Task. Estas tareas pueden ser descritas como pruebas más ecológicas

ya que el individuo tiene información sobre las potenciales consecuencias de sus acciones

(Brand, Labudda, & Markowitsch, 2006). En los últimos años, diversos estudios han

mostrado que la utilización de índices de toma de decisiones en condiciones ambiguas y de

riesgo son de gran utilidad en la predicción de recaídas (Bowden-Jones, McPhillips,

Rogers, Hutton, & Joyce, 2005; Passetti, Clark, Mehta, Joyce, & King, 2008; Paulus, Tapert,

& Schuckit, 2005).

2.3.1.5. Procesos emocionales

Las investigaciones llevadas a cabo en pacientes drogodependientes han

demostrado que estos pacientes presentan alteraciones tanto en la percepción como en la

experiencia emocional.

Uno de los aspectos claves del funcionamiento emocional es la capacidad para

identificar y reconocer señales emocionales en las caras de otras personas. Se ha

demostrado que el reconocimiento emocional es fundamental para la conducta prosocial,

la socialización y la interacción normal (Blair, 2003) y que la alteración de dicho proceso

cursa con alteraciones en la experiencia emocional (Calder & Young, 2005). El paradigma

más utilizado en la evaluación de la percepción emocional consiste en la presentación de

un conjunto de fotografías de personas que expresan diferentes emociones que la persona

tiene que identificar. De entre las diferentes tareas, la prueba más utilizada dentro del

campo de las adicciones ha sido el test de Ekman. Esta prueba evalúa mediante la

P á g i n a | 3 0

presentación de un conjunto de estímulos procedentes del banco de “Expresiones Faciales

Emocionales: Estímulos y Tests” (Facial Expressions of Emotion: Stimuli and Tests –

FEEST) (Young, Perrett, Calder, Sprengelmeyer & Ekman, 2002) la capacidad del individuo

para identificar expresiones faciales representativas de las seis emociones básicas

(felicidad, tristeza, miedo, asco, ira y sorpresa).

2.3.2. Hallazgos neuropsicológicos relacionados con las funciones ejecutivas en

consumidores de distintas drogas

Una revisión de los estudios llevados a cabo hasta el momento en consumidores de

drogas de diferente tipo, indica que aunque existen ciertos mecanismos neuropsicológicos

que parecen verse afectados por todas las drogas de abuso, también existen ciertas

alteraciones que podrían ser especificas al consumo de cada sustancia (Fernández-

Serrano, Pérez-García, & Verdejo-García, 2011). En este sentido, aunque todas las drogas

han sido asociadas con la presencia de alteraciones en los procesos de memoria,

actualización, toma de decisiones y procesamiento emocional, el uso de psicoestimulantes

y alcohol ha sido asociado con el control inhibitorio y la flexibilidad, el consumo de alcohol

y MDMA con el procesamiento espacial, la velocidad perceptiva y la atención selectiva, el

uso de cannabis y metanfetaminas con la presencia de déficits en memoria prospectiva y

por último, el uso de cannabis y MDMA ha sido asociado con la presencia de alteraciones

en velocidad de procesamiento y planificación.

Por otra parte, la mayoría de las investigaciones que han tratado de evaluar la

percepción emocional en sujetos consumidores de drogas se han centrado en los efectos

producidos por el consumo de alcohol, y en la mayor parte de los casos en el estudio de la

habilidad para estimar la intensidad de las emociones expuestas pero no la exactitud o

precisión en el reconocimiento emocional. De entre aquellos estudios dirigidos a estudiar

la eficacia del reconocimiento emocional en estas poblaciones, aunque algunos estudios no

han encontrado diferencias significativas entre consumidores de drogas y controles

P á g i n a | 3 1

(Salloum et al., 2007; Woicik et al., 2009), la mayoría de los estudios llevados han

encontrado diferencias significativas en el reconocimiento de las emociones en general y

de las emociones negativas en particular. Asimismo, los estudios que han tratado de

evaluar la experiencia emocional en pacientes drogodependientes coinciden en señalar

que estos pacientes se activan menos y muestran mayor control que los individuos no

consumidores (Aguilar de Arcos et al., 2008; Aguilar de Arcos, Verdejo-García, Peralta-

Ramírez, Sánchez-Barrera, & Pérez-García, 2005), resultados que coinciden con la

creciente evidencia que demuestra que los individuos drogodependientes presentan una

sobreactivación de las regiones cerebrales implicadas en el procesamiento emocional de

recompensas en respuesta a estímulos asociados con la sustancia de consumo (Childress

et al., 1999; George et al., 2001; Kilts, Gross, Ely, & Drexler, 2004; Tapert, Brown, Baratta, &

Brown, 2004) y una reducción de la actividad cerebral en estas mismas regiones en

respuesta a reforzadores naturales como imágenes de contenido sexual (Asensio et al.,

2010; Garavan et al., 2000).

2.3.2.1. Cocaína

El consumo de cocaína ha sido consistentemente asociado con la presencia de

alteraciones neuropsicológicas en los procesos cognitivos de atención y memoria,

funcionamiento ejecutivo y procesos emocionales. Una revisión de los estudios llevados a

cabo en los últimos años indica que las alteraciones más frecuentemente encontradas en

consumidores de cocaína y policonsumidores con consumo principal de cocaína son las

alteraciones de los procesos de memoria, flexibilidad e inhibición (Fernández-Serrano et

al., 2011; Ruth Janke van Holst & Schilt, 2011). La alteración de la memoria de trabajo

constituye una de las características más frecuentemente reportadas en estudios de

cocaína (Fernández-Serrano et al., 2011) y ha sido recientemente asociada con la

neurotoxicidad de esta sustancia (Albein-Urios et al., 2012). Asimismo se han descrito

alteraciones en atención (Pace-Schott et al., 2008; Verdejo-García & Pérez-García, 2007;

P á g i n a | 3 2

Woicik et al., 2009), memoria y aprendizaje verbal (De Oliveira et al., 2009; Fox, Jackson, &

Sinha, 2009; Goldstein et al., 2004; Jovanovski, Erb, & Zakzanis, 2005; Verdejo-García &

Pérez-García, 2007; Woicik et al., 2009), fluidez (Verdejo-García & Pérez-García, 2007),

inhibición (Verdejo-García & Pérez-García, 2007), flexibilidad cognitiva (Goldstein et al.,

2004; Woicik et al., 2009), toma de decisiones (Fernández-Serrano et al., 2010b; Verdejo-

Garcia et al., 2007a) y reconocimiento emocional (Fernández-Serrano, Lozano, Pérez-

García, & Verdejo-García, 2010c; Verdejo-García, Rivas-Pérez, Vilar-López, & Pérez-García,

2007d).

2.3.2.2. Heroína

La mayoría de los estudios llevados a cabo en consumidores de opiáceos o heroína

defienden la existencia de alteraciones en los dominios de fluencia, inhibición y toma de

decisiones (van Holst & Schilt, 2011). En una revisión llevada a cabo en 2007, Gruber y

colaboradores concluyeron que el consumo de opiáceos producía alteraciones a largo

plazo en los procesos de atención, concentración, memoria y habilidades visuo-

perceptivas. Otros autores han descrito alteraciones en los dominios de memoria de

trabajo, flexibilidad cognitiva e inhibición (Brand, Rothbauer, Driessen, & Markowitsch,

2008; Prosser et al., 2008). Asimismo se han descrito alteraciones en fluidez (Fernández-

Serrano et al., 2010a), razonamiento (Verdejo et al., 2005a), toma de decisiones (Brand et

al., 2008; Verdejo-García & Pérez-García, 2007) y procesos emocionales (Kornreich et al.,

2003; Wang et al., 2010).

2.3.2.3. Alcohol

El consumo de alcohol ha sido consistentemente asociado con la presencia de

alteraciones neuropsicológicas en los procesos de atención y memoria, funcionamiento

ejecutivo y procesos emocionales. Una reciente revisión de los estudios que han tratado de

evaluar las alteraciones neuropsicológicas producidas por el consumo de alcohol en

individuos drogodependientes en situación de abstinencia concluye que las alteraciones

P á g i n a | 3 3

más frecuentemente reportadas son las encontradas en memoria de trabajo, memoria

verbal, inhibición y habilidades visuo-perceptivas (van Holst & Schilt, 2011). El consumo

de alcohol y psicoestimulantes han sido asociado con la presencia de alteraciones del

control inhibitorio y la flexibilidad cognitiva y el consumo de esta sustancia con MDMA ha

sido asociado con alteraciones en tareas de procesamiento espacial, velocidad perceptiva y

atención selectiva (Fernández-Serrano et al., 2011). En general, aunque parece que se

produce una mejora de los procesos cognitivos tras periodos de abstinencia relativamente

breves (Manning et al., 2008; Moriyama, Muramatsu, Kato, Mimura, & Kashima, 2006;

Tedstone & Coyle, 2004; Zinn, Stein, & Swartzwelder, 2004), los procesos de

funcionamiento ejecutivo se han encontrado alterados incluso tras varios años de

abstinencia (Davies et al., 2005; Fein et al., 2004).

A nivel emocional, se ha encontrado que los alcohólicos tienden a sobreestimar la

intensidad de la emoción que aparece en las expresiones faciales de felicidad, ira y asco

(Foisy et al., 2007; Kornreich et al., 2001; Townshend & Duka, 2003) y que presentan

alteraciones en el reconocimiento de las emociones en general y de las emociones

negativas en particular (Frigerio, Burt, Montagne, Murray, & Perrett, 2002; Townshend &

Duka, 2003).

2.3.2.4. Éxtasis (MDMA)

El consumo de éxtasis o MDMA ha sido asociado con alteraciones en todos los

procesos neuropsicológicos descritos. Parece que las alteraciones más frecuentemente

encontradas son aquellas que se producen en tareas de memoria, concretamente en

aquellas que evalúan memoria verbal (Kalechstein, De La Garza, Mahoney, Fantegrossi, &

Newton, 2007; Parrott, 2012; Ruth Janke van Holst & Schilt, 2011). También se han

descrito alteraciones en el reconocimiento emocional (Yip & Lee, 2006).

En los últimos años, varios estudios han llamado la atención sobre el efecto que el

co-abuso de otras sustancias y en especial del cannabis podría tener sobre los efectos

P á g i n a | 3 4

neuropsicológicos asociados con el consumo de éxtasis (p.e. Schulz, 2011). En este sentido,

mientras que el uso de MDMA y alcohol ha sido asociado con la presencia de alteraciones

en los procesos de atención selectiva, procesamiento espacial y velocidad perceptiva en

consumidores de estas sustancias, el uso de MDMA y cannabis ha sido asociado con la

planificación y la velocidad de procesamiento (Fernández-Serrano et al., 2011).

2.3.2.5. Cannabis

Todavía hoy existe un importante debate acerca de las alteraciones producidas por

el consumo de cannabis. Aunque el consumo de esta sustancia ha sido consistentemente

asociado con la presencia de alteraciones neuropsicológicas a corto plazo sobre todo en

los procesos de memoria y atención, las revisiones publicadas hasta el momento de los

estudios que han tratado de evaluar los efectos a largo plazo de esta sustancia coinciden

en señalar la ausencia de alteraciones significativas en consumidores de esta sustancia en

situación de abstinencia (Grant, Gonzalez, Carey, Natarajan, & Wolfson, 2003; van Holst &

Schilt, 2011). En los últimos años, algunos autores han reportado alteraciones tras

importantes periodos de abstinencia en consumidores severos de la sustancia (Bolla et al.,

2005; Montgomery, Seddon, Fisk, Murphy, & Jansari, 2012; Verdejo-García & Pérez-García,

2007) y en aquellos que iniciaron su consumo en la adolescencia (Fontes et al., 2011;

Medina et al., 2007; Pope et al., 2003). Parece que estas alteraciones se producen

fundamentalmente en los procesos de toma de decisiones, la formación de conceptos y la

planificación (Crean, Crane, & Mason, 2011).

P á g i n a | 3 5

JUSTIFICACIÓN Y OBJETIVOS [ ]

P á g i n a | 3 6

La aplicación de las técnicas de neuroimagen cerebral al ámbito de las

drogodependencias ha permitido en los últimos años examinar “in vivo” los efectos agudos

de la administración de sustancias psicoactivas, los correlatos cerebrales del deseo intenso

por consumir las drogas (craving), o las alteraciones a largo plazo en regiones cerebrales y

sistemas neuroquímicos implicados en el consumo crónico, la dependencia y la recaída.

En los últimos años, un aspecto metodológico que ha suscitado un creciente interés

es la dirección causal de la relación entre las alteraciones neuropsicológicas encontradas y

el consumo de drogas, ya que diversos estudios han puesto de manifiesto que las

alteraciones neuropsicológicas (en especial las alteraciones de las funciones ejecutivas)

pueden predecir el inicio del consumo de drogas y contribuir al desarrollo de las

adicciones (Giancola & Tarter, 1999; Verdejo-García et al., 2008). Sin embargo, esta noción

no es incompatible con la evidencia empírica que indica que el consumo prolongado de

drogas puede generar alteraciones neuropsicológicas en individuos drogodependientes.

Por otro lado, el estudio de las alteraciones neuropsicológicas asociadas al consumo de

drogas y sus correlatos neurobiológicos puede verse afectado por varias limitaciones

metodológicas inherentes al contexto de la investigación en drogodependencias que

dificultan en gran medida la interpretación de los resultados obtenidos y que pueden dar

lugar a inconsistencias y contradicciones. En este sentido, es importante destacar el hecho

de que la mayoría de los individuos drogodependientes son policonsumidores de drogas y

suelen presentar patrones de intensidad, frecuencia y duración del consumo o de

duración de la abstinencia muy heterogéneos. El policonsumo y en particular la

combinación de drogas ilegales con alcohol y, en ocasiones, con medicamentos y

sustancias no reguladas, se ha convertido en la pauta dominante del consumo de drogas en

Europa (EMCDDA, 2011), por lo que la investigación de los efectos diferenciales de

distintas drogas de abuso está frecuentemente limitada por el hecho de que la práctica

totalidad de los individuos drogodependientes no son consumidores exclusivos de una

P á g i n a | 3 7

única sustancia, sino policonsumidores de diversas sustancias (Fernández-Serrano et al.,

2011; Verdejo-García et al., 2004).

Aunque puede resultar complejo afrontar diseños de investigación capaces de

controlar el conjunto de las limitaciones mencionadas, el conocimiento y la consideración

de estas limitaciones en la interpretación de los resultados puede incrementar

significativamente la capacidad explicativa de las investigaciones en el ámbito de las

drogodependencias.

En el contexto de aportar conocimientos sobre esta temática y superar las

limitaciones existentes en estudios previos de adicción, la presente Tesis Doctoral tiene

como Objetivo General el estudio de las alteraciones cerebrales estructurales y

funcionales asociadas con variables de personalidad que incrementan la predisposición al

riesgo de consumo y la dependencia, con el consumo crónico de diferentes drogas y con el

funcionamiento neuropsicológico y emocional de un grupo de pacientes

drogodependientes en situación de abstinencia.

De este Objetivo General se derivaron tres Objetivos Específicos, que se articularon

a través de tres estudios. En primer lugar, se abordaron las posibles alteraciones

estructurales de pacientes con consumo principal de cocaína. En este sentido, el Primer

Objetivo fue estudiar la asociación entre una medida multidimensional de impulsividad

rasgo y las alteraciones de la estructura cerebral presentadas por un grupo de pacientes

policonsumidores de drogas con consumo principal de cocaína. Una vez estudiadas las

alteraciones estructurales encontradas en esta población y su asociación con medidas de

personalidad, los dos siguientes estudios se dirigieron a investigar las alteraciones

funcionales asociadas con la severidad del consumo (estudio 2) y las alteraciones

neuropsicológicas encontradas en esta población (estudio 3). Por tanto, el Segundo

Objetivo fue estudiar la asociación entre la severidad del consumo de diversas drogas y el

metabolismo cerebral en reposo de un grupo de pacientes policonsumidores de drogas en

P á g i n a | 3 8

situación de abstinencia y el Tercer Objetivo estudiar la asociación del funcionamiento

neuropsicológico y el metabolismo cerebral en situación de reposo en este mismo grupo

de policonsumidores de drogas.

Las Hipótesis que se derivaron de cada uno de estos Objetivos fueron:

Para el Primer Objetivo: Esperamos encontrar una asociación significativa entre la

personalidad impulsiva y el volumen cerebral de la corteza prefrontal y los ganglios

basales específica de los individuos consumidores de cocaína.

Para el Segundo Objetivo: Esperamos encontrar una asociación generalizada de la

severidad del consumo de todas las drogas estudiadas con el metabolismo cerebral de

regiones fronto-estriadas y el cerebelo en individuos policonsumidores, así como una

asociación específica del consumo de cocaína con la corteza parietal y del de heroína con la

corteza temporal una vez controlado el co-abuso de las demás drogas consumidas.

Para el Tercer Objetivo: Esperamos encontrar una asociación entre el rendimiento

en pruebas de función ejecutiva “frías” y el metabolismo cerebral de la corteza prefrontal

dorsolateral, la parte anterior del cíngulo y estructuras temporales y entre pruebas de

función ejecutiva “caliente” y el metabolismo cerebral de la corteza orbitofrontal, la parte

anterior del cíngulo y el sistema límbico en el grupo de policonsumidores.

El seguimiento de estos tres Objetivos e Hipótesis ha dado lugar a la realización de

tres trabajos de investigación y a la publicación de tres Manuscritos en revistas

internacionales de alto impacto tras un proceso de revisión por pares y cuyas referencias

se citan a continuación:

Moreno-López, L., Catena, A., Fernández-Serrano, M.J., Delgado-Rico, E., Stamatakis,

E.A., Pérez-García, M., Verdejo-García, A. (2012). Trait impulsivity and prefrontal

gray matter reductions in cocaine dependent individuals. Drug and Alcohol

Dependence, en prensa.

P á g i n a | 3 9

Moreno-López, L., Stamatakis, E.A., Fernández-Serrano, M.J., Gómez-Río, M.,

Rodríguez-Fernández, A., Pérez-García, M., Verdejo-García, A. (2012) Neural

Correlates of the Severity of Cocaine, Heroin, Alcohol, MDMA and Cannabis Use

in Polysubstance Abusers: A Resting-PET Brain Metabolism Study. PLoS ONE,

en prensa.

Moreno-López, L., Stamatakis, E.A., Fernández-Serrano, M.J., Gómez-Río, M.,

Rodríguez-Fernández, A., Pérez-García, M., Verdejo-García, A. (2012) Neural

correlates of hot and cold executive functions in polysubstance addiction:

Association between neuropsychological performance and resting-PET brain

metabolism. Psychiatry Research: Neuroimaging, en prensa.

P á g i n a | 4 0

P á g i n a | 4 1

[MEMORIA DE TRABAJOS]

P á g i n a | 4 2

P á g i n a | 4 3

* ARTÍCULO 1

Trait impulsivity and prefrontal gray matter reductions in cocaine dependent individuals

* Moreno-López, L., Catena, A., Fernández-Serrano, M. J., Delgado-Rico, E., Stamatakis, E. A.,

Pérez-García, M., & Verdejo-García, A. (2012). Trait impulsivity and prefrontal gray matter

reductions in cocaine dependent individuals. Drug and Alcohol Dependence, en prensa.

P á g i n a | 4 4

1. Introduction

Cocaine addiction is a worldwide major public health problem for which current

prevention and treatment options are not fully satisfactory (EMCDDA, 2009; Substance

Abuse and Mental Health Services Administration, 2010). Neuroimaging studies in

cocaine-dependent individuals have revealed significant brain volume reductions, most

notably in a priori selected regions of interest, such as the striatum, the amygdala or the

prefrontal cortex (Barrós-Loscertales et al., 2011; Makris et al., 2004; Matochik et al.,

2003; Tanabe et al., 2009). These regions are thought to contribute to critical aspects of

the addictive cycle, including reinforcement learning, craving, and inhibitory control

(Koob and Volkow, 2010). Nonetheless, the findings from nonbiased automated

techniques, such as Voxel Based Morphometry (VBM), have yielded inconsistent results.

Despite previous positive findings showing significant gray matter (GM) reductions in

cocaine-dependent individuals, including the medial prefrontal cortex, superior temporal

cortex, insula, thalamus and cerebellum (Franklin et al., 2002; Sim et al., 2007), a recent

VBM study failed to detect any structural change (GM or white matter –WM– reductions)

in cocaine users compared to non-drug using controls (Narayana et al., 2010). Perhaps

more critically, most of these studies have failed to find any correlation between estimates

of drug use patterns (e.g., amount, duration or age at onset) and GM and WM reductions

(Franklin et al., 2002; Makris et al., 2004; Matochik et al., 2003; but see Ersche et al., 2011).

This apparent lack of association between cocaine exposure and brain attrition raises the

possibility that certain personality traits, such as impulsivity, may relate to GM and WM

abnormalities in cocaine-dependent individuals.

Impulsivity is viewed as a multifaceted trait that varies normally across the

population, but that in high levels may predispose to a range of dysfunctional behaviors,

including addiction (Cyders et al., 2007; Verdejo-García et al., 2008). Animal studies have

shown that increased impulsivity is associated with reduced dopamine receptors

P á g i n a | 4 5

availability and cocaine use escalation and progression to dependence (Belin et al., 2008;

Dalley et al., 2007). In humans, high levels of trait impulsivity (indexed with the Barratt

Impulsivity Scale) are associated with lower midbrain dopamine autoreceptor binding,

and greater amphetamine-induced dopamine release in the striatum (Buckholtz et al.,

2010). In addition, trait impulsivity is negatively correlated with orbitofrontal cortex

(OFC) GM volumes in healthy individuals (Matsuo et al., 2009). Nonetheless, in addicted

individuals, continued use of stimulants is thought to further exacerbate impulsive traits

(Ersche et al., 2010), and to possibly modify its neural underpinnings. In fact, the

association between trait impulsivity levels and striatal dysfunction is stronger in

methamphetamine-dependent individuals compared to healthy controls (Lee et al., 2009).

Conversely, Ersche et al. (2011) failed to find significant correlations between particular

aspects of trait impulsivity (impulsive reward-seeking) and GM reductions in cocaine-

dependent subjects. Therefore, more studies are needed to specifically explore if, as

expected, different facets of trait impulsivity are specifically linked to GM volumes in

frontostriatal regions in cocaine-dependent individuals.

The aims of this study were (i) to use whole-brain VBM analyses to examine

possible GM and WM reductions in a sample of currently abstinent (>1 month) cocaine

dependent individuals, as compared to a non-drug using control group; and (ii) to

measure differences in the way impulsivity relates to GM and WM volumes in cocaine

users vs. controls. Because impulsive personality in the normal population is negatively

associated with GM volume in the OFC (Matsuo et al., 2009), which is also impacted by

lifetime cocaine use (Alia-Klein et al., 2011), we expected cocaine-dependent individuals to

have reduced GM volumes in the OFC and richly interconnected regions, such as insula,

amygdala, striatum, and WM adjacent to these regions (Bechara, 2005). Because

progression of cocaine use is thought to provoke neuroadaptations from the ventral

striatum to the dorsal striatum, and from the OFC to more extensive prefrontal regions

like the anterior cingulate and dorsolateral prefrontal cortex (Koob and Volkow, 2010), we

P á g i n a | 4 6

expected impulsivity to be uniquely associated with more lateral aspects of the prefrontal

cortex and more posterior aspects of the striatum in cocaine-dependent individuals.

2. Materials and Methods

2.1. Participants

Thirty-eight cocaine-dependent individuals, mean age=29.58, SD=6.53, and 38

non-drug using controls, mean age=31.08, SD=5.14, participated in this study. All

participants were male due to the low prevalence of women entering drug treatment

during the recruitment period. Cocaine users were recruited in an inpatient therapeutic

community (“Proyecto Hombre”), in the city of Granada, Spain. All of them reported

cocaine as their main drug of choice and the one for which they requested treatment.

Clinical interviews based on Diagnostic and Statistical Manual version IV (DSM-IV) criteria

confirmed cocaine dependence diagnosis; nonetheless, they also had regular use of

tobacco, alcohol, cannabis and MDMA (see Table 1). To enter the study, cocaine users had

to be abstinent for at least 30 days (for any drug but tobacco), as confirmed by weekly

urine tests; in this way, we could rule out acute and residual effects of previously used

drugs on brain structure, with the exception of nicotine (84.2% of cocaine-dependent

individuals and 44.7% of controls were current smokers). None of the cocaine patients

were currently following pharmacological substitution treatments. Potential participants

who had previously been diagnosed with any disorder from DSM- IV Axes I and II (other

than substance dependence), or had neurological or systemic diseases affecting central

nervous system (CNS) functioning were excluded.

Non-drug using controls were recruited through adverts distributed by a local

employment agency, and therefore they were also matched to cocaine participants in

terms of unemployment status. Selection criteria for control participants were: (i) absence

of current or past substance use, excluding past or current social drinking (less than ten

standard alcohol units per week), (ii) absence of documented major psychiatric disorders,

P á g i n a | 4 7

(iii) absence of documented head injury or neurological disorder, and (iv) not being taking

medication with effects on the CNS.

For both groups, evidence of stroke or space-occupying lesions observed on

conventional clinical MR images, any contraindications to MRI scanning (including

claustrophobia and implanted ferromagnetic objects), and history of loss of consciousness

(LOC) for longer than 30 minutes or LOC with any neurological consequence were

exclusionary.

2.2. Instruments and assessment procedures

The study was approved by the Ethics Committee for Research in Humans of the

University of Granada. All participants signed an informed consent form certifying their

voluntary participation. Controls, but not patients, received a €40 compensation for

participating in the study. Assessments were conducted across two sessions separated by

less than one week. During the first session we administered the Interview for Research

on Addictive Behavior, the Hamilton Rating Scales for Depression and Anxiety (HAM-D

and HAM-A), and the UPPS-P Impulsive Behavior Scale, along with a battery of cognitive

tests which results will be reported separately. The second session involved the MRI

scanning. The MRI scans lasted approximately 6 minutes.

2.2.1. Patterns of Drug Use

Data regarding lifetime amount and duration of use of the different drugs was self-

reported by participants and collected using the Interview for Research on Addictive

Behavior (Verdejo-García et al., 2005). This interview provides an estimation of monthly

use of each substance during regular use (amount per month) and total duration of use of

each substance (in years). The descriptive scores for these variables in the sample are

presented in Table 1.

P á g i n a | 4 8

Table 1 Descriptive information about patterns of drug use in cocaine-dependent individuals.

Substance Ever used (%) Unit Amount per month Duration of use Abstinence

Cocaine 100 # Grams 18.92 ± 29.46 4.05 ± 3.07 36.76 ± 22.60

Cannabis 76.31 # Joints 109.90 ± 111.40 2.35 ± 3.17 147.45 ± 203.66

MDMA 50 # Tablets 9.29 ± 10.12 2.27 ± 1.90 227.68 ± 225.78

Alcohol 100 # SAU 107.58 ± 114.03 7.38 ± 5.77 33.11 ± 21.77

Tobacco 86.8 # Cigarettes 488.48 ± 307.40 9.19 ± 7 7.8 ± 15.82

Note: Mean ± Standard deviation of the mean. Duration of use is expressed in years. Abstinence duration is expressed in number of weeks.

P á g i n a | 4 9

2.2.2. Trait Impulsivity

We used the Spanish version of the UPPS-P Scale (Verdejo-García et al., 2010;

Whiteside and Lynam, 2001). This is a 59-item self-report inventory designed to measure

five distinct personality pathways to impulsive behavior: sensation seeking, (lack of)

perseverance, (lack of) premeditation, negative urgency, and positive urgency. The first 4

dimensions were included in the original version of the UPPS scale (Whiteside and Lynam,

2001); the fifth dimension has been included based on recent work by (Cyders et al., 2007;

Smith et al., 2007). Sensation seeking (12 items) incorporates two aspects: 1) a tendency

to enjoy and pursue activities that are exciting, and 2) an openness to trying new

experiences that may or may not be dangerous; (lack of) perseverance (10 items) refers to

the individual’s ability to remain focused on a task that may be boring or difficult; (lack of)

premeditation (11 items) refers to the tendency to think and reflect on the consequences

of an act before engaging in that act; and finally urgency (12 items) refers to the tendency

to experience strong impulses under conditions of negative affect (negative urgency, 12

items) or positive affect (positive urgency, 14 items). Each item on the UPPS is rated on a

4-point scale ranging from 1 (strongly agree) to 4 (strongly disagree). We obtained the

total scores of each of these five UPPS-P dimensions for analysis.

2.3. MRI acquisition

Participants were scanned on a 3T whole body MRI scanner (Phillips Achieva)

operating with 8 channels phased-array head coil for reception. For each participant, a T1-

weighted 3D volume was acquired using a T1-weighted 3D-turbo-gradient-echo sequence

(3D-TFE), in sagittal orientation with 0.94x0.94x1.0 mm resolution (160 slides,

FOV=240x240 mm2, matrix 256x256x160) with repetition time of 8 ms, echo time 4 ms,

inversion delay = 1022.6264 ms, flip angle of 8°, band with 191 Hz/pixel. The sequence

was optimal for reducing motion sensitivity, susceptibility artifacts and field

inhomogeneities.

P á g i n a | 5 0

2.4. Image analysis

Before automatic preprocessing, the images were checked for artifacts and

manually aligned to the AC-PC line. Data were processed and analyzed using SPM8

(http://www.fil.ion.ucl.ac.uk/spm) and VBM8 toolbox (http://dbm.neuro.unijena.

de/vbm.html), for which we used the default parameters. Thus, within a unified

segmentation model (Ashburner and Friston, 2005), images were corrected for biasfield

inhomogeneities, registered using linear (12 parameters affine) and non-linear

transformations (warped), and tissue was clustered into GM, WM, and cerebrospinal fluid

(CSF). Segments were further refined by using adaptive maximum a posteriori

estimations, which account for partial volume effects, and by applying a hidden Markov

random field model, as implemented in VBM8. Importantly, to preserve local GM/WM

values, we multiplied the segments by the Jacobian determinants of the deformation field

to create modulated images. Segments were smoothed by an 8 mm full-width at half-

maximum (FWHM) using an isotropic Gaussian kernel. Afterwards, we conducted analysis

on modulated GM and WM segments.

2.4.1. Global effects of patterns of drug use

Total gray and white matter volumes (TGMV, TWMV) were computed by adding up

the voxel values of GM/WM segments, including cerebellum and sub-cortical structures.

Next, the total brain volume (TBV=TGMV+TWMV) was used for computing the TGMV/TBV

ratios, which served to evaluate the global effects of patterns of drug use on brain tissue

volumes using the backward step-wise procedure in the general linear model, including as

predictors age, years of education, cocaine amount per month during regular use, duration

of cocaine use, and age of onset of cocaine use. T-tests were used to estimate the

significance of the regression parameters. Significance threshold was set at p<0.05.

P á g i n a | 5 1

2.4.2. Regional GM and WM differences between cocaine-dependent individuals and non-

drug using controls

The general linear model implemented in SPM8 was used to conduct voxel-wise

comparisons between cocaine-dependent individuals and controls, and to determine the

relationship between consume severity factors (i.e. amount, duration and age of onset)

and regional GM and WM volume variations in cocaine-dependent individuals. In both

analyses the total volume of GM, or WM, was modeled as a linear confound in order to

account for residual global volume variability. The significance threshold was set at p<0.05

after family-wise correction for multiple comparisons (pFWE<0.05). Significant peaks

from the t-test maps are given in MNI space.

2.4.3. Regional relationships between GM and WM and measures of impulsivity and

estimates of drug use

We performed SPM8 multiple regression analysis, in which we tested the

significances of the Group (cocaine-dependent individuals vs. controls) x UPPS-P

dimensions interactions across the entire brain. We only included the UPPS-P dimensions

showing significant differences between the groups. The differences in the regression

coefficients as a function of group were tested using a significance threshold of p<0.05

after family-wise correction for multiple comparisons (pFWE<0.05).

3. Results

3.1. Participants’ characteristics

Both groups were matched on gender, ethnicity and language, and had statistically

equivalent distributions for age. Controls, compared to cocaine-dependent individuals, had

more years of education [(M=17.58, SD=4.56) vs. (M=12.03, SD=3.62)]; but analysis

including education as a covariate did not alter the results. Cocaine-dependent individuals

had greater scores for depression [(M=5.08, SD=3.98) vs. (M=1.32, SD=2.15)] and anxiety

P á g i n a | 5 2

[(M=7.05, SD=6.40) vs. (M=1.29, SD=2.29)], but on average both groups were below cut-

offs for clinical significance.

3.2. Trait Impulsivity

Cocaine-dependent individuals had higher scores than controls across the five

dimensions of the UPPS-P Scale. T-tests showed these differences were significant for lack

of premeditation (t=2.46, p=0.016, Cohen´s d=0.6), lack of perseverance (t=2.56, p=0.013,

Cohen´s d=0.6), negative urgency (t=5.14, p=0.000, Cohen´s d=1.2), and positive urgency

(t=5.86, p=0.000, Cohen´s d=1.4). We found no significant differences for sensation

seeking, such that this dimension was not further used in the subsequent analysis.

Table 2 Descriptive scores –means and standard deviations (in parentheses) of cocaine

users and non-drug using controls in the five UPPS-P dimensions.

UPPS-P Dimensions Cocaine-dependent

individuals Non-drug using

Controls Test

Negative Urgency 31.45 (6.79) 23.16 (6.99) (t46 = 5.14, p = 0.000)

Positive Urgency 32.91 (8.79) 21.66 (7.63) (t46 = 5.86, p = 0.000)

Lack of Premeditation 23.83 (5.95) 21 (3.68) (t46 = 2.46, p = 0.016)

Lack of Perseverance 21.26 (4.42) 18.84 (3.63) (t46 = 2.56, p = 0.013)

Sensation Seeking 30.63 (8.73) 29.89 (7.37) (t46 = 0,34, p = 0.73)

3.3. Imaging analysis

3.3.1. Global effects

For cocaine-dependent individuals, the TGMV/TBV ratio variability was accounted

for by monthly cocaine intake during regular use, t(34)=2.19, p=0.035, and age of onset,

t(34)=-2.33, p=0.026; the effect of duration of cocaine use also showed a trend to

significance, t(34)=-1.80, p=0.080. These variables explained about a 32% of the

P á g i n a | 5 3

variability in TGMV/TBV ratios, F(3,34)=5.21, p=0.005. Noteworthy, higher amounts of

cocaine (r=0.41, p=.01) and younger age of onset of cocaine use (r=-.34, p=0.04) were

associated with larger TGMV/TBV. The first result is associated with a significant decrease

in WM volumes (TWMV-amount of cocaine: r=0.43, p=0.006), but the second showed a

trend in relation to reductions in GM (TGMV-age of onset of cocaine abuse: r=-0.18, NS).

3.3.2. Regional GM and WM differences between cocaine-dependent individuals and non-

drug using controls

The cocaine group had significantly lower GM and WM volumes than controls.

Cocaine-dependent individuals showed lower GM volumes in a number of sections of the

OFC (left superior and medial frontal gyrus), right inferior frontal gyrus, right insula, left

amygdala and parahippocampal gyrus, left inferior and middle temporal gyrus, and

bilateral caudate (Table 3 and Figure 1). Likewise, cocaine dependent individuals showed

lower WM volumes in the left inferior and medial frontal gyrus, superior temporal gyrus,

right anterior cingulate cortex, insula and caudate (Table 4 and Figure 2).

P á g i n a | 5 4

Table 3 Summary of the results obtained by the gray matter voxel-wise analysis of volume reductions in cocaine-dependent individuals and non-

drug using controls.

Anatomical Region k Talairach Labels for peaks BA T p x y z

Frontal Lobe 3342 L Medial Frontal Gyrus Cingulate Gyrus

11, 10 32, 25, 24

5.81 5.75

0.004 0.005

-9 27 -13

Frontal Lobe Caudate

2802 R Inferior Frontal Gyrus R Caudate Head 47, 11, 34

6.39 5.53

0.000 0.010

28 9

21 19

-15 -15

Caudate 1118 L Caudate Head 5.52 0.010 6 -1 3 Limbic Lobe / Parahippocampal Gyrus

972 L Amygdala/Hippocampus 6.42 0.000 -33 -7 -19

Sub-Lobar 581 R Insula 13 5.81 0.004 52 -19 23

Temporal Lobe 3095 L Middle/Inferior Temporal Gyrus 21, 22, 20 5.98 0.002 -56 -14 -18

P á g i n a | 5 5

Figure 1 Brain regions showing significant gray matter decreases in cocaine-dependent

individuals with respect to non-drug using controls.

Table 4 Summary of the results obtained by the white matter voxel-wise analysis of

volume reductions in cocaine-dependent individuals and non-drug using controls.

Anatomical Region k Talairach Labels for peaks T p x y z

Frontal Lobe 1181 L Inferior Frontal Gyrus Sub-Gyral

6.07 0.001 -26 18 -17

Frontal Lobe 623 L Medial Frontal Gyrus 5.46 0.003 -20 57 -3

Temporal Lobe 389 L Superior Temporal Gyrus 5.24 0.007 -48 -25 -3

Frontal Lobe 1651 R Anterior Cingulate R Medial Frontal Gyrus

6.81 0.001 7 33 -15

Sub-Lobar 1531 Caudate Extra-Nuclear Lentiform Nucleus

5.86 0.001 12 12 -3

Temporal Lobe 674 Insula R Superior Temporal Gyrus R Transverse Temporal Gyrus

5.82 0.001 46 -26 12

Figure 2 Brain regions showing significant white matter decreases in cocaine-dependent

individuals with respect to non-drug using controls.

P á g i n a | 5 6

3.3.3. Regional relationships between GM and WM and measures of impulsivity and

estimates of drug use

We found a number of regions in which GM volumes showed differential

associations with the impulsivity dimensions of lack of premeditation and negative

urgency; in all cases, there was a significant Group x Impulsivity interaction (see Table 5).

The associations between impulsivity levels and GM volumes were significantly different

for cocaine-dependent individuals vs. controls in the left inferior frontal gyrus, right insula

and left putamen (in relation to lack of premeditation), and left middle frontal gyrus and

sub-gyral (in relation to negative urgency). Left inferior frontal gyrus was positively

associated with lack of premeditation in cocaine users, whereas it was negatively

correlated with this dimension in controls. Conversely, right insula and left putamen were

negatively correlated with lack of premeditation in cocaine users, but did not correlate

with this dimension in controls. With regard to negative urgency, the subgyral showed

significant positive correlations in the cocaine group, whereas the association was

negative in controls. Finally, in the case of the left middle frontal gyrus, none of the slopes

significantly differed from zero, but the trend was positive in cocaine-dependent

individuals, and negative in controls.

We did not find significant associations between WM volumes and the different

dimensions of impulsivity in the cocaine or control groups. Finally, we did not find

significant associations between the main patterns of drug use (amount, duration and age

of onset of cocaine abuse) and GM and WM volumes.

P á g i n a | 5 7

Table 5 Regression slopes showing different associations between impulsivity and gray matter volumes between cocaine-dependent individuals

(CDI) and non-drug-using controls (NDC) (pFWE=0.05).

Anatomical Region k Talairach Label BA p (β) Slopes Volume x y z

Lack of premeditation

Sub-lobar 179 R Insula BA 47/BA 13 0.02 CDI-/NDC+ CDI<NDC 30 10 -3

358 L Putamen 0.05 CDI-/NDC+ CDI<NDC -21 10 -4

Fontal Lobe 249 L Inferior Frontal Gyrus BA 10 0.005 CDI+/NDC- CDI<NDC -37 41 1

Negative Urgency

Frontal Lobe 228 L Middle Frontal Gyrus BA10/46 0.038 CDI+/NDC- CDI<NDC -39 37 19

381 R Sub-Gyral BA 8 0.024 CDI+/NDC- CDI<NDC 25 24 36

P á g i n a | 5 8

4. Discussion

Our results showed that cocaine-dependent individuals, compared to controls,

have lower GM volumes in the OFC cortex, anterior cingulate, inferior frontal gyrus, insula,

amygdala, temporal gyrus, and caudate. We also found WM reductions in adjacent regions

of the anterior cingulate, inferior/middle frontal gyrus, insula and putamen. As expected,

we found significant differences in the way that impulsivity dimensions correlates with

GM volumes in cocaine users vs. controls: cocaine patients had elevated levels of trait

impulsivity, and these scores were differentially associated with GM volumes in the left

inferior frontal gyrus, insula and putamen (lack of premeditation), and left middle frontal

gyrus (negative urgency). Inferior and medial superior frontal clusters positively

correlated with impulsivity in cocaine users, and negatively in controls; conversely, insula

and putamen correlated negatively with impulsivity in cocaine patients and showed no

correlations in controls. These results point to a distinctive link between trait impulsivity

and regional GM among cocaine-dependent individuals, which specifically relates to

frontostriatal systems regions. The measures of patterns of cocaine use correlated with

general indices of GM/TBV ratios but had no significant influence on GM volumes at the

stricter voxel level. Overall, our findings are in agreement with the notion that GM

reductions in cocaine-dependent individuals are at least partly mediated by trait

impulsivity.

Our findings of GM reductions in cocaine-dependent individuals are largely

consistent with previous neuroimaging studies in revealing structural abnormalities in

brain regions relevant to reinforcement learning (amygdala, OFC) and inhibitory control

(anterior cingulate, inferior frontal gyrus, insula and caudate) (Ersche et al., 2011;

Franklin et al., 2002; Makris et al., 2004; Matochik et al., 2003). Recent findings from fMRI

connectivity analyses also provide support to the notion that neural networks connecting

the medial frontal cortex with paralimbic and temporal regions are significantly less

P á g i n a | 5 9

functional in cocaine users (Gu et al., 2010). Biological abnormalities in these regions are

thought to underlie psychological alterations leading to overestimation of the motivational

relevance of drug-related reinforcement (the impulsive system) and to breakdown of

inhibitory processes necessary to achieve long-term abstinence (the reflective system)

(Bechara, 2005; Goldstein and Volkow, 2002; Redish et al., 2008; Verdejo-García and

Bechara, 2009). These alterations are proposed to anchor pervasive drug taking in the face

of increasing aversive consequences, one of the key features of addiction (DSM-IV).

A key unresolved issue is that if these structural abnormalities are reflective of

cocaine induced brain attrition, or are related to personality traits, or reflect a

combination of both factors. Our findings, in agreement with growing neuroscience

evidence, are supportive of the impact of personality traits. However, these traits may

interact with cocaine use in producing GM alterations, as suggested by the global TGMV

analysis. Recent studies indicate that, in healthy individuals, impulsivity is negatively

correlated with midbrain and striatal D2/D3 receptors availability (Buckholtz et al., 2010;

Lee et al., 2009), but this association is stronger in methamphetamine dependent

individuals (Lee et al., 2009) and correlates with amphetamine-induced dopamine release

in the striatum (Buckholtz et al., 2010). Striatal D2 availability is associated with metabolic

rate in the prefrontal cortex of psychostimulant users (Volkow et al., 2001). Therefore,

different strands of evidence indicate that frontostriatal systems mediate impulsive

behavior and thereby promote psychostimulant addiction. Nonetheless, we acknowledge

that these issues can only be tested by longitudinal imaging studies in individuals at high

risk of developing psychostimulant dependence, or by gene association imaging studies

(Kreek et al., 2005; Verdejo-García et al., 2008).

In accordance with the notion that impulsivity is a multifaceted construct (Smith et

al., 2007; Whiteside and Lynam, 2001), our results showed differential links between

impulsivity dimensions and GM volumes. Lack of premeditation was positively associated

P á g i n a | 6 0

with left inferior frontal gyrus GM in cocaine users but not controls. This region is

importantly involved in the representation of expected values (Bermpohl et al., 2010);

such that persistent stimulation by drug-expected rewards could have induced GM

enlargements linked to an increased tendency to rapidly engage in non-adequately

forethought behavior. On the other hand, both insula and putamen have been involved in

drug-related neuroadaptations related to attentional bias and craving (Childress et al.,

2008; Luijten et al., 2011; Volkow et al., 2006), such that reduced volumes may predict

increased prepotency of drug-related action tendencies. On the other hand, negative

urgency was positively associated with GM in subgyral BA 8 in cocaine users, negatively in

controls. This region has been previously associated with the experience of uncertainty

(Volz et al., 2003), which is a key aspect of the associative learning processes mediating

expected drug effects (D’Souza and Duvauchelle, 2008). In fact, adjacent regions of the

medial frontal gyrus (BA 10) display increased activation during uncertain outcome

selection in psychostimulant users (Leland et al., 2006). Therefore, in this case the

rationale could be similar to that posited for inferior frontal gyrus; repeated exposure to

uncertainty-loaded scenarios may be associated with increased subgyral volumes and

higher risk to trigger impulsive acts when under emotionally uneasy conditions.

Strengths of this study include the adequate sample size, the sound clinical

characterization and the agreement between results from the different analytical

approaches (between-group differences and regressions testing the link between

impulsivity and GM). On the other hand, we should also acknowledge a number of

limitations. The cross-sectional design precludes us from drawing conclusions about the

causality of GM deficits; one way out of addressing this issue would be by using

longitudinal designs, but these studies are costly and convey important ethical concerns

and methodological complexities. Furthermore, we cannot completely rule out the

influence of depression and anxiety levels, although they were on average below the cut-

off for clinical significance and they had very low impact on regression models when

P á g i n a | 6 1

included. Finally, although we discuss our results as pertaining to cocaine dependence,

polysubstance abuse is almost ubiquitous among cocaine-dependent individuals.

However, the current sample was carefully selected for relative specificity of cocaine

related problems and low use of other drugs. The illegal drugs more frequently and

intensely co-abused in the sample were cannabis and MDMA, but estimates of the use of

these drugs failed to predict regional GM.

P á g i n a | 6 2

References

Alia-Klein, N., Parvaz, M.A., Woicik, P.A., Konova, A.B., Maloney, T., Shumay, E., et al., 2011.

Gene x disease interaction on orbitofrontal gray matter in cocaine addiction. Arch Gen

Psychiatry 68, 283-294.

Ashburner, J., Friston, K.J., 2005. Unified segmentation. Neuroimage 26, 839-851.

Barrós-Loscertales, A., Garavan, H., Bustamante, J.C., Ventura-Campos, N., Llopis, J.J.,

Belloch, V., et al., 2011. Reduced striatal volume in cocaine-dependent patients.

Neuroimage 56, 1021-1026.

Bechara, A., 2005. Decision making, impulse control and loss of willpower to resist drugs:

a neurocognitive perspective. Nat Neurosci. 8, 1458-1463.

Belin, D., Mar, A.C., Dalley, J.W., Robbins, T.W., Everitt, B.J., 2008. High impulsivity predicts

the switch to compulsive cocaine-taking. Science 320, 1352-1355.

Bermpohl, F., Kahnt, T., Dalanay, U., Hägele, C., Sajonz, B., Wegner, T., et al., 2010. Altered

representation of expected value in the orbitofrontal cortex in mania. Hum Brain Mapp.

31, 958-969.

Buckholtz, J.W., Treadway, M.T., Cowan, R.L., Woodward, N.D., Li, R., Ansari, M.S., et al.,

2010. Dopaminergic network differences in human impulsivity. Science 329, 532.

Childress, A.R., Ehrman, R.N., Wang, Z., Li, Y., Sciortino, N., Hakun, J., et al., 2008. Prelude to

passion: limbic activation by "unseen" drug and sexual cues. Plos One 3, e1506.

Cyders, M.A., Smith, G.T., Spillane, N.S., Fischer, S., Annus, A.M., Peterson, C., 2007.

Integration of impulsivity and positive mood to predict risky behavior: development and

validation of a measure of positive urgency. Psychol Assess. 19, 107-118.

P á g i n a | 6 3

Dalley, J.W., Fryer, T.D., Brichard, L., Robinson, E.S., Theobald, D.E., Lääne, K., et al., 2007.

Nucleus accumbens D2/3 receptors predict trait impulsivity and cocaine reinforcement.

Science 315, 1267-1270.

D'Souza, M.S., Duvauchelle, C.L., 2008. Certain or uncertain cocaine expectations influence

accumbens dopamine responses to self-administered cocaine and non-rewarded operant

behavior. Eur Neuropsychopharmacol. 18, 628-638.

EMCDDA European Monitoring Centre for Drugs and Drug Addiction (2009): Annual

report 2009: the state of the drugs problem in Europe Luxembourg. Publications Office of

the European Union.

Ersche, K.D., Barnes, A., Simon Jones, P., Morein-Zamir, S., Robbins, T.W., Bullmore, E.T.,

2011. Abnormal structure of frontostriatal brain systems is associated with aspects of

impulsivity and compulsivity in cocaine dependence. Brain 134, 2013-2024.

Ersche, K.D., Bullmore, E.T., Craig, K.J., Shabbir, S.S., Abbott, S., Müller, U., et al., 2010.

Influence of compulsivity of drug abuse on dopaminergic modulation of attentional bias in

stimulant dependence. Arch Gen Psychiatry 67, 632-644.

Franklin, T.R., Acton, P.D., Maldjian, J.A., Gray, J.D., Croft, J.R., Dackis, C.A., et al., 2002.

Decreased gray matter concentration in the insular, orbitofrontal, cingulate, and temporal

cortices of cocaine patients. Biol Psychiatry 51, 134-142.

Goldstein, R.Z., Tomasi, D., Alia-Klein, N., Zhang, L., Telang, F., Volkow, N.D., 2007. The

effect of practice on a sustained attention task in cocaine abusers. Neuroimage 35, 194-

206.

Goldstein, R.Z., Volkow, N.D., 2002. Drug addiction and its underlying neurobiological

basis: neuroimaging evidence for the involvement of the frontal cortex. Am J Psychiatry

159, 1642-1652.

P á g i n a | 6 4

Gu, H., Salmeron, B.J., Ross, T.J., Geng, X., Zhan, W., Stein, E.A., et al., 2010.

Mesocorticolimbic circuits are impaired in chronic cocaine users as demonstrated by

resting-state functional connectivity. Neuroimage 53, 593-601.

Koob, G.F., Volkow, N.D., 2010. Neurocircuitry of addiction. Neuropsychopharmacology 35,

217-238.

Kreek, M.J., Bart, G., Lilly, C., LaForge, K.S., Nielsen, D.A., 2005. Pharmacogenetics and

human molecular genetics of opiate and cocaine addictions and their treatments.

Pharmacol Rev. 57, 1-26.

Lee, B., London, E.D., Poldrack, R.A., Farahi, J., Nacca, A., Monterosso, J.R., et al., 2009.

Striatal dopamine d2/d3 receptor availability is reduced in methamphetamine

dependence and is linked to impulsivity. J Neurosci. 29, 14734-14740.

Leland, D.S., Arce, E., Feinstein, J.S., Paulus, M.P. 2006. Young adult stimulant users'

increased striatal activation during uncertainty is related to impulsivity. Neuroimage 33,

725-731.

Luijten, M., Veltman, D.J., van den Brink, W., Hester, R., Field, M., Smits, M., et al., 2011.

Neurobiological substrate of smoking-related attentional bias. Neuroimage 54, 2374-2381.

Makris, N., Gasic, G.P., Seidman, L.J., Goldstein, J.M., Gastfriend, D.R., Elman, I., et al., 2004.

Decreased absolute amygdala volume in cocaine addicts. Neuron 44, 729-740.

Matochik, J.A., London, E.D., Eldreth, D.A., Cadet, J.L., Bolla, K.I., 2003. Frontal cortical tissue

composition in abstinent cocaine abusers: a magnetic resonance imaging study.

Neuroimage 19, 1095-1102.

Matsuo, K., Nicoletti, M., Nemoto, K., Hatch, J.P., Peluso, M.A., Nery, F.G., et al., 2009. A voxel-

based morphometry study of frontal gray matter correlates of impulsivity. Hum Brain

Mapp. 30, 1188-1195.

P á g i n a | 6 5

Narayana, P.A., Datta, S., Tao, G., Steinberg, J.L., Moeller, F.G., 2010. Effect of cocaine on

structural changes in brain: MRI volumetry using tensor-based morphometry. Drug

Alcohol Depend. 111, 191-199.

Redish, A.D., Jensen, S., Johnson, A., 2008. A unified framework for addiction:

vulnerabilities in the decision process. Behav Brain Sci. 31, 415-437.

Sim, M.E., Lyoo, I.K., Streeter, C.C., Covell, J., Sarid-Segal, O., Ciraulo, D.A., et al., 2007.

Cerebellar gray matter volume correlates with duration of cocaine use in cocaine-

dependent subjects. Neuropsychopharmacology 32, 2229-2237.

Smith, G.T., Fischer, S., Cyders, M.A., Annus, A.M., Spillane, N.S., 2007. On the validity and

utility of discriminating among impulsivity-like traits. Assessment 14, 155-170.

Substance Abuse and Mental Health Services Administration, 2010. Results from the 2009

National Survey on Drug Use and Health: Mental Health Findings (Office of Applied

Studies, NSDUH Series H-39, HHS Publication No. SMA 10-4609). Rockville, MD.

Tanabe, J., Tregellas, J.R., Dalwani, M., Thompson, L., Owens, E., Crowley, T., et al., 2009.

Medial orbitofrontal cortex gray matter is reduced in abstinent substance-dependent

individuals. Biol Psychiatry 65, 160-164.

Verdejo-García, A., Bechara, A., 2009. A somatic marker theory of addiction.

Neuropharmacology 56, 48-62.

Verdejo-García, A., Lawrence, A.J., Clark, L., 2008. Impulsivity as a vulnerability marker for

substance-use disorders: review of findings from high-risk research, problem gamblers

and genetic association studies. Neurosci Biobehav Rev. 32, 777-810.

Verdejo-García, A., Lozano, O., Moya, M., Alcázar, M.A., Pérez-García, M., 2010.

Psychometric properties of a Spanish version of the UPPS-P impulsive behavior scale:

P á g i n a | 6 6

reliability, validity and association with trait and cognitive impulsivity. J Pers Assess. 92,

1532-7752.

Verdejo-García, A.J., López-Torrecillas, F., Aguilar de Arcos, F., Pérez-García, M., 2005.

Differential effects of MDMA, cocaine, and cannabis use severity on distinctive components

of the executive functions in polysubstance users: a multiple regression analysis. Addict

Behav. 30, 89-101.

Volkow, N.D., Wang, G.J., Telang, F., Fowler, J.S., Logan, J., Childress, A.R. et al., 2006.

Cocaine cues and dopamine in dorsal striatum: mechanism of craving in cocaine addiction.

J Neurosci. 26, 6583-6588.

Volkow, N.D., Chang, L., Wang, G.J., Fowler, J.S., Ding, Y.S., Sedler, M., et al., 2001. Low level

of brain dopamine D2 receptors in methamphetamine abusers: association with

metabolism in the orbitofrontal cortex. Am J Psychiatry 158, 2015-2021.

Volz, K.G., Schubotz, R.I., von Cramon, D.Y. 2003. Predicting events of varying probability:

uncertainty investigated by fMRI. Neuroimage 19, 271-280.

Whiteside, S.P., Lynam, D.R., 2001. The five factor model and impulsivity: using a structural

model of personality to understand impulsivity. Pers Individ Dif. 30, 669-689.

P á g i n a | 6 7

* ARTÍCULO 2

Neural Correlates of the Severity of Cocaine, Heroin, Alcohol, MDMA and Cannabis

Use in Polysubstance Abusers: A Resting-PET Brain Metabolism Study

* Moreno-López, L., Stamatakis, E. A., Fernández-Serrano, M. J., Gómez-Río, M., Rodríguez-

Fernández, A., Pérez-García, M., & Verdejo-García, A. (2012). Neural Correlates of the

Severity of Cocaine, Heroin, Alcohol, MDMA and Cannabis Use in Polysubstance Abusers: A

Resting-PET Brain Metabolism Study. PLoS ONE, en prensa.

P á g i n a | 6 8

1. Introduction

Drug addiction has been associated with neuroadaptations in brain systems

involved in motivation, memory and executive control [1]. Functional neuroimaging

studies have revealed that the use of different classes of drugs is associated with

dysfunctions in a range of overlapping brain regions including ventromedial and

dorsolateral prefrontal cortex (DLPFC), anterior cingulate cortex, inferior frontal gyrus,

insula, amygdala, basal ganglia and cerebellum [2]. These dysfunctions may explain the

overlapping cognitive deficits observed in users of different drugs, including working

memory, inhibitory control, flexibility or decision-making deficits [3]. In addition, some

specific effects have been described for the association between cocaine use and the

parietal cortex [4,5,6], heroin use and the temporal cortex [7], MDMA use and occipital,

hippocampus and thalamic regions [8,9], and cannabis use and the premotor cortex

[10,11]. These associations are congruent with relatively specific neuropsychological

deficits pertaining to attention and cognitive control in cocaine users, long-term memory

in opiate users, visuospatial memory in MDMA users, and psychomotor function in

cannabis users [3,7]. In spite of the available evidence, in vivo studies have not been

systematically conducted to look at associations between severity of use of different drugs

and brain dysfunction. In fact, several neuroimaging studies have failed to detect such link

or have provided controversial results (reviews in 12). These negative findings and

controversies might be linked to the impact of relevant confounding variables. One of

these is the fact that neuroimaging studies have frequently tested drug users currently

using drugs or having brief periods of abstinence (24-48 h) [13, 14]; under these

conditions, the presumed link between lifetime drug use and brain dysfunction could be

masked by several other factors, including recent drug use or psychological symptoms

associated with withdrawal and short-term abstinence [2]. Another key variable for

consideration is the concurrent use of multiple types of substances, which can introduce

significant confounds in the interpretation of the data.

P á g i n a | 6 9

The current study aimed to specifically address the association between lifetime

estimates of use of different types of drugs and brain functioning (as measured by 18F-

fluorodeoxyglucose Positron Emission Tomography (FDG-PET), statistically controlling for

the concurrent use of other drugs in a sample of polysubstance users with prolonged

abstinence from all drugs. Specifically, we examined the link between lifetime estimates of

amount and duration of use of cocaine, heroin, alcohol, MDMA and cannabis and brain

metabolism (BM). Based on previous findings we hypothesized that exposure to all

substances would be significantly associated with overlapping reductions of metabolism

in frontal, limbic, striatal and cerebellar regions, whereas cocaine and heroin use would

specifically correlate with parietal and temporal cortices respectively.

2. Methods

2.1 Participants

Forty-nine substance dependent individuals (SDI), forty-one men and eight women

with a mean age of 32.67 years, were recruited during their stay in an inpatient treatment

program at the centre “Proyecto Hombre” in Granada (Spain). This centre provides a

controlled setting for the treatment of substance use disorders. Selection criteria for

participants in this study were (i) meeting the DSM-IV criteria for substance dependence;

(ii) absence of documented comorbid mood or personality disorders, as assessed by

clinical reports; (iii) absence of documented head injury or neurological disorders; and

(iv) to have minimum abstinence duration from all drugs consumed (except for tobacco)

of 15 days before testing. In fact, SDI participants had overall much longer abstinence

duration (mean of 32.94, SD=11.25 weeks); for that reason it was possible to rule out the

presence of withdrawal symptoms or brain function alterations associated with the acute

or short term effects of the drugs. None of the participants were experiencing withdrawal

symptoms as assessed by routine medical examination or were enrolled in opioid

P á g i n a | 7 0

substitution treatments with methadone or other pharmacological treatments (e.g.,

benzodiazepines) during the course of the assessment interview/PET evaluations.

2.2 Testing protocols and procedures

This study was approved by the Human Subjects Committee of the University of

Granada. One-step Syva urine drug screens for amphetamines, benzodiazepines, cannabis,

cocaine and opiates were conducted to confirm abstinence at time of testing. Participants

were evaluated individually in two different sessions. During the first session, the

participants were provided with information about the study and were encouraged to ask

questions to ascertain they fully understood the rational for the study and were clear on

the details for participation. The participants then provided written consent and were

assessed with the Interview for Research on Addictive Behaviours [15] in the “Proyecto

Hombre” installations”. During the second session, each patient was accompanied by a

therapist to the nuclear medicine service of the “Virgen de las Nieves” hospital, where the

PET neuroimaging session took place. This second session usually took place one week

after the first session was completed.

2.3 Tools

2.3.1 Drug Use Information

The Interview for Research on Addictive Behaviour [15] was used to examine the

severity of drug use. This interview evaluates, by means of a brief interview, the quantity

(average dosing), frequency (number of drug taking episodes per month), and duration

(years of duration) of the use of different substances that can produce physical or

psychological dependence, including cocaine, heroin, alcohol, MDMA and cannabis which

were the main drugs of choice in the present study. For every substance the participant

had actually used, the following information was requested: (i) the average dose of each

target drug taken in each episode of use (number of grams for cocaine and heroin, number

P á g i n a | 7 1

of units for alcohol, considering that 25 cl. of hard liquor (e.g., scotch) equals 2 units, while

25 cl. of wine or beer equals 1 unit, number of pills for MDMA and number of cigarettes for

cannabis); (ii) the frequency of these consumption episodes per month (daily, between

one and three times upon a week, once a week, between one and three times upon a

month, or once a month); and (iii) the number of years that elapsed since the onset of use.

From this information, we obtained two indices for each of the drugs used: (i) Amount

(average dose x frequency), an index of monthly use of the drug; and (ii) Duration of drug

use, measuring number of years of exposure to the drug. The main sociodemographic and

clinical features from the sample are displayed in Table 1.

Table 1 Sociodemographic data and clinical features of the participants

Subjects Characteristics Gender Male 41

Female 8

Mean S.D. Maximum Minimum

Age (years) 32.67 8.03 53 21

Education (years) 9.73 2.34 17 5

Amount of drug usea Substance Mean S.D. Maximum Minimum Cocaine (g) (46) 48.69 44.83 180 0

Heroin (g) (16) 9.12 20.63 120 0

Alcohol (units) (45) 571.24 489.34 1800 0

MDMA (pills) (23) 13.41 24.31 120 0

Cannabis (cigarettes) (37) 148.57 190.64 750 0

Duration of drug use (years) Substance Mean S.D. Maximum Minimum Cocaine 7.95 5.95 23 0

Heroin 1.76 3.69 17 0

Alcohol 10.85 7.67 27 0

MDMA 1.40 2.28 8 0

Cannabis 8.31 8.13 29 0

Abstinence (weeks) 32.94 11.25 60 12 a Amount of monthly use of the drug calculated using the average dose (per episode) × frequency (per month).

P á g i n a | 7 2

2.3.2 PET Image Acquisition

PET scans were acquired with an ECAT/931 scanner (Siemens CTI ECAT EXACT),

at the service of nuclear medicine centre of the hospital “Virgen de las Nieves” in Granada

(Spain). For each individual, we obtained 20 minute emission scans 30 minutes after the

injection of one dose of 200 MBq of FDG administrated only after levels of glycemia had

been checked (they must be below 120 mg/dl) [16]. The subjects were scanned with their

eyes open and ears unplugged in a dimly illuminated quiet room. Raw data were

processed using an iterative reconstruction technique (OSEM method: 10 iterations, 32

subsets). Images were reoriented in transaxial, coronal and saggital planes. For the

analyses reported here the reconstructed images used had 47 axial slices each with an in-

plane resolution of 2.57x2.57 and a slice thickness of 3.38mm.

2.4 Data analysis

2.4.1 Preprocessing of PET images

PET images were converted from DICOM to NIFTI format and were spatially

normalized to the SPM PET template (Wellcome Department of Cognitive Neurology,

London, UK) using linear (12 affine) and non-linear normalization. An average image of all

the normalized images was obtained and this acted as a study specific PET template. In

this manner we constructed a template in the same modality, from the same scanner as

the images we wanted to normalize. The raw PET images were then spatially normalized

with linear affine and nonlinear parameters to this study-specific template. Visual

inspection revealed this optimized spatial normalization technique produced satisfactory

spatial normalization for the PET image of every volunteer.

2.4.2 Statistical Analysis

P á g i n a | 7 3

The spatially normalized PET images were smoothed with an 8mm3 isotropic

Gaussian filter and were statistically modeled using the General Linear Model in SPM5.

Linear regressions were used to carry out several analyses in order to identify brain areas

with a significant correlation between uptake values and measures of the amount and

duration of cocaine, heroin, alcohol, MDMA and cannabis use. Not all subjects used all the

drugs but since virtually all participants had regularly abused two or more substances, the

analyses were conducted on the whole group and results were interpreted in relation to

polysubstance use. In each of the analysis, we controlled for the effects of the other

substances concurrently used (e.g., analysis of severity of cocaine use controlled for the

effects of co-abuse of heroin, alcohol, MDMA and cannabis), age and years of education.

We report local maximum peaks that survive a voxel threshold of p<0.001 (uncorrected

for multiple comparisons, cluster size ≥100 voxels). We chose to report our findings at this

threshold based on numerous similar studies of PET imaging in drug addiction [17,18].

3. Results

The main results are summarized in Table 2.

3.1 Amount of use

Voxel-based analyses showed negative correlations between lifetime amount of

cocaine, alcohol and cannabis use and several brain regions. The amount of cocaine used

was negatively correlated with a cluster encompassing the left inferior parietal lobule

extending to the left postcentral gyrus (Figure 1.1). The amount of alcohol used was

negatively associated with three clusters: one cluster included the left middle and superior

temporal cortex; a second cluster included the bilateral superior frontal cortex extending

to the left DLPFC and the right supplementary motor area; and a third cluster included the

right DLPFC extending to the superior frontal gyrus (Figure 1.2). The amount of cannabis

use showed a negative correlation with two clusters: one cluster included the left inferior

frontal gyrus (pars triangularis) extending to the DLPFC, and a second cluster included the

P á g i n a | 7 4

right DLPFC extending to the superior frontal cortex (Figure 1.3). We did not find

significant correlations between the amount of heroin or MDMA use and BM measures.

3.2 Duration of use

Voxel-based analyses showed negative correlations between duration of heroin,

alcohol and MDMA use and several brain regions. Duration of heroin use showed a

negative correlation with three clusters, which encompassed the right inferior, middle and

superior temporal cortex extending to the right supramarginal and inferior frontal gyrus

(Figure 1.4). Duration of alcohol use was negatively correlated with four clusters. One

cluster included the right inferior temporal cortex extending to the middle aspect of this

area. A second cluster included the right putamen and pallidum. A third cluster

encompassed the right precentral and postcentral gyrus and the fourth cluster included

the right fusiform and parahippocampal gyrus (Figure 1.5). Finally, we found negative

correlations between the duration of MDMA use and BM in three clusters: the first cluster

included the left postcentral and inferior parietal gyrus; the second cluster included the

right inferior frontal gyrus (pars triangularis) extending to the DLFPC; and the third

cluster included the right superior temporal pole extending to the middle part of this

region (Figure 1.6).

P á g i n a | 7 5

Table 2 Relation between severity of drug use and PET uptake.

Substance Voxels-Based

Analysis

BA KE T MNI

Amount

Cocaine Parietal Inf L 40 176 3.98 -58 -38 54

Alcohol Temporal Mid L 21 267 4.04 -66 -2 -10

Frontal Sup L 9 308 3.96 -20 30 36

Frontal Sup R 6 365 3.73 22 2 60

Frontal Mid R 46 372 3.82 26 22 34

Cannabis Frontal Inf Tri L 45 355 4.40 -44 22 30

Frontal Mid R 46 125 4.02 24 48 28

Duration

Heroin Frontal Mid R 46 876 3.77 44 36 34

Temporal Inf R 20 257 4.58 62 -16 -32

Temporal Sup R 22 1090 4.19 66 -44 20

Alcohol Temporal Inf R 20 238 4.22 64 -16 -30

Putamen R 472 3.71 20 -2 8

Precentral R 6 274 3.67 50 2 34

Fusiform R 20 299 3.65 30 -6 -40

MDMA Postcentral L 40 131 3.73 -28 -40 48

Frontal Inf Tri R 45 139 3.67 44 34 20

Temporal Pole Sup R 20 561 3.67 32 12 -30 Voxel-based analyses at p<0.001 (uncorrected, cluster size > 100 voxels). Labels were obtained using the aal toolbox (Tzourio-Mazoyer et al., 2002). Broadmann areas were obtained using MRICron (Rorden and Brett, 2000). KE indicate the number of voxels included in the cluster. Stereotaxic coordinades are those listed in SPM5. Inf, inferior; Sup, superior; L, left; R, right; Supp, supplementary; Mid, middle; Orb, orbitalis.

P á g i n a | 7 6

Figure 1. Areas of reduced metabolism predicted by the amount or duration of drug

use: (1) amount of cocaine, (2) amount of alcohol, (3) amount of cannabis (4) duration of

heroin (5) duration of alcohol and (6) duration of MDMA. Areas of reduced metabolism are

superimposed on a T1 weighted MRI image in MRICRON. MNI coordinates are shown

underneath each panel.

P á g i n a | 7 7

4. Discussion

Our findings showed that the estimates of severity of use of heroin, alcohol, MDMA

and cannabis were negatively associated with DLPFC and temporal cortex regional

metabolism in polysubstance dependent individuals with prolonged abstinence from all

these drugs. Alcohol use was further associated with lower metabolism in frontal

premotor cortex and putamen. Furthermore, severity of stimulants use (cocaine and

MDMA) was uniquely associated with inferior parietal/postcentral cortex metabolism.

These associations were established after a prolonged period of abstinence, and after

controlling for multiple confounding variables, thus surpassing the limitations of previous

functional neuroimaging studies that were not specifically designed to study prolonged

abstinence, and were confounded by important variables, such as withdrawal related

symptoms or concurrent use of multiple substances. The associations found are in

agreement with our initial predictions, and with the neuropsychological deficits typically

aligned with the chronic use of these substances; these include reduced competency in

executive functions (related to all classes of drugs), episodic memory (mainly observed on

heroin, alcohol, MDMA and cannabis users), motor control (alcohol and cannabis users),

and visuospatial attention (stimulants users) (see meta-analyses in 3,8,19,20).

As expected, the estimates of amount or duration of different classes of drugs were

negatively associated with overlapping regions within the DLPFC (BA 9, 45, 46). This

region has been linked with several of the key neuroadaptations associated with drug

addiction, including drug conditioning, loss of self-control, and stimulus-driven

compulsive behavior [2]. This finding is also in agreement with results from several

neuropsychological studies showing cognitive deficits on working memory, planning or

inhibition processes, which are associated with DLPFC functioning, across users of several

drugs [3]. We also found negative associations between estimates of duration of heroin,

alcohol, and MDMA use and regional metabolism on partially overlapping sections of the

P á g i n a | 7 8

temporal cortex. These associates may substantiate the links between the intensity and

duration of involvement with these drugs and the degree of memory attrition, which has

been observed in longitudinal studies with users of these substances [21,22]. It is worth

mentioning that duration of heroin use was shorter to that of cocaine and alcohol use;

however, previous brain structural findings indicate that duration of heroin use is a

critical factor leading to brain damage, even in users with periods of use below five years

[23]. Surprisingly, we failed to find significant associations between the severity of the

substances used and overlapping limbic, striatal and cerebellar regions but this result

could be explained from the fact that these regions play a greater role during current use

and short-term abstinence [24]. In addition, we found relatively specific negative

correlations between amount of cocaine use and regional metabolism in the inferior

parietal cortex, and between alcohol duration and regions involved in motor programming

and control (frontal precentral and putamen). The link between cocaine use and parietal

dysfunction has been observed on previous functional imaging studies during cognitive

challenges of attention and executive control [6,25,26]. The link between alcohol use and

precentral and basal ganglia dysfunction is in agreement with structural imaging findings

showing that function of these regions is significantly reduced in alcohol users compared

to other forms of addiction [27]. Additional evidence comes from volumetric studies which

associated volumetric measures of these regions and deficient motor skills in alcoholics

[28]. These more specific findings could be used to design tailored interventions aimed to

address specific aspects of frontal-subcortical executive functions in different profiles of

substance abusers.

An important implication of this study is the fact that the link between severity of

drug use and BM is still observable in substance abusers with prolonged periods of

abstinence (an average of eight months in this sample). Therefore, quantity and duration

of drug use may still be affect brain functioning after several months of successful

cessation of drug use. This is particularly relevant in view of recent findings indicating that

P á g i n a | 7 9

certain patterns of brain dysfunction can reliably predict mid-term and long-term relapse.

Permanent brain alterations may also have important repercussions for the rehabilitation

of SDI, since these alterations can affect the ability of SDI to assimilate the contents and

participate in the activities of the rehabilitation programs [29,30]. In fact, preliminary

evidence suggests that cognitive rehabilitation techniques, such as errorless learning may

be successful in compensating chronic cognitive deficits in substance abusers [31]. Thus,

the brain networks impacted by severity of drug abuse can play a central role in clinical

outcome and relapse and should become priority targets for therapeutic interventions.

Several possible limitations of our study should be considered and addressed by

future research. Our findings were not obtained in “pure” users of each of the substances

but in mostly polysubstance users. Although this limitation is inherent to the clinical

literature on addiction (i.e., “pure” users of one single drug are quite rare among substance

abusers enrolled in treatment programs), we attempted to control for the co-abuse of

other drugs using statistics. A related confounding variable was the use of nicotine, which

is ubiquitous among drug users but was not specifically assessed or controlled for.

Furthermore, due to the use of a correlational design, this study cannot yield conclusions

about cause-effect relationships between the use of drug and resting BM. However, given

the agreement of our results with both resting state PET, activation PET/fMRI and

addiction neuropsychology studies, we could hypothesize that, at least in part, the

alterations observed in this population could be due to the quantity and duration related

effects of the consumption of the substances under consideration in this study.

One last point to consider is that our findings could be due to premorbid brain

alterations or the results of the interaction between the premorbid alterations and the

neurotoxic effects of drug use. This question should be addressed through longitudinal

designs. A related limitation is that no comparison with a healthy control database was

performed. As we argued before, a healthy control group was not needed to test the main

P á g i n a | 8 0

predictions of our study; however, the lack of controls has stopped us from making any

assumption about the direction of these findings in relation to the healthy population.

P á g i n a | 8 1

References

1. Koob GF, Volkow ND (2010) Neurocircuitry of addiction. Neuropsychopharmacology 35:

217-238.

2. Goldstein RZ, Volkow ND (2011) Dysfunction of the prefrontal cortex in addiction:

neuroimaging findings and clinical implications. Nat Rev Neurosci 12: 652-669.

3. Fernández-Serrano MJ, Pérez-García M, Verdejo-García A (2011) What are the specific

vs. generalized effects of drugs of abuse on neuropsychological performance? Neurosci

Biobehav Rev 35: 377-406.

4. Aron JL, Paulus MP (2007) Location, location: using functional magnetic resonance

imaging to pinpoint brain differences relevant to stimulant use. Addiction 1: 33-43.

5. Lane SD, Steinberg JL, Ma L, Hasan KM, Kramer LA, et al. (2010) Diffusion tensor

imaging and decision making in cocaine dependence. PLoS One 5: e11591.

6. Tomasi D, Goldstein RZ, Telang F, Maloney T, Alia-Klein N, et al. (2007) Widespread

disruption in brain activation patterns to a working memory task during cocaine

abstinence. Brain Res 1171: 83-92.

7. Gruber SA, Silveri MM, Yurgelun-Todd DA (2007) Neuropsychological consequences of

opiate use. Neuropsychol Rev 17: 299-315.

8. Kish SJ, Lerch J, Furukawa Y, Tong J, McCluskey T, et al. (2010) Decreased cerebral

cortical serotonin transporter binding in ecstasy users: a positron emission

tomography/[(11)C]DASB and structural brain imaging study. Brain 133: 1779-1797.

9. de Win MM, Jager G, Booij J, Reneman L, Schilt T, et al. (2008) Neurotoxic effects of

ecstasy on the thalamus. Br J Psychiatry 193: 289-296.

P á g i n a | 8 2

10. King GR, Ernst T, Deng W, Stenger A, Gonzales RM, et al. (2011) Altered brain

activation during visuomotor integration in chronic active cannabis users: relationship to

cortisol levels. J Neurosci 31: 17923-17931.

11. Pillay SS, Rogowska J, Kanayama G, Gruber S, Simpson N, et al. (2008) Cannabis and

motor function: fMRI changes following 28 days of discontinuation. Exp Clin

Psychopharmacol 16: 22-32.

12. Martín-Santos R, Fagundo AB, Crippa JA, Atakan Z, Bhattacharyya S, et al. (2010)

Neuroimaging in cannabis use: a systematic review of the literature. Psychol Med 40: 383-

98.

13. Goldstein RZ, Alia-Klein N, Tomasi D, Carrillo JH, Maloney T, et al. (2009) Anterior

cingulate cortex hypoactivations to an emotionally salient task in cocaine addiction. Proc

Natl Acad Sci U S A 106: 9453-9458.

14. Zubieta JK, Heitzeg MM, Xu Y, Koeppe RA, Ni L, et al. (2005) Regional cerebral blood

flow responses to smoking in tobacco smokers after overnight abstinence. Am J Psychiatry

162: 567-577.

15. Verdejo-García AJ, López-Torrecillas F, Aguilar de Arcos F, Pérez-García M (2005)

Differential effects of MDMA, cocaine, and cannabis use severity on distinctive components

of the executive functions in polysubstance users: a multiple regression analysis. Addict

Behav 30: 89-101.

16. Boellaard R, O’Doherty MJ, Weber WA, Mottaghy FM, Lonsdale MN, et al. (2010) FDG

PET and PET/CT: EANM procedure guidelines for tumour PET imaging: version 1.0. Eur J

Nucl Med Mol Imaging 37: 181-200.

P á g i n a | 8 3

17. Fehr C, Hohmann N, Gründer G, Dielentheis TF, Buchholz HG, et al. (2007) Tiagabine

does not attenuate alcohol-induced activation of the human reward system.

Psychopharmacology (Berl) 191: 975-983.

18. Volkow ND, Fowler JS, Wang GJ, Telang F, Logan J, et al. (2010) Cognitive control of

drug craving inhibits brain reward regions in cocaine abusers. Neuroimage 49: 2536-

2543.

19. Jovanovski D, Erb S, Zakzanis KK (2005) Neurocognitive deficits in cocaine users: a

quantitative review of the evidence. J Clin Exp Neuropsychol 27: 189-204.

20. Zakzanis KK, Campbell Z, Jovanovski D (2007) The neuropsychology of ecstasy

(MDMA) use: a quantitative review. Hum Psychopharmacol 22: 427-435.

21. Hanson KL, Cummins K, Tapert SF, Brown SA (2011) Changes in neuropsychological

functioning over 10 years following adolescent substance abuse treatment. Psychol Addict

Behav 25: 127-142

22. de Sola S, Tarancón T, Peña-Casanova J, Espadaler JM, Langohr K, et al. (2008) Auditory

event-related potentials (P3) and cognitive performance in recreational ecstasy polydrug

users: evidence from a 12-month longitudinal study. Psychopharmacology (Berl) 200:

425-437

23. Yuan Y, Zhu Z, Shi J, Zou Z, Yuan F, et al. (2009) Gray matter density negatively

correlates with duration of heroin use in young lifetime heroin-dependent individuals.

Brain Cogn 71: 223-228.

24. Breiter HC, Rosen BR (1999) Functional magnetic resonance imaging of brain reward

circuitry in the human. Ann N Y Acad Sci 877: 523-547.

P á g i n a | 8 4

25. Barrós-Loscertales A, Bustamante JC, Ventura-Campos N, Llopis JJ, Parcet MA, et al.

(2011) Lower activation in the right frontoparietal network during a counting Stroop task

in a cocaine-dependent group. Psychiatry Res 194: 111-118.

26. Bustamante JC, Barrós-Loscertales A, Ventura-Campos N, Sanjuán A, Llopis JJ, et al.

(2011) Right parietal hypoactivation in a cocaine-dependent group during a verbal

working memory task. Brain Res 1375: 111-119.

27. van Holst RJ, de Ruiter MB, van den Brink W, Veltman DJ, Goudriaan AE (2012) A

voxel-based morphometry study comparing problem gamblers, alcohol abusers, and

healthy controls. Drug Alcohol Depend. In press.

28. Sullivan EV, Deshmukh A, De Rosa E, Rosenbloom MJ, Pfefferbaum A (2005) Striatal

and forebrain nuclei volumes: contribution to motor function and working memory

deficits in alcoholism. Biol Psychiatry 57: 768-776.

29. Johnson-Greene D, Adams KM, Gilman S, Junck L (2002) Relationship between

neuropsychological and emotional functioning in severe chronic alcoholism. Clin

Neuropsychol 16: 300-309.

30. Zinn S, Stein R, Swartzwelder HS (2004) Executive functioning early in abstinence from

alcohol. Alcohol Clin Exp Res 28: 1338-1346.

31. Pitel AL, Perruchet P, Vabret F, Desgranges B, Eustache F, et al. (2009) The advantage

of errorless learning for the acquisition of new concepts' labels in alcoholics. Psychol Med

40: 497-502.

Tzourio-Mazoyer N, Landeau B, Papathanassiou D, Crivello F, Etard O, et al. (2002)

Automated anatomical labeling of activations in SPM using a macroscopic anatomical

parcellation of the MNI MRI single-subject brain. Neuroimage 15: 273-289.

P á g i n a | 8 5

* ARTÍCULO 3

Neural correlates of hot and cold executive functions in polysubstance addiction:

Association between neuropsychological performance and resting-PET brain

metabolism

* Moreno-López, L., Catena, A., Fernández-Serrano, M. J., Delgado-Rico, E., Stamatakis, E. A.,

Pérez-García, M., & Verdejo-García, A. (2012). Neural correlates of hot and cold executive

functions in polysubstance addiction: Association between neuropsychological

performance and resting-PET brain metabolism. Psychiatry Research: Neuroimaging, en

prensa.

P á g i n a | 8 6

1. Introduction

Contemporary neurobiological models consider addiction to be a brain disorder

that involves long-term neuroadaptations leading to persistent drug use despite

increasing negative consequences (Baler and Volkow, 2006; Goldstein and Volkow, 2002;

Verdejo-García et al., 2004). These neuroadaptations are thought to affect neural systems

involved in motivation, emotion, learning, memory and executive functioning (Ersche et

al., 2008; Verdejo-García and Bechara, 2009). From this perspective, it is crucial to

understand both the long-term neuropsychological consequences of drug abuse and their

neural substrates.

There is compelling evidence that substance dependent individuals (SDI) present

widespread deficits in neuropsychological functioning, and that these deficits are

especially prominent in so-called executive functions (Fernández-Serrano et al., 2011).

Executive functions are defined as a set of higher-order abilities involved in the

employment, monitoring, and regulation of goal-directed behaviors. Recent evidence

suggests that these functions can be classified into two broad domains: (i) cold executive

functions, involved in the processing of relatively abstract, context-free information, which

are linked to more dorsal and lateral regions of the prefrontal cortex, and (ii) hot executive

functions, which come to play when emotionally laden information is concerned, and are

linked to the functioning of the orbitofrontal cortex (OFC) (Kerr and Zelazo, 2004).

Although executive dysfunction in SDI has been extensively described, the study of

the neurobiological substrates of executive dysfunction in the context of drug addiction is

still a growing research area (Goldstein and Volkow, 2002; Lundqvist, 2010). A number of

functional magnetic resonance imaging (fMRI) studies have shown significant associations

between performance on cognitive tasks indexing cold aspects of executive functions and

abnormal activation of brain systems in SDI. Polysubstance abusers have shown

abnormally increased or decreased activation of the dorsolateral prefrontal cortex

P á g i n a | 8 7

(DLPFC), orbitofrontal cortex (OFC), temporo-parietal regions and the cerebellum during

performance on working memory (Desmond et al., 2003; Jager et al., 2006; Tomasi et al.,

2007), response inhibition (Bolla et al., 2004; Dao-Castellana et al., 1998; Eldreth et al.,

2004; Forman et al., 2004; Lee et al., 2005), and cognitive flexibility tasks (Gilman et al.,

1990; Goldstein et al., 2004). Nevertheless, very few studies have investigated the neural

correlates of indices of hot executive functions in SDI. These abilities have been associated

with the OFC and the cingulate cortex in cognitive neuroscience studies (Anderson et al.,

2006; Hartstra et al., 2010; Heberlein et al., 2008), and they have been found to be

dysfunctional in polysubstance abusers even after mid-term abstinence (circa six months)

(Fernández-Serrano et al., 2011). For example, two neuropsychological studies have

revealed deficits in goal-driven self-regulation in psychostimulant users using the Revised

Strategy Application Test (R-SAT) (Halpern et al., 2004; Verdejo-García et al., 2007), but

no studies to date have investigated the neural substrates of self-regulation in addiction.

Another key factor for addiction initiation and maintenance is adaptive decision-making,

which is typically indexed using the Iowa Gambling Task (IGT) (Bechara et al., 1994).

Performance on this task has been associated with hyperperfusion in the anterior

cingulate cortex (ACC), middle and superior frontal gyrus and DLPFC in small samples of

shortly abstinent cocaine-dependent individuals (Adinoff et al., 2003; Tucker et al., 2004).

However, because of the key role of decision-making in mid and long-term drug relapse

(Passetti et al., 2008; Paulus et al., 2005), there is a need to extend these findings in larger

samples of mid-term abstinent SDI. Similarly, the neural correlates of emotional

processing and perception must be considered in light of their association with hot aspects

of executive functions in SDI (Verdejo-García et al., 2007), and their predictive role in drug

relapse (Lubman et al., 2009). Only two previous neuroimaging studies have focused on

this domain, and they have found decreased ACC and amygdala activations in alcohol users

exposed to facial emotional expressions (Marinkovic et al., 2009; Salloum et al., 2007).

P á g i n a | 8 8

In this study we aimed to investigate the neuropsychological performance of a

sample of SDI and healthy controls (HC) on tests tapping on cold (i.e., updating, inhibition

and flexibility) and hot (i.e., self-regulation, decision-making and emotion perception)

aspects of executive functions, and the association between the neuropsychological

performance of SDI on those tests and PET-indexed regional metabolism. We hypothesize

that neuropsychological performance will be poorer in SDI than HC and that it will be

associated with DLPFC, ACC and temporal regions for tests of cold executive functions, and

with OFC, ACC and limbic regions for tests of hot executive function.

2. Methods

2.1. Participants

Forty-nine SDI were recruited from an inpatient treatment program at the centre

“Proyecto Hombre” in Granada (Spain). The treatment has a mean duration of 6 months

and integrates key elements of cognitive behavioral psychology and systemic approaches

to provide complete social rehabilitation and integration. Selection criteria for participants

in this study were (1) meeting the DSM-IV criteria for substance dependence at the onset

of the treatment; (2) absence of documented comorbid mood or personality disorders, as

assessed by clinical reports; (3) absence of documented head injury or neurological

disorders, as assessed by clinical reports and PET images; (4) to be literate enough to

perform any literacy tests, as assessed by the TAP reading skills test (Del Ser et al., 1997)

and (5) to have a minimum abstinence duration of 15 days before testing. In fact, the SDI

included in the sample had overall much longer abstinence duration (M=32.94, SD=11.25

weeks), allowing us to rule out acute or short term effects of drug intake. Since virtually all

participants had regularly abused two or more substances, the analyses were conducted

on the whole group and results were interpreted in relation to polysubstance use. Urine

toxicology screening (one-step Syva rapid tests) for the different drugs used were

conducted routinely in the SDI, so we can rule out the use of these substances during the

P á g i n a | 8 9

entire period of abstinence. None of the participants were experiencing withdrawal

symptoms before or during neuropsychological/PET testing as assessed by routine

medical examination. Likewise, none of the participants were enrolled in opioid

substitution treatments with methadone or other pharmacological treatments during the

course of the neuropsychological/PET testing. A group of drug-free healthy individuals

was recruited as a comparison group for neuropsychological performance, but these

participants were not PET-scanned. HC were recruited through local advertisements and

snowball communication among adult people from the community. Selection criteria for

these group were: (1) absence of current or past substance abuse, excluding past cannabis

use and past or current smoking or social drinking (in cases of current social drinking,

exclusion criteria included acute use of alcohol during the last 24 h before evaluation), (2)

absence of documented major psychiatric disorders, (3) absence of documented head

injury or neurological disorder and (4) not being on any medication that could affect the

normal neuropsychological functioning. This group only had exposure to alcohol use,

having a regular mean use of 8.85 standard units per week (SD=20.7) and a mean duration

of use of 6.7 years (SD=7.3). The main sociodemographic and clinical features from the

samples are displayed in Table 1.

Table 1 Sociodemographic data and clinical features of the participants.

SDI (n=49) HC (n=30)

Variable N % N %

Gender Male 41 84 24 80 Female 8 16 6 20

Mean SD Mean SD

Age (years) 32.67* 8.03 26.4* 8.03 Education (years) 9.73* 2.34 11.63* 2.04

* p ≤ 0.001

P á g i n a | 9 0

2.2. Testing protocols and procedures

This study was approved by the Human Subjects Committee of the University of

Granada. All participants signed an informed consent form before participating in the

study. The testing protocol, including the neuropsychological assessment and the PET

scanning were administered into two sessions that lasted approximately 70 minutes each.

The control group only participated in the first session. During the first session,

participants completed the Interview for Research on Addictive Behaviors and performed

the neuropsychological tests protocol. During the second session (up to 7 days later), SDI

were scanned at the nuclear medicine service of the “Virgen de las Nieves” hospital.

2.3. Instruments

2.3.1. Patterns of drug use

The information about the severity of drug addiction was collected by the

Interview for Research on Addictive Behavior (IRAB; (López-Torrecillas et al., 2001) (see

supplementary material). This is a brief interview that collects information about the

quantity (average dose), frequency (consumption episodes by month), and duration (years

of duration) of the use of a series of drugs of abuse, including cocaine, heroin, and alcohol.

The main drug use information of the SDI group is displayed in Table 2.

P á g i n a | 9 1

Table 2 Amount and duration of drug use in the substance dependence individuals.

Quantity of use

Substance Averagea S.D. Maximum Minimum

Cocaine (46) 48.69 44.83 180 0 Cannabis (37) 148.57 190.64 750 0 Heroin (16) 9.12 20.63 120 0 MDMA (18) 13.41 24.31 120 0 Alcohol (45) 571.24 489.34 1800 0

Duration of use (years)

Substance Duration S.D. Maximum Minimum Age of onset

Cocaine 7.95 5.95 23 0 23.26 Cannabis 8.31 8.13 29 0 20.91 Heroin 1.76 3.69 17 0 24.75 MDMA 1.40 2.28 8 0 23.66 Alcohol 10.85 7.67 27 0 20.87 Abstinenceb 32.94 11.25 60 12

a Average dose (per episode) × Frequency (per month). A dose is units in alcohol, grams in cocaine and heroin, fags in cannabis and pills in MDMA. b Abstinence in weeks. The number in brackets is the number of substance dependence individuals that consumed the substance.

2.3.2. Neuropsychological tests

We administered a battery of tests designed to assess several domains related to

cold and hot executive functions, including working memory, response inhibition,

cognitive flexibility, self-regulation, decision-making and emotion processing.

The tests used were Letter Number Sequencing (LNS) (Wechsler Adult Intelligence

Scale - WAIS-III, Wechsler, 1997), Stroop (Golden, 1978), Category Test (DeFilippis, 2002),

Revised Strategy Application Test (R-SAT) (Levine et al., 2000), Iowa Gambling Task (IGT)

(Bechara et al., 1994) and Ekman Faces Test (EFT) (Young et al., 2002). All tests were

administered in a fixed order and according to standard instructions. The description of

the instruments is provided in the supplementary material.

Due to technical problems during data acquisition, we could not obtain the

performance indices for some of the tests in the SDI group. Therefore, the number of SDI

P á g i n a | 9 2

participants included in the analyses of LNS was 40, 39 for Stroop, 47 for Category Test, 49

for R-SAT and IGT, and 48 for EFT.

2.3.3. PET image acquisition

PET scans were acquired with the ECAT/931 scanner (Siemens CTI ECAT EXACT),

at the nuclear medicine facility of the “Virgen de las Nieves” hospital in Granada, (Spain).

For each individual, we obtained 20 minute emission scans 30 minutes after injection of

FDG resting in quiet conditions. The subjects were scanned with their eyes open and ears

unplugged in a dimly illuminated quiet room. Raw data were processed using an iterative

reconstruction (OSEM method: 10 iterations, 32 subsets). Images were reoriented in

transaxial, coronal and saggital planes. For the analyses reported here the reconstructed

images used had 47 axial slices each with an in-plane resolution of 2.57 x 2.57 and a slice

thickness of 3.38mm.

2.4. Data analysis

2.4.1. Preprocessing of PET images

The PET images were converted from DICOM to NIFTI format, were spatially

normalized to the SPM5 PET template (Wellcome Department of Cognitive Neurology,

London, UK) using linear (12 affine) and non-linear normalization. An average of all the

normalized images was obtained and this acted as a study specific PET template. In this

manner we constructed a template in the same modality, from the same scanner as the

images we wanted to normalize. The raw PET images were then spatially normalized with

linear affine and nonlinear parameters to this study-specific template. Visual inspection

revealed this optimized spatial normalization technique produced satisfactory spatial

normalization for the PET image of every volunteer. Finally, the images were smoothed

with an 8mm3 isotropic Gaussian filter.

2.4.2. ROI Analysis

P á g i n a | 9 3

Having achieved satisfactory spatial normalisation we used the Marsbar tool

(http://marsbar.sourceforge.net- Brett et al., 2002) to extract 90 regions of interest (ROIs)

in the entire brain. We used the AAL atlas as a reference for these ROIs (Tzourio-Mazoyer

et al., 2002).

2.4.3. Statistical Analyses

2.4.3.1. Behavioral Analysis

ANCOVA models were conducted to test neuropsychological performance

differences between the groups while controlling for age and education in SPSS 15.0 for

Windows (SPSS Inc., Chicago IL). Cohen´s d estimations of effect sizes were calculated

according to formulations proposed in (Zakzanis, 2001). The association between length

of abstinence and neuropsychological performance was explored using bivariate

correlations in SPSS 15.0.

2.4.3.2. ROI analysis

Relative glucose uptake on standardized uptake value (SUV) were used to obtain

the mean uptake values for each ROI. Next, these values were correlated with the indices

of neuropsychological performance. In a second analysis, those ROIs that showed

significant associations with neuropsychological indices were also correlated with

estimates of severity of drug use. In both cases, we used partial correlations and age as

confounding variable in SPSS 15.0. Results are showed at p≤0.01 and p≤0.05 respectively

(uncorrected for multiple comparisons).

2.4.3.3. Voxel based analysis

Linear regressions were used to identify clusters with a significant correlation

between uptake values and indices of neuropsychological performance using age as

confounding factor. Scans were corrected for differences in global activity by including

P á g i n a | 9 4

proportional scaling in the regression model. We report local maximum peaks that survive

a voxel threshold of p<0.001 (uncorrected for multiple comparisons, cluster size > 100

voxels). Although this threshold offers only modest protection again risk of type I error, it

has been chosen in multiples PET studies under similar assumptions (i.e. Brooks et al.,

2009; Yaouhi et al., 2009). The relationship between duration of abstinence and resting

glucose metabolism was explored by using linear regression models but no associations

survived at a threshold of p<0.001, and therefore it was not further considered. We did not

use treatment seeking as a variable in our analyses because all our volunteers were on

treatment, and more specifically on the same modality of treatment (i.e., therapeutic

community). Finally, we carried out a second analysis in which the mean uptake of the

regions we found to be associated with neuropsychological performance in whole brain

voxel-based analysis were correlated with estimates of drug use using age as confounding

variable in SPSS 15.0. Results are showed at p≤0.01 and p≤0.05 respectively (uncorrected

for multiple comparisons).

3. Results

3.1. Neuropsychological performance

Results showed that SDI had significantly poorer performance than HC on all of the

neuropsychological tests administered (Table 3). Calculation of effect sizes showed large

effect sizes (d>0.8) for LNS (working memory), Stroop (response inhibition), Category

Test (cognitive flexibility), IGT (decision-making) and EFT (emotion processing), and

medium effect sizes (d>0.5) for R-SAT (self-regulation) (Table 3). We only found a

significant negative correlation between length of abstinence and performance on LNS (r=-

0.303).

P á g i n a | 9 5

Table 3 Descriptive scores on the different indices of executive functioning.

Variable Group p-Valuea F-Value d-Value

SDI HC

LNSb 9.75±2.27 15.13±2.33 0.000* 32.12 -2.34

Stroopc -1.76±5.76 4.15±7.24 0.000* 7 -0.92

Category testd 65.67±26 31.80±25.33 0.000* 11.48 1.32

R-SATe 83.78±17.14 93.98±5.71 0.013* 3.83 -0.73

IGTf -1.98±21.06 37.20±26.16 0.000* 17.90 -1.70

Ekman faces testg 47.50±4.49 53±4.61 0.000* 10.03 -1.21

a p<0.05. b Letter and Number sequence total score. c Stroop interference score. d Category test total errors. e Revised strategy application test percentage brief items. f Iowa gambling test total score. g

Ekman faces test total identification.

3.2. ROI analyses

3.2.1. Association between neuropsychological performance and ROIs

These results are shown in Table 4. We found significant correlations between

number of hits on LNS (working memory) and cerebral metabolism (CM) at the right

middle temporal pole (r=-0.367), between R-SAT proportion of brief items (self-

regulation) at the right calcarine (r=0.391) and posterior cingulum bilaterally (left

r=0.440) (right r=0.454), and between IGT net score (decision-making) and the right

middle (r=0.390) and superior frontal cortex (r=0.376). There were not significant

correlations between the remaining tests and the ROIs at p≤0.01.

3.2.2. Association between the ROIs associated with neuropsychological performance and

estimates of drug use

We failed to find significant associations between measures of severity of drug use

and cerebral metabolism at a p<0.01 threshold. By relaxing the alpha level to p<0.05, we

only found a positive correlation between amount of cocaine use and brain metabolism in

the right middle temporal pole (r=0.301).

P á g i n a | 9 6

3.3. Voxel-based analyses

The main results from these analyses are summarized in Table 4. The results

presented in text and table correspond to an alpha level of p<0.001 (uncorrected, cluster

size > 100 voxels), but for demonstration purposes the figures display statistical

parametric maps thresholded at uncorrected voxel level of p<0.01.

P á g i n a | 9 7

Table 4 Relationship between neuropsychological performance and PET uptake using both ROI and voxel based methods.

L, left; R, right; (+) positive association; (-) negative association. Correlation ROI analyses are reported at p≤0.01 (uncorrected). Voxel-based analyses at p<0.001 (uncorrected, cluster size > 100 voxels). Talairach coordinates were obtained by applying the mni2tal function in MATLABR2006a (http://imaging.mrc-cbu.cam.ac.uk/imaging/MniTalairach) to the MNI coordinates obtained from SPM5. The corresponding anatomical names were obtained using the TalairachClient application – version 2.4.2. – (http://www.talairach.org/client.html) for peak and nearest grey matter.

Instrument ROIs Voxel-Based Analysis BA KE t p MNI Talairach

LNS R MiddleTemporal Pole (+) R Superior Temporal gyrus (+) BA22 621 5.23 0.000 52 -30 2 51 -29 3 R Inferior Temporal gyrus (+) BA37 261 3.77 0.000 52 -74 -4 51 -72 0 L Paracentral lobule (-) BA5 502 4.67 0.000 -20 -38 52 -20 -34 50 L Superior Frontal gyrus (-) BA8 449 3.95 0.000 -20 28 52 -20 30 46

Stroop

L Pallidus (-) 191 3.82 0.000 -16 -12 -4 -16 -12 -3 L Middle Frontal gyrus (-) BA10 160 3.72 0.000 -34 64 10 -34 62 6

Category test

L Superior Frontal gyrus (+) BA10 127 4.25 0.000 -26 68 4 -26 66 0 R Middle Frontal gyrus (-) BA6 268 3.59 0.000 50 10 48 50 12 44 L Inferior Parietal Lobule (-) BA40 198 3.74 0.000 -46 -48 38 -46 -45 37

R-SAT R Calcarine (+) R Posterior cingulate (+) L Posterior cingulate (+)

R Posterior cingulate (+) BA30 5811 4.12 0.000 18 -68 10 18 -65 12 R Anterior cingulate (+) BA32 702 3.75 0.000 14 32 16 14 32 13 R Precentral gyrus (+) BA6 416 3.96 0.000 22 -20 64 22 -16 60 R Inferior Temporal gyrus (+) BA20 152 3.58 0.000 40 2 -48 40 -0 -40 L Middle Temporal gyrus (-) BA21 464 4.48 0.000 -56 10 -20 -55 9 -17

IGT R Middle Frontal (+) R Superior Frontal (+)

R Middle Frontal gyrus (+) BA6 945 4.91 0.000 26 -6 60 26 -3 55 L Middle Frontal gyrus (+) BA6 377 3.71 0.000 -28 2 54 -28 4 50 R Middle Frontal gyrus (+) BA6 222 3.94 0.000 54 -12 56 53 14 51 L Superior Frontal gyrus (+) BA10 212 3.67 0.000 -18 68 14 -18 67 10 L Superior Frontal gyrus (+) BA8 141 3.72 0.000 -20 44 52 -20 45 46

Ekman Faces Test Lingual gyrus (-) BA18 308 4.10 0.000 0 -102 -14 0 -99 -7 Fusiform gyrus (-) BA19 191 3.90 0.000 -36 -78 -18 -36 -76 -11

P á g i n a | 9 8

3.3.1. Cold executive functions –Whole brain voxel-based correlates in SDI

For working memory, there was a positive correlation between LNS and one

cluster with peaks at the superior temporal gyrus (51 -29 3) and inferior temporal gyrus

(51 -72 0); this cluster included right superior and inferior lateral temporal regions

extending medially. LNS negatively correlated with a cluster in the left postcentral region

extending to parietal cortex, precuneus and paracentral lobule (peak at paracentral lobule,

-20 -34 50), and a second cluster which encompassed regions in the left frontal cortex

extending to the left DLPFC (peak at superior frontal gyrus, -20 30 46) (Figure 1 –Panel A).

For response inhibition, there was a negative correlation between the Stroop

interference score and two different clusters. The first cluster correspond to the left

thalamus, hippocampus and pallidum (peak at globus pallidus, -16 -12 -3), and the second

cluster encompassed the DLPFC bilaterally extending to the superior frontal cortex (peak

at middle frontal gyrus, -34 62 6) (Figure 1 –Panel B).

For cognitive flexibility, results showed a positive correlation between Category

Test total number of errors and a cluster which included the DLPFC and superior frontal

gyrus (peak at superior frontal gyrus, -26 66 0). There were also negative correlations

between task performance and two other clusters. The first was at the DLPFC extending to

the precentral gyrus with peak at middle frontal gyrus (50 12 44), and the second in the

parietal cortex extending to the supramarginal and angular area (peak at inferior parietal

lobule, -46 -45 37) (Figure 1 –Panel C).

P á g i n a | 9 9

Figure 1 Images showing the areas where cold executive functions correlated with brain

metabolism.

A. Letter Number sequencing negative (blue) and positive (yellow) correlations.

L -20 52 R

B. Stroop negative correlations (blue).

L -34 -16 R

C. Category Test negative (blue) and positive (yellow) correlations.

L -45 -26 45 R

In summary, working memory and cognitive flexibility performance negatively

correlated with cerebral metabolism in the left DLPFC and left parietal regions, whereas

response inhibition performance negatively correlated with bilateral DLPFC and basal

ganglia. We found positive correlations between working memory performance and

P á g i n a | 1 0 0

cerebral metabolism in the temporal gyrus, and between cognitive flexibility and cerebral

metabolism in the superior frontal gyrus.

3.3.2. Hot executive functions –Whole brain voxel-based correlates in SDI

For the R-SAT (self-regulation) results showed significant positive correlations

with four different clusters. One of them had a peak at the posterior cingulate (18 -65 12),

including bilateral middle and posterior cingulate cortex and right precuneus. A second

cluster had a statistical peak at the anterior cingulate (14 32 13) extending to the insula. A

third cluster encompassed regions of the precentral gyrus extending to superior frontal

cortex and to the postcentral gyrus and supplementary motor area (peak at precentral

gyrus 22 -16 60). The fourth cluster had a statistical peak at the inferior temporal gyrus

(40 -0 -40) extending to the temporal pole and fusiform. We also found a negative

correlation between R-SAT performance and a cluster localized in the temporal cortex

which included the middle and superior temporal gyrus and the temporal pole (peak at

the middle temporal gyrus -55 9 -17) (Figure 2 –Panel A).

For decision-making, there was a positive correlation between IGT scores and

three clusters. The first cluster included the DLPFC extending to posterior areas (peak at

the right middle frontal gyrus 26 -3 55), a second cluster which included the DLPFC

bilaterally (peak at left middle frontal gyrus -28 4 50 and right middle frontal gyrus 53 14

51) and two clusters with peak at the superior frontal gyrus (-18 67 10; -20 45 46) (Figure

2 –Panel B).

For emotion processing, there was a negative correlation between EFT emotion

recognition hits and two main clusters. One including the lingual gyrus (peak at lingual

gyrus, 0 -99 -7) and a second cluster including the cerebellum and fusiform and lingual

gyrus (peak at the fusiform gyrus -36 -76 -11) (Figure 2 –Panel C).

P á g i n a | 1 0 1

Figure 2 Images showing the areas where hot executive functions correlated with brain

metabolism.

A. Revised Strategy Application Test negative (blue) and positive (yellow) correlations.

L -56 8 35 R

B. Iowa Gambling Task positive correlations (yellow).

L -26 8 35 R

C. Ekman Faces Test negative correlations (blue).

L -37 -32 R

In summary, we found positive correlations between self-regulation performance

and cerebral metabolism within different sections of the cingulate cortex, superior frontal

cortex and inferior temporal gyrus, and between decision-making performance and

DLPFC/superior frontal metabolism. Conversely, we found negative correlations between

P á g i n a | 1 0 2

self-regulation performance and cerebral metabolism in the superior temporal gyrus, and

between emotion recognition performance and cerebral metabolism in the fusiform,

cerebellum and lingual gyrus.

3.3.3. The relationship between voxel-based derived regions associated with

neuropsychological performance and estimates of drug use

We only found a significant positive correlation between amount of heroin used

and metabolism in the left middle temporal gyrus (r=0.377) at p<0.01. By relaxing the

alpha level to p<0.05, we found positive correlations between duration of cannabis

(r=0.330) and cocaine used (r=0.296) and the right precentral gyrus, and negative

correlations between the amount of cannabis used and the left middle frontal gyrus (r=-

0.370) and left inferior parietal lobe (r=-0.307), the amount of heroin used and the right

inferior temporal gyrus (r=-0.308), and the amount of alcohol used and the right middle

frontal gyrus (r=-0.297).

4. Discussion

This study investigated the association between the neuropsychological

performance of SDI on measures of cold and hot executive functions (as measured outside

the scanner) and PET indexed regional brain glucose metabolism. SDI were also compared

to non-drug using controls with respect to behavioral performance in these tasks.

We found that SDI had significantly poorer performance than HC individuals on all

of the neuropsychological tests administered. Likewise, the neuropsychological

performance in the SDI was just affected by the duration of abstinence in the LNS task,

allowing us to rule out the effects of this variable on all other tasks. On the other hand, in

accordance with our initial hypotheses, we found that both cold and hot executive tests

were significantly associated with common brain regions, including the DLPFC, mid-

superior frontal gyrus and superior, inferior and middle temporal gyrus. All these regions

P á g i n a | 1 0 3

are thought to play important roles in the executive control of goal-driven behavior

(Bechara and Van Der Linden, 2005), and in neuroadaptations related to addiction (Koob

and Volkow, 2010); therefore, our findings highlight the relevance of this set of regions for

a wide range of cognitive functions that are significantly impaired in SDI. In terms of

differential associations, cold executive tests mainly correlated with frontal, temporal and

parietal regions, whereas hot executive functions correlated with midline mid-superior

frontal and cingulate regions in both ROI and whole brain voxel-based analyses. These

results are in agreement with the notion that different executive processes rely in both

common and partially dissociable neural networks (Collette et al., 2006), and that different

neural circuits are relevant to various aspects of abnormal cognitive functioning in SDI

(Verdejo-García et al., 2004). .

Specifically, our results showed that verbal working memory (indexed by LNS) is

mainly associated with DLPFC and temporo-parietal regions (Figure 1 –Panel A), which is

consistent with previous findings suggesting these regions are the key neural substrates

for memory updating and manipulation (Collette and Van der Linden, 2002; D’Esposito,

2007). For response inhibition, results showed associations between Stroop performance

and DLPFC and basal ganglia metabolism (Figure 1 –Panel B). This finding is in agreement

with evidence from cognitive and neuroimaging studies showing that fronto-striatal

alterations related to inhibitory control are ubiquitous across different forms of addiction,

and are associated with difficulties in ceasing drug use (Feil et al., 2010). For cognitive

flexibility results, we showed associations between the number of errors and DLPFC and

parietal cortex resting metabolism (Figure 1 –Panel C), replicating previous findings using

this probe and other switching perseveration tests (e.g., the Wisconsin Card Sorting Test)

in alcoholics (Adams et al., 1993) and healthy subjects (Thomas et al., 2011).

With regard to hot executive functions, results showed that R-SAT performance

was associated with resting metabolism in the mid-posterior and ACC, insula and temporal

P á g i n a | 1 0 4

regions (Figure 2 –Panel A). Both insula and ACC are involved in task-level control and

self-regulation of behavior (Nelson et al., 2010; Posner et al., 2007), the key cognitive

processes needed to develop and implement goal-driven adjustment of performance (as

indexed by this task). ACC and insula dysfunctions have been associated with stress-

induced craving, poor inhibitory control, impaired insight and less motivation to change in

SDI (Goldstein et al., 2009; Li and Sinha, 2008; Romero et al., 2010; Sinha and Li, 2007);

this range of symptoms may relate to core deficits of self-regulation (i.e., difficulties to set

and adjust behaviour according to long-term goals) in addicted individuals. Furthermore,

both anterior and posterior cingulate regions have been previously associated with

abnormal performance on multi-tasking measures, specifically with deficits in task

learning and remembering, rule-breaking and task switching (Burgess et al., 2000).

According to previous studies using similar methodologies, IGT performance in

mid-term abstinent SDI is associated with resting metabolism in both OFC and DLPFC

(Adinoff et al., 2003; Krain et al., 2006; Tucker et al., 2004) (Figure 2 –Panel B). This

finding is indicative of the notion that both cold processes, such as working memory or

switching skills relying on DLPFC functioning, and hot processes, such as monitoring of

reward value, relying on OFC functioning play a role in decision-making performance in

SDI (Bolla et al., 2003). Finally, although we failed to find significant associations between

ventral fronto-striatal regions and emotional perception, we found robust associations

between scores in this task and two areas (lingual and fusiform gyrus) (Figure 2 –Panel C)

that have been consistently associated with face processing in numerous studies

(Dinkelacker et al., 2011; Kanwisher and Yovel, 2006). The association between emotion

processing and the ventromedial frontal stream could be better grasped by fMRI studies

monitoring brain activation during actual emotional arousal.

Finally, we found significant associations between patterns of drug use and the

regions associated with executive performance in both ROI and whole brain voxel-based

P á g i n a | 1 0 5

analysis. Although mostly exploratory, some of these effects are informative about the link

between the use of particular substances and the brain regions supporting different

executive functions. For example, higher metabolism in the middle temporal gyrus was

associated with both R-SAT performance and severity of heroin use, in agreement with

data from resting fMRI showing left middle temporal gyrus dysfunction in chronic heroin

users (Jiang et al., 2011). Higher metabolism in the right precentral gyrus was associated

with both severity of cocaine use and cognitive flexibility, such that we may speculatively

suggest a role of this region in cocaine-induced motor perseveration deficits (Ersche et al.,

2008). Lower metabolism in the superior frontal gyrus was associated with both severity

of alcohol use and performance on cold executive tasks and on decision-making; this result

is fitting with the findings of Goldstein et al. (2004) who found that severity of alcohol use

was the main predictor of a composite index of executive impairment, which was

predicted by regional metabolism in the anterior cingulate. Finally, lower metabolism in

the adjacent middle frontal gyrus was linked to severity of cannabis use and performance

on response inhibition and decision-making, fitting with evidence of impairment of these

functions in chronic cannabis users (Battisti et al., 2010; Verdejo-Garcia et al., 2007).

These associations are in agreement with the notion that amount and duration of drug use

impact brain functioning in key regions for executive control, supporting the relevance of

the correlations between executive performance and brain metabolism in this clinical

group of SDI.

Several possible limitations of our study should be considered and addressed by

future research. In the first place, the methodology used in this study does not allow us to

monitor brain activity as a direct response to neuropsychological challenges; instead it

shows deficits in resting glucose metabolism that were related to test performance.

Therefore, this study cannot attribute NP function to an inability of the substance

dependent group to activate specific regions of the brain. FMRI or PET designs that involve

scanning during actual test performance may be better suited to test the neuronal

P á g i n a | 1 0 6

substrates of executive dysfunction in SDI. On the other hand, this design allowed us to

explore the brain correlates of highly complex neuropsychological paradigms, such as the

R-SAT that need to be simplified to fit them in fMRI designs. Furthermore, PET was

altogether not obtained in the control group and therefore it is not clear whether

performance differences on the NP tasks were associated with actual abnormalities (i.e.

compared to health) in glucose metabolism at rest. Future studies should examine if there

are differences in these associations between a clinical group and a control group. A fourth

limitation was the lack of estimates of nicotine use/dependence, considering that previous

studies have observed significant alterations in PET-indexed cerebral metabolism

associated with nicotine use (i.e. Rose et al., 2003; Zubieta et al., 2005) and should be

assess in any PET study of drug use. The voxel based analyses proved to be overall more

sensitive than the ROI method, this is an expected outcome because the voxel based

method is used to carry out statistical tests in smaller brain units (voxels) than the ROI

analysis (whole gyri). Finally, due to its correlational design, this study cannot yield

conclusions about cause-effect relationships in the area of the neuropsychology of drug

addiction.

Taken together, our results show that SDI have neuropsychological deficits in both

cold and hot executive functions. These deficits are associated with brain metabolism in

both common and partially dissociable neural networks encompassing frontal, parietal,

temporal and basal ganglia regions.

P á g i n a | 1 0 7

References

Adams, K.M., Gilman, S., Koeppe, R.A., Kluin, K.J., Brunberg, J.A., Dede, D., Berent, S., Kroll,

P.D., 1993. Neuropsychological deficits are correlated with frontal hypometabolism in

positron emission tomography studies of older alcoholic patients. Alcoholism, Clinical

and Experimental Research 17, 205-210.

Adinoff, B., Devous, M.D., Cooper, D.B., Best, S.E., Chandler, P., Harris, T., Cervin, C.A.,

Cullum, C.M., 2003. Resting regional cerebral blood flow and gambling task

performance in cocaine-dependent subjects and healthy comparison subjects. The

American Journal of Psychiatry 160, 1892-1894.

Anderson, S.W., Barrash, J., Bechara, A., Tranel, D., 2006. Impairments of emotion and real-

world complex behavior following childhood- or adult-onset damage to ventromedial

prefrontal cortex. Journal of the International Neuropsychological Society 12, 224-235.

Baler, R.D., Volkow, N.D., 2006. Drug addiction: the neurobiology of disrupted self-control.

Trends in Molecular Medicine 12, 559-566.

Battisti, R.A., Roodenrys, S., Johnstone, S.J., Pesa, N., Hermens, D.F., Solowij, N., 2010.

Chronic cannabis users show altered neurophysiological functioning on Stroop task

conflict resolution. Psychopharmacology (Berl) 212, 613-624.

Bechara, A., Damasio, A.R., Damasio, H., Anderson, S.W., 1994. Insensitivity to future

consequences following damage to the human prefrontal cortex. Cognition 50, 7-15.

Bechara, A., Van Der Linden, M., 2005. Decision-making and impulse control after frontal

lobe injuries. Current Opinion in Neurology 18, 734-739.

Bolla, K., Ernst, M., Kiehl, K., Mouratidis, M., Eldreth, D., Contoreggi, C., Matochik, J., Kurian,

V., Cadet, J., Kimes, A., Funderburk, F., London, E., 2004. Prefrontal cortical dysfunction

P á g i n a | 1 0 8

in abstinent cocaine abusers. The Journal of Neuropsychiatry and Clinical

Neurosciences 16, 456-464.

Bolla, K.I., Eldreth, D.A., London, E.D., Kiehl, K.A., Mouratidis, M., Contoreggi, C., Matochik,

J.A., Kurian, V., Cadet, J.L., Kimes, A.S., Funderburk, F.R., Ernst, M., 2003. Orbitofrontal

cortex dysfunction in abstinent cocaine abusers performing a decision-making task.

NeuroImage 19, 1085-1094.

Brett, M., Anton, J.L., Valabregue, R., Poline, J.B., 2002. Region of interest analysis using an

SPM toolbox. NeuroImage 16, 2.

Brooks, J.O. 3rd, Rosen, A.C., Hoblyn, J.C., Woodard, S.A., Krasnykh, O., Ketter, T.A., 2009.

Resting prefrontal hypometabolism and paralimbic hypermetabolism related to verbal

recall deficits in euthymic older adults with bipolar disorder. The American Journal of

Geriatric Psychiatry 17, 1022-1029.

Burgess, P.W., Veitch, E., de Lacy Costello, A., Shallice, T., 2000. The cognitive and

neuroanatomical correlates of multitasking. Neuropsychologia 38, 848-863.

Collette, F., Hogge, M., Salmon, E., Van der Linden, M., 2006. Exploration of the neural

substrates of executive functioning by functional neuroimaging. Neuroscience 139,

209-221.

Collette, F., Van der Linden, M., 2002. Brain imaging of the central executive component of

working memory. Neuroscience and Biobehavioral Reviews 26, 105-125.

Dao-Castellana, M.H., Samson, Y., Legault, F., Martinot, J.L., Aubin, H.J., Crouzel, C., Feldman,

L., Barrucand, D., Rancurel, G., Féline, A., Syrota, A., 1998. Frontal dysfunction in

neurologically normal chronic alcoholic subjects: metabolic and neuropsychological

findings. Psychological Medicine 28, 1039-1048.

P á g i n a | 1 0 9

DeFilippis, N.A., 2002. The Category Test, Computer Version. 2nd ed. Odessa, Florida,

Psychological Assessment Resources.

Del Ser, T., García-Montalvo, J.I., Martínez Espinosa, S., Delgado Villapalos, C., Bermejo, F.,

1997. The estimation of premorbid intelligence in Spanish people with the "Word

Accentuation Test" and its application to the diagnosis of dementia. Brain & Cognition

33, 343-356.

Desmond, J.E., Chen, S.H., DeRosa, E., Pryor, M.R., Pfefferbaum, A., Sullivan, E.V., 2003.

Increased frontocerebellar activation in alcoholics during verbal working memory: an

fMRI study. NeuroImage 19, 1510-1520.

D'Esposito, M., 2007. From cognitive to neural models of working memory. Philosophical

Transactions of the Royal Society B: Biological Sciences 29, 761-772.

Dinkelacker, V., Grüter, M., Klaver, P., Grüter, T., Specht, K., Weis, S., Kennerknecht, I., Elger,

C.E., Fernandez, G., 2010. Congenital prosopagnosia: multistage anatomical and

functional deficits in face processing circuitry. Journal of Neurology 258, 770-782.

Eldreth, D.A., Matochik, J.A., Cadet, J.L., Bolla, K.I., 2004. Abnormal brain activity in

prefrontal brain regions in abstinent marijuana users. NeuroImage 23, 914-920.

Ersche, K.D., Roiser, J.P., Robbins, T.W., Sahakian, B.J., 2008. Chronic cocaine but not

chronic amphetamine use is associated with perseverative responding in humans.

Psychopharmacology 197, 421-431.

Feil, J., Sheppard, D., Fitzgerald, P.B., Yücel, M., Lubman, D.I., Bradshaw, J.L., 2010.

Addiction, compulsive drug seeking, and the role of frontostriatal mechanisms in

regulating inhibitory control. Neuroscience and Biobehaviorals Reviews 35, 248-275.

P á g i n a | 1 1 0

Fernández-Serrano, M.J., Pérez-García, M., Verdejo-García, A., 2011. What are the specific

vs. generalized effects of drugs of abuse on neuropsychological performance?.

Neuroscience and Biobehavioral Reviews 35, 377-406.

Forman, S.D., Dougherty, G.G., Casey, B.J., Siegle, G.J., Braver, T.S., Barch, D.S., Stenger, V.A.,

Wick-Hull, C., Pisarov, L.A., Lorensen, E., 2004. Opiate addicts lack error-dependent

activation of rostral anterior cingulate. Biological Psychiatry 55, 531-537.

Gilman, S., Adams, K., Koeppe, R.A., Berent, S., Kluin, K.J., Modell, J.G., Kroll, P., Brunberg,

J.A., 1990. Cerebellar and frontal hypometabolism in alcoholic cerebellar degeneration

studied with positron emission tomography. Annals of Neurology 28, 775-785.

Golden, C.J., 1978. Stroop Color and Word Tets. A manual for clinical and experimental

uses. Wood Dale, Illinois, Stoelting Co.

Goldstein, R.Z., Craig, A.D., Bechara, A., Garavan, H., Childress, A.R., Paulus, M.P., Volkow,

N.D., 2009. The neurocircuitry of impaired insight in drug addiction. Trends in

Cognitive Sciences 9, 372-380.

Goldstein, R.Z., Leskowjan, A.C., Hoff, A.L., Hitzemann, R., Bashan, F., Khalsa, S., Wang, G.J.,

Fowler, J.S., Volkow, N.D., 2004. Severity of neuropsychological impairment in cocaine

and alcohol addiction: association with metabolism in the prefrontal cortex.

Neuropsychologia 42, 1447-1458.

Goldstein, R.Z., Volkow, N.D., 2002. Drug addiction and its underlying neurobiological

basis: neuroimaging evidence for the involvement of the frontal cortex. The American

Journal of Psychiatry 159, 1642-1652.

Halpern, J.H., Pope, H.G. Jr., Sherwood, A.R., Barry, S., Hudson, J.I., Yurgelun-Todd, D., 2004.

Residual neuropsychological effects of illicit 3,4-methylenedioxymethamphetamine

P á g i n a | 1 1 1

(MDMA) in individuals with minimal exposure to other drugs. Drug and Alcohol

Dependence 75, 135-147.

Hartstra, E., Oldenburg, J.F., Van Leijenhorst, L., Rombouts, S.A., Crone, E.A., 2010. Brain

regions involved in the learning and application of reward rules in a two-deck

gambling task. Neuropsychologia 48, 1438-1446.

Heberlein, A.S., Padon, A.A., Gillihan, S.J., Farah, M.J., Fellows, L.K., 2008. Ventromedial

frontal lobe plays a critical role in facial emotion recognition. Journal of Cognitive

Neuroscience 20, 721-733.

Jager, G., Kahn, R.S., Van Den Brink, W., Van Ree, J.M., Ramsey, N.F., 2006. Long-term effects

of frequent cannabis use on working memory and attention: an fMRI study.

Psychopharmacology 185, 358-368.

Jiang, G.H., Qiu, Y.W., Zhang, X.L., Han, L.J., Lv, X.F., Li, L.M., Lin, C.L., Zhuo, F.Z., Hu, S.Y., Tian,

J.Z., 2011. Amplitude low-frequency oscillation abnormalities in the heroin users: a

resting state fMRI study. Neuroimage 57,149-154.

Kanwisher, N., Yovel, G., 2006. The fusiform face area: a cortical region specialized for the

perception of faces. Philosophical Transactions of the Royal Society of London. Series

B, Biological Sciences 361, 2109-2128.

Kerr, A., Zelazo, P.D., 2004. Development of "hot" executive function: the children's

gambling task. Brain and Cognition 55, 148-157.

Koob, G.F., Volkow, N.D., 2010. Neurocircuitry of addiction. Neuropsychopharmacology 35,

217-238.

Krain, A.L., Wilson, A.M., Arbuckle, R., Castellanos, F.X., Milham, M.P., 2006. Distinct neural

mechanisms of risk and ambiguity: a meta-analysis of decision-making. NeuroImage

32, 477-484.

P á g i n a | 1 1 2

Lee, T.M., Zhou, W.H., Luo, X.J., Yuen, K.S., Ruan, X.Z., Weng, X.C., 2005. Neural activity

associated with cognitive regulation in heroin users: A fMRI study. Neuroscience

Letters 382, 211-216.

Levine, B., Dawson, D., Boutet, I., Schwartz, M.L., Stuss, D.T., 2000. Assessment of strategic

self-regulation in traumatic brain injury: its relationship to injury severity and

psychosocial outcome. Neuropsychology 14, 491-500.

Li, C.S., Sinha, R., 2008. Inhibitory control and emotional stress regulation: neuroimaging

evidence for frontal-limbic dysfunction in psycho-stimulant addiction. Neuroscience

and Biobehavioral Reviews 32, 581-597.

López-Torrecillas, F., Godoy, J.F., Pérez-García, M., Godoy, D., Sánchez-Barrera, M., 2001.

Variables modulating stress and coping that discriminate drug consumers from low or

not drug consumers. Addictive Behaviors 25, 161-165.

Lubman, D.I., Yücel, M., Kettle, J.W., Scaffidi, A., Mackenzie, T., Simmons, J.G., Allen, N.B.,

2009. Responsiveness to drug cues and natural rewards in opiate addiction:

associations with later heroin use. Archives of General Psychiatry 66, 205-212.

Lundqvist, T., 2010. Imaging cognitive deficits in drug abuse. Current Topics in Behavioral

Neurosciences 3, 247-275.

Marinkovic, K., Oscar-Berman, M., Urban, T., O’Reilly, C.E., Howard, J.A., Sawyer, K., Harris,

G.J., 2009. Alcoholism and dampened temporal limbic activation to emotional faces.

Alcoholism, Clinical and Experimental Research 33, 1880-1892.

Nelson, S.M., Dosenbach, N.U., Cohen, A.L., Wheeler, M.E., Schlaggar, B.L., Petersen, S.E.,

2010. Role of the anterior insula in task-level control and focal attention. Brain

Structure and Function 214, 669-680.

P á g i n a | 1 1 3

Passetti, F., Clark, L., Mehta, M.A., Joyce, E., King, M., 2008. Neuropsychological predictors

of clinical outcome in opiate addiction. Drug and Alcohol Dependence 94, 82-91.

Paulus, M.P., Tapert, S.F., Schuckit, M.A., 2005. Neural activation patterns of

methamphetamine-dependent subjects during decision making predict relapse.

Archives of General Psychiatry 62, 761-768.

Posner, M.I., Rothbart, M.K., Sheese, B.E., Tang, Y., 2007. The anterior cingulate gyrus and

the mechanism of self-regulation. Cognitive, Affective and Behavioral Neuroscience 7,

391-395.

Romero, M.J., Asensio, S., Palau, C., Sanchez, A., Romero, F.J., 2010. Cocaine addiction:

diffusion tensor imaging study of the inferior frontal and anterior cingulate white

matter. Psychiatry Research 181, 57-63.

Rose, J.E., Behm, F.M., Westman, E.C., Mathew, R.J., London, E.D., Hawk, T.C., Turkington,

T.G., Coleman, R.E., 2003. PET studies of the influences of nicotine on neural systems in

cigarette smokers. American Journal of Psychiatry 160, 323-333.

Salloum, J.B., Ramchandani, V.A., Bodurka, J., Rawlings, R., Momenan, R., George, D.,

Hommer, D.W., 2007. Blunted rostral anterior cingulate response during a simplified

decoding task of negative emotional facial expressions in alcoholic patients.

Alcoholism, Clinical and Experimental Research 31, 1490-1504.

Sinha, R., Li, C.S., 2007. Imaging stress- and cue-induced drug and alcohol craving:

association with relapse and clinical implications. Drug and Alcohol Review, 26, 25-31.

Thomas, L.A., Hall, J.M., Skup, M., Jenkins, S.E., Pine, D.S., Leibenluft, E., 2011. A

developmental neuroimaging investigation of the change paradigm. Developmental

Science 14, 148-161.

P á g i n a | 1 1 4

Tomasi, D., Goldstein, R.Z., Telang, F., Maloney, T., Alia-Klein, N., Caparelli, E.C., Volkow,

N.D., 2007. Widespread disruption in brain activation patterns to a working memory

task during cocaine abstinence. Brain Research 1171, 83-92.

Tucker, K.A., Potenza, M.N., Beauvais, J.E., Browndyke, J.N., Gottschalk, P.C., Kosten, T.R.,

2004. Perfusion abnormalities and decision making in cocaine dependence. Biological

Psychiatry 56, 527-530.

Tzourio-Mazoyer, N., Landeau, B., Papathanassiou, D., Crivello, F., Etard, O., Delcroix, N.,

Mazoyer, B., Joliot, M., 2002. Automated anatomical labelling of activations in SPM

using a macroscopic anatomical parcellation of the MNI MRI single subject brain.

NeuroImage 15, 273-289.

Verdejo-García, A., Bechara, A., 2009. A somatic marker theory of addiction.

Neuropharmacology 56, 48-62.

Verdejo-Garcia, A., Benbrook, A., Funderburk, F., David, P., Cadet, J.L., Bolla, K.I., 2007. The

differential relationship between cocaine use and marijuana use on decision-making

performance over repeat testing with the Iowa Gambling Task. Drug and Alcohol

Dependence 90, 2-11.

Verdejo-García, A., López-Torrecillas, F., Orozco, C., Pérez-García, M., 2004. Clinical

implications and methodological challenges in the study of the neuropsychological

correlates of cannabis, stimulant and opioid abuse. Neuropsychology Review 14, 1-41.

Verdejo-García, A., Rivas-Pérez, C., Vilar-López, R., Pérez-García, M., 2007. Strategic self-

regulation, decision-making and emotion processing in poly-substance abusers in their

first year of abstinence. Drug and Alcohol Dependence 12, 139-146.

Wechsler, D., 1997. Wechsler adult intelligence scale. 3rd ed. San Antonio, Texas, The

Psycological Corporation.

P á g i n a | 1 1 5

Yaouhi, K., Bertran, F., Clochon, P., Mézenge, F., Denise, P., Foret, J., Eustache, F.,

Desgranges, B., 2009. A combined neuropsychological and brain imaging study of

obstructive sleep apnea. Journal of Sleep Research 18, 36-48.

Young, A.W., Perrett, D.I., Calder, A.J., Sprengelmeyer, R., Ekman, P., 2002. Facial

expressions of emotion: stimuli and tests (FEEST). Bury St. Edmunds, Thames Valley

Test Company.

Zakzanis, K.K., 2001. Statistics to tell the truth, the whole truth, and nothing but the truth:

formulae, illustrative numerical examples, and heuristic interpretation of effects size

analyses for neuropsychological researchers. Archives of Clinical Neuropsychology 16,

653-677.

Zubieta, J.K., Heitzeg, M.M., Xu, Y., Koeppe, R.A., Ni, L., Guthrie, S., Domino, E.F., 2005.

Regional cerebral blood flow responses to smoking in tobacco smokers after overnight

abstinence. American Journal of Psychiatry 162, 567-577.

P á g i n a | 1 1 6

Supplementary Material

This document included the description of the Interview for Research on Addictive

Behaviour and the neuropsychological tasks used in the study.

The Interview for Research on Addictive Behaviour (IRAB; López-Torrecillas et al.,

2001) was used to examine the severity of drug use. This test evaluates, by means of a

brief interview, the quantity (average dose), frequency (consumption episodes by month),

and duration (years of duration) of the use of a series of drugs of abuse, including cocaine,

heroin, and alcohol, which were the drugs of abuse that motivated treatment requests in

the substance dependent individuals (SDI) included in this sample. For every substance

the participant had actually used, the following information was requested: (1) the

average dose of each target drug taken in each episode of use (number of grams for

cocaine and heroin and number of units for alcohol, considering that 25 cl. of hard liquor

(e.g., scotch) equals 2 units, while 25 cl. of wine or beer equals 1 unit); (2) the frequency of

substance use per month (daily, between one and three times upon a week, once a week,

between one and three times upon a month, or once a month); and (3) the number of

years that elapsed since the onset of the use. From this information, we obtained two

indices for each of the drugs used: (1) Amount (average dose x frequency), an index of

monthly use of the drug; and (2) Duration of drug use, measuring number of years of

exposure to the drug.

The description of the different neuropsychological tests used in the study is

detailed below.

Working memory: Letter Number Sequencing (LNS) (Wechsler Adult Intelligence

Scale -WAIS-III, Wechsler, 1997). The participant is read a sequence in which letters and

numbers are combined, and he or she is asked to reproduce the sequence heard, first

placing the numbers in ascending order and then the letters in alphabetical order. The test

consists of a series of elements of increasing complexity (progressively higher working

P á g i n a | 1 1 7

memory demands). Each element includes three trials, and in each trial the sequence is

read at one letter or number per second. The test is stopped when the participant misses

three trials of the same element. The main dependent variable used on this test was the

total number of hits.

Response inhibition: Stroop (Golden, 1978). This test consists of three forms, each

containing 100 elements distributed in five columns of 20 elements each. The first form is

made up of the words “RED”, “GREEN”, and “BLUE” ordered randomly and printed in black

ink. In this condition the participant is asked to read aloud, as quickly as possible, the

words written on this page in a time set at 45 seconds. The second form consists of strings

of “XXXX” printed in red, blue, or green ink. In this condition, the participant is asked to

overtly name as quickly as possible the color of these elements with a time limit of 45

seconds. The third form introduces the condition of interference, and it consists of the

words from the first form printed in the colors of the second. In this condition, the subject

is asked to name the color of the ink the word is written in, ignoring the word, also in 45

seconds. The main dependent variable used in this test was the interference score.

Cognitive flexibility: Category Test (DeFilippis, 2002). A computerized version of

this test was administered. The task consists of 208 stimuli that have different types of

designs grouped in seven subtests. For all the stimuli included in the same subtest, there is

an underlying rule that determines the appropriateness of the responses throughout this

subtest. However, this rule changes in the next subtest, so that the performance on the test

depends on the ability of the participant to infer these rules, and modify them when they

are no longer valid. Test instructions are intentionally ambiguous: we explained to the

participants that different types of designs will consecutively show up on the screen, and

that each design was associated with one of the first four numbers: 1, 2, 3, or 4. The

participant must press, for each stimulus, the key she/he thinks is associated with that

design; then the computer provides auditory feedback about the correctness of the

P á g i n a | 1 1 8

response. Therefore, participants have to use feedback to update response strategies

according to the changing rules, making this test a solid index of cognitive flexibility. The

main index of performance on the test was the total number of errors across the seven

subtests.

Self-regulation: Revised Strategy Application Test (R-SAT) (Levine et al., 2000). This

task is an unstructured paper and pencil test of strategy usage that is sensitive to

disturbed self-regulation. The R-SAT consisted of three simple activities: figure tracing,

sentence copying and object numbering. Each activity was displayed in two different

stacks (A and B), each containing 120 items. The items differed in two dimensions: size

(large vs. small) and the amount of time required to complete them (brief vs. medium vs.

long). Participants were instructed to perform these three activities during a 10 minute

time frame, trying to do some work on each of the stacks. The main goal of the task was to

win as many points as possible, considering that large items scored 0 points and small

items scored 100 points each. Additionally, completion of some items (those printed on

pages containing hand-drawn faces) led to the loss of all the points previously achieved.

Points were used in the instructions to see if participants would respond accordingly, but

the dependent variable in this task is the number of items and not the points. Given the

limited time, the most efficient strategy on this test (which the participant must discover)

is to complete the brief items and exclude the lengthy items. This requires the inhibition of

a tendency to complete all the items in sequence, which is established on the early pages,

where all the items are brief. The main dependent variable from the R-SAT was the

proportion of brief items completed with regard to the total number of items attempted.

Decision-making: Iowa Gambling Task (IGT) (Bechara et al., 1994). This task has

been proposed to be sensitive to real-life decision-making deficits. We used a

computerized version of the test (Bechara et al., 2001) four decks labelled A’, B’, C’, and D’,

with 60 cards each are displayed on the computer screen. The participant has to choose

P á g i n a | 1 1 9

cards from these decks during 100 trials with the aim of winning as much money as

possible, or in case she/he cannot win, at least try not to lose. Every time the participant

chooses a card she/he receives a varying amount of money and positive feedback (a

smiling face); however, in some cards, after winning she/he can also lose a varying

amount of money, accompanied by a frowning face. Decks C and D are considered

advantageous, because they provide little reward but also minimal punishment; leading to

substantial winning over time. On the other hand, decks A and B are considered

disadvantageous, because they provide high reward but also disproportionate losses;

leading to bankruptcy over time. The participant can actually track the money she/he wins

or loose watching a green bar on top of the screen. Good performance on the IGT implies

that the participant selects on average more cards from decks C’ and D’ (advantageous

decks: lower immediate gain but smaller future loss) than from decks A’ and B’

(disadvantageous decks: high immediate gain but larger future loss). Therefore, the

primary dependent variable from this task was the difference between the number of

cards selected from the advantageous minus the number of cards selected from the

disadvantageous decks: [(C+D)-(A+B)].

Emotion Recognition: Ekman Faces Test (EFT). This test is a computer task

assessing recognition of facial emotional expressions. The task uses stimuli from the

standardized test: “Facial Expressions of Emotion: Stimuli and Test” (FEEST; Young et al.,

2002). A series of 60 stimuli featuring faces portraying basic emotions were presented and

participants were asked to identify the emotion displayed by each face. Faces depicted

expressions of anger, disgust, fear, happiness, sadness and surprise (6 emotions, 10 faces

each). The total number of correct identifications was the main dependent variable from

this task.

P á g i n a | 1 2 0

P á g i n a | 1 2 1

[DISCUSIÓN]

P á g i n a | 1 2 2

1. Discusión General

La revisión de la literatura realizada al inicio de nuestra investigación nos permitió

subrayar la importancia que ciertos rasgos de personalidad como la impulsividad tenían

en las diferentes fases del proceso adictivo y la delimitación de ciertas inconsistencias y

limitaciones presentes en muchos de los estudios llevados a cabo en individuos

drogodependientes. Estas limitaciones, entre las que cabe destacar la repercusión

conjunta de factores disposicionales (rasgos premórbidos) y deterioros inducidos por el

consumo de drogas en la neurobiología de la adicción, los efectos del policonsumo de

distintas sustancias y los patrones heterogéneos de severidad de consumo de estas

sustancias presentes en estas poblaciones, han dificultado en gran medida la

interpretación de los resultados obtenidos por estudios previos, así como el

establecimiento de conclusiones sólidas sobre la asociación entre el consumo de drogas y

el cerebro.

Teniendo en cuenta estos antecedentes, el Objetivo General de esta Tesis Doctoral

fue el de investigar la asociación entre las alteraciones cerebrales estructurales y

funcionales presentes en individuos policonsumidores de drogas y (i) las variables de

personalidad que incrementan la predisposición al consumo y la dependencia, (ii) las

estimaciones de cantidad y duración de consumo de diferentes drogas, y (iii) el

funcionamiento neuropsicológico de los procesos ejecutivos típicamente afectados por el

consumo de drogas.

Este Objetivo General fue articulado en tres Objetivos Específicos que dieron lugar

a tres trabajos de investigación empírica y que han sido presentados en apartados

anteriores. Estos objetivos fueron: (i) estudiar la asociación entre las alteraciones

estructurales presentadas por un grupo de pacientes policonsumidores de drogas en

situación de abstinencia con consumo principal de cocaína y un conjunto de medidas de

personalidad, (ii) estudiar la asociación de la severidad del consumo de drogas con el

P á g i n a | 1 2 3

metabolismo cerebral en reposo de un grupo de pacientes policonsumidores de drogas en

situación de abstinencia y (iii) estudiar la asociación del funcionamiento neuropsicológico

y el metabolismo cerebral en situación de reposo en este mismo grupo de

policonsumidores de drogas.

Respecto al primer objetivo, nuestros resultados mostraron un patrón de

interacción diferente entre el volumen de materia gris regional del grupo de consumidores

y el grupo de individuos no consumidores y dos de las dimensiones de impulsividad

evaluadas: la urgencia negativa y la falta de premeditación. Estos resultados fueron

interpretados como el resultado de un patrón disfuncional en el grupo de consumidores

asociado con la alteración de los sistemas frontales y subcorticales propios de esta

población. Asimismo, se pudo demostrar la existencia de reducciones significativas en el

volumen regional de materia gris y materia blanca cerebral en estos consumidores, la

presencia de mayores puntuaciones en cuatro de las cinco dimensiones de personalidad

evaluadas y la asociación de los patrones de severidad de consumo de cocaína con índices

globales de materia gris cerebral.

El estudio de la asociación de la severidad del consumo de drogas con el

metabolismo cerebral permitió demostrar la existencia de asociaciones significativas entre

la cantidad de cocaína, alcohol y cannabis consumida y la duración del consumo de

heroína, alcohol y MDMA y el metabolismo regional cerebral de diferentes regiones

cerebrales. En concreto, se encontró que todas las drogas consumidas (excepto la cocaína)

mostraban una correlación negativa con el metabolismo de la corteza prefrontal

dorsolateral y que la severidad del consumo de alcohol y cocaína correlacionaba con el

metabolismo de la corteza promotora y el putamen y la corteza parietal respectivamente.

Estas asociaciones fueron interpretadas como el resultado de la existencia de mecanismos

de acción comunes y específicos de las distintas drogas sobre un conjunto de sistemas

neuroquímicos y estructuras cerebrales.

P á g i n a | 1 2 4

Por último, el estudio de la asociación del funcionamiento neuropsicológico y el

metabolismo cerebral permitió demostrar la existencia de asociaciones significativas entre

el rendimiento neuropsicológico de estos pacientes en tareas de funcionamiento ejecutivo

“frías” y “calientes” y un conjunto de regiones cerebrales. En este sentido, se encontró una

asociación significativa entre el rendimiento en las tareas de funcionamiento ejecutivo

“frías” y el metabolismo cerebral de la corteza prefrontal dorsolateral, la parte medial del

giro frontal superior, las partes superior e inferior de la corteza temporal y la corteza

parietal, y entre el rendimiento en las tareas de funcionamiento ejecutivo “calientes” y la

corteza prefrontal dorsolateral, la parte medial del giro frontal superior, la parte anterior y

posterior del cíngulo, la corteza temporal y el giro fusiforme. Estos resultados fueron

interpretados como el resultado de la diferente implicación de los sistemas cerebrales en

el funcionamiento ejecutivo y sus diferentes componentes. Asimismo, se pudo demostrar

la asociación entre el metabolismo cerebral y diferentes patrones de consumo y la

existencia de alteraciones neuropsicológicas en todos los procesos neuropsicológicos

evaluados al comparar la ejecución de nuestros pacientes con la de un grupo de individuos

no consumidores.

De estos resultados se derivan una serie de conclusiones y recomendaciones. En

primer lugar, la realización de estos tres estudios nos ha permitido profundizar en el

estudio de los factores explicativos de la asociación entre el consumo de drogas y las

alteraciones cerebrales y de funcionamiento neuropsicológico que se observan en esta

población. Los resultados encontrados demuestran que tanto las variables de

personalidad, en concreto las dimensiones de falta de premeditación y urgencia negativa,

como las variables relacionadas con el consumo de drogas se asocian con las alteraciones

encontradas en esta población. Así, la presencia de ciertos rasgos de personalidad como la

impulsividad, pueden predisponer al individuo vulnerable al inicio del consumo de drogas,

promover la transición del consumo recreativo a la dependencia, impedir que la persona

pueda abandonar el consumo a pesar de intentarlo en repetidas ocasiones y por último

P á g i n a | 1 2 5

provocar la recaída del consumidor abstinente (Verdejo-García et al., 2008), pero además,

el consumo prolongado de estas sustancias puede traer consigo la alteración del

funcionamiento cerebral y la exacerbación de la impulsividad resultado de la alteración de

las áreas asociadas con el control de la conducta (Ersche et al., 2010a).

La consideración de la impulsividad como un factor de riesgo para iniciarse en el

consumo de drogas ha sido avalada por diferentes tipos de estudios. En este sentido, se

han llevado a cabo investigaciones con diferentes poblaciones caracterizadas por

presentar un alto riesgo de iniciarse en el consumo de drogas (p.e. adolescentes, pacientes

con trastorno por déficits de atención e hiperactividad (TDAH) y otros trastornos

externalizantes o hijos de padres que han sido consumidores) o en jugadores patológicos,

un tipo de “adicción conductual” que comparte con las adicciones por sustancias

mecanismos etiológicos y de vulnerabilidad pero que se diferencia de estas en que en este

tipo de adicción no se produce la administración de una sustancia que puede causar

efectos dañinos a nivel cerebral (Bechara, 2003; Potenza, 2001) y se ha demostrado que

estas poblaciones presentan de forma consistente altas puntuaciones de impulsividad y

alteraciones en los procesos de control inhibitorio (Verdejo-García et al., 2008).

Sin embargo, tanto la impulsividad como la presencia de alteraciones

neuropsicológicas pueden incrementar como consecuencia directa del consumo de drogas.

Y es que, son múltiples los estudios que demuestran la existencia de alteraciones

significativas en los procesos de control inhibitorio y toma de decisiones en consumidores

de diferentes tipos de drogas y la asociación de estas con los diferentes patrones de

severidad del consumo evaluados.

Por otra parte, observamos que las diferentes sustancias evaluadas afectan de

forma generalizada a una serie de circuitos que han sido consistentemente asociados con

las alteraciones encontradas en esta población pero que también muestran cierta

especificidad. En este sentido, los resultados de éste y otros estudios parecen indicar que

P á g i n a | 1 2 6

múltiples sustancias producen deterioros globales similares sobre diversos mecanismos

de control ejecutivo pero que además, determinadas sustancias pueden producir

alteraciones más severas en componentes ejecutivos específicos (Fernández-Serrano et al.,

2011).

Es importante subrayar el hecho de que tanto el rendimiento en todas las pruebas

neuropsicológicas evaluadas como la severidad del consumo de todas las drogas

estudiadas (excepto de la cocaína), fueron asociadas con el metabolismo cerebral del área

prefrontal dorsolateral. La implicación de esta estructura en nuestra población coincide

con la evidencia que indica la asociación de esta estructura con múltiples procesos

ejecutivos y su alteración en consumidores de psicoestimulantes (Goldstein et al., 2004;

Makris et al., 2008b), opiáceos (Gerra et al., 1998), alcohol (Dao-Castellana et al., 1998;

Makris et al., 2008a) o cannabis (Eldreth et al., 2004). Es importante destacar la asociación

encontrada entre la cantidad de cannabis consumida y el metabolismo cerebral en esta

misma estructura. Y es que, hasta ahora, la mayoría de los estudios llevados a cabo con

esta población no han encontrado alteraciones significativas asociadas con el consumo de

cannabis en consumidores abstinentes de esta sustancia. Por lo que, estos resultados

subrayan la importancia de llevar a cabo nuevos estudios en esta población haciendo uso

de herramientas de neuroimagen cerebral, ya que estas han mostrado una mayor

sensibilidad en la detección de alteraciones en el funcionamiento cerebral aún en ausencia

de alteraciones a nivel conductual (p.e. Barrós-Loscertales et al., 2011a).

Asimismo, encontramos una asociación generalizada entre la severidad del

consumo de heroína, alcohol y MDMA y el metabolismo cerebral de la corteza temporal, un

área que ha sido encontrada alterada en los estudios llevados a cabo en estas poblaciones

y que coincide con las alteraciones en memoria encontradas en estos pacientes

(Fernández-Serrano et al., 2011).

P á g i n a | 1 2 7

Por último, es importante subrayar que los individuos drogodependientes

mostraron alteraciones en todos los procesos neuropsicológicos evaluados. La evidencia

indica que los consumidores de drogas presentan importantes déficits a nivel

neuropsicológico y que estos déficits son especialmente pronunciados en las llamadas

funciones ejecutivas (Fernández-Serrano et al., 2011). En este sentido, se ha demostrado

que los pacientes en tratamiento por drogodependencias con alteraciones en estos

procesos presentan un peor pronóstico y mayores tasas de recaídas que aquellos que no

presentan dichas alteraciones (Aharonovich et al., 2006; Grohman & Fals-Stewart, 1994;

Streeter et al., 2008). Es más, mientras que un adecuado funcionamiento ejecutivo ha sido

asociado con mayores probabilidades de mantenerse abstinente (Passetti et al., 2008), la

presencia de estas alteraciones se ha asociado con un menor nivel de implicación y

participación en los programas de tratamiento y con la severidad de los problemas

presentados por estos individuos en diferentes ámbitos de su funcionamiento cotidiano

(Kornreich et al., 2002; Verdejo-García, Bechara, Recknor, & Pérez-García, 2006) y ha

demostrado su validez en la predicción de recaídas de consumidores de diversas

sustancias (Aharonovich, Nunes, & Hasin, 2003; Passetti et al., 2008; Streeter et al., 2008;

Turner et al., 2009; Verdejo-García et al., 2012). Y es que, en comparación con otras

variables, las variables neuropsicológicas se asocian irremediablemente con las

neuroadaptaciones cerebrales que definen la naturaleza crónica de la adicción (Paulus et

al., 2005) y que se asocian tanto con la vulnerabilidad previa como con el daño provocado

por el consumo de las sustancias.

En este sentido, el conocimiento de los procesos neuropsicológicos afectados y sus

correspondientes correlatos neuroanatómicos puede llevarnos a desarrollar estrategias

más eficaces de tratamiento y rehabilitación, estrategias dirigidas al tratamiento de los

efectos generales y específicos asociados con cada tipo de droga y con las alteraciones en

el funcionamiento ejecutivo que sabemos que han sido asociadas con las recaídas. Un

tratamiento que podría verse favorecido por el uso combinado de estrategias dirigidas

P á g i n a | 1 2 8

específicamente al tratamiento de las funciones neuropsicológicas afectadas y de

tratamientos farmacológicos que hayan mostrado su validez en la mejora de ciertas

alteraciones cognitivas (Brady, Gray, & Tolliver, 2011; Sofuoglu, 2010) y que sin duda

alguna se verá favorecido por el uso de herramientas de neuroimagen cerebral.

2. Conclusiones

De los resultados obtenidos se derivan las siguientes conclusiones:

1. Los individuos policonsumidores de drogas con consumo principal de cocaína presentan

reducciones significativas del volumen regional de materia gris y materia blanca cerebral

en regiones fronto-límbicas y fronto-estriadas relacionadas con sus niveles de

impulsividad.

2. La asociación entre la impulsividad y el volumen regional de materia gris cerebral

muestra un patrón diferencial entre individuos policonsumidores de drogas con consumo

principal de cocaína y controles no consumidores: los niveles de impulsividad presentan

correlaciones positivas con el volumen de regiones relacionadas con la valoración de

reforzadores, una asociación que es inversa en el caso de los controles.

3. La severidad del consumo de diferentes drogas de abuso se asocia de forma general con

el metabolismo regional de la corteza prefrontal dorsolateral y la corteza temporal en

individuos policonsumidores de drogas.

4. La severidad del consumo de cocaína se asocia específicamente con el metabolismo

cerebral de la corteza parietal, la del alcohol con la corteza prefrontal y el putamen y la del

MDMA con el área postcentral en individuos policonsumidores de drogas.

5. Existen correlaciones significativas entre el rendimiento neuropsicológico de los

individuos policonsumidores de drogas en pruebas de funcionamiento ejecutivo y el

metabolismo cerebral de diversas regiones frontales y subcorticales.

P á g i n a | 1 2 9

6. Los individuos policonsumidores de drogas presentan deterioros equivalentes en las

funciones ejecutivas “frías” y “calientes” pero el rendimiento en las funciones ejecutivas

“frías” correlaciona específicamente con reducciones del metabolismo regional en áreas

frontales, temporales y parietales, mientras que el rendimiento en las funciones ejecutivas

“calientes” se correlaciona con el metabolismo regional de áreas frontales, temporales,

cíngulo y giro fusiforme.

3. Perspectivas de futuro

De los resultados obtenidos se derivan importantes perspectivas de investigación futura,

de entre las que cabe destacar:

1. Profundizar en la determinación de la dirección de causalidad de las alteraciones del

funcionamiento ejecutivo en individuos drogodependientes. Se hace necesario el diseño de

estudios longitudinales y en poblaciones de alto riesgo de iniciarse en el consumo de

drogas (por ejemplo en adolescentes, hijos de padres que han sido consumidores y niños y

adolescentes con trastornos de conducta o trastorno por déficit de atención e

hiperactividad), así como estudios transversales que incluyan variables genéticas,

neuropsicológicas y de neuroimagen cerebral.

2. Estudiar los efectos a largo plazo del consumo de diferentes sustancias utilizando

pruebas de funcionamiento neuropsicológico y de neuroimagen cerebral en consumidores

“puros” de las sustancias. En este sentido, aunque en nuestro estudio intentamos controlar

el efecto del co-abuso de las diferentes sustancias, el efecto del policonsumo no puede ser

completamente descartado (Halpern et al., 2004).

3. Profundizar en la investigación de los posibles efectos de programas de rehabilitación

específicos para cada tipo de consumidor.

P á g i n a | 1 3 0

4. Desarrollar paradigmas de neuroimagen funcional sensibles a la detección de

alteraciones en el funcionamiento cerebral de consumidores de cannabis en situación de

abstinencia.

5. Evaluar las alteraciones en los procesos emocionales de percepción y experiencia

emocional en consumidores de cannabis en situación de abstinencia.

6. Estudiar la validez de diferentes técnicas de neuroimagen cerebral en la predicción de

distintos índices de funcionamiento clínico, así como los potenciales beneficios de la

aplicación de estrategias de intervención sobre la rehabilitación de los individuos

drogodependientes.

7. Desarrollar programas de tratamiento dirigidos al reestablecimiento o la compensación

de las funciones neuropsicológicas y emocionales encontradas alteradas en los individuos

drogodependientes.

P á g i n a | 1 3 1

[DOCTORADO EUROPEO]

P á g i n a | 1 3 2

1. Summary

The review of the literature conducted at the beginning of our investigation

allowed us to emphasize the potential impact that premorbid personality traits, such as

impulsivity, may have on different stages of the addictive process and its neurobehavioral

correlates. We also observed a number of inconsistencies and limitations in the way that

polysubstance use and duration of abstinence was handled in many of the studies

conducted in this population (poor control of co-abuse of other drugs and predominance

of studies in acute or shortly abstinent drug users). In this study we attempted to address

some of these limitations by studying the association between brain structure and

function and impulsivity traits, severity of use of different drugs, and executive control

associated dysfunction in polysubstance dependent individuals during protracted

abstinence.

The main aim of this Thesis was therefore to investigate the association between

brain structure and function in substance dependent individuals and (i) trait impulsivity,

as a main risk factor and clinical symptom associated with addiction, (ii) lifetime estimates

of amount and duration of drugs used and (iii) performance on tests of executive

functioning, which are thought to be associated with drug-induced neuroadaptations.

This aim was subsequently decomposed in three specific aims, which correspond

to the three studies collected in this Thesis. The first aim was to investigate the association

between volumetric measures of gray and white matter (measured with Magnetic

Resonance Imaging) and trait impulsivity and estimates of cocaine use in cocaine-

dependent individuals. Here we predicted that both greater impulsivity and higher and

longer exposure to cocaine use would correlate with less gray and white matter in

prefrontal and subcortical structures. Our results showed a different pattern of interaction

between regional gray matter volume of the cocaine-dependent individuals and the

control group and two of the dimensions of impulsivity assessed: negative urgency and

P á g i n a | 1 3 3

lack of premeditation. These results were interpreted as a result of a dysfunctional pattern

in the cocaine-dependent group associated with the disruption of frontal and subcortical

systems in this population. Also, we found significant reductions in the volume of regional

gray and white matter in the cocaine-dependent individuals, the presence of higher scores

on four of the five personality dimensions assessed and the association of patterns of

severity of cocaine use and overall rates of cerebral gray matter. The results of this study

can be found in the article entitled “Trait impulsivity and prefrontal gray matter

reductions in cocaine dependent individuals” published in Drug and Alcohol Dependence.

The second aim was to investigate the association between brain regional

metabolism (measured with FDG-PET) and lifetime estimates of amount and duration of

cocaine, heroin, alcohol, MDMA and cannabis use in a sample of polysubstance dependent

individuals. Here we predicted that exposure to all substances would be significantly

associated with overlapping reductions of metabolism in frontal, limbic, striatal and

cerebellar regions, whereas cocaine and heroin use would specifically correlate with

parietal and temporal cortices respectively. We found that whereas all drugs consumed

(except cocaine) showed a negative correlation with the metabolism of the dorsolateral

prefrontal cortex (DLPFC), the severity of alcohol, cocaine and MDMA correlated with

regional metabolism in premotor cortex and putamen, parietal cortex and postcentral

gyrus respectively. These associations were interpreted as a result of the existence of

common and specific mechanisms of action of the different drugs studied on the

neurotransmitter and functional brain systems involved in addiction. The results of this

study can be found in the article entitled “Neural Correlates of the Severity of Cocaine,

Heroin, Alcohol, MDMA and Cannabis Use in Polysubstance Abusers: A Resting-PET Brain

Metabolism Study” published in PloS One.

Finally, the third aim was to investigate the association between brain regional

metabolism (measured with FDG-PET) and performance measures of executive

P á g i n a | 1 3 4

functioning in a sample of polysubstance dependent individuals. Here we predicted that

executive performance would be poorer in polysubstance users, and that it would be

associated with lateral prefrontal and temporal regions for tests of “cold” executive

functions, and with orbitofrontal cortex and limbic regions for tests of “hot” executive

function. Regression models showed that substance dependent individuals performance in

cold executive test was associated with regional brain metabolism in the DLPFC, mid-

superior frontal gyrus, superior and inferior temporal gyrus and inferior parietal cortex,

whereas performance in hot executive functioning was associated with DLPFC, mid-

superior frontal gyrus, anterior and mid-posterior cingulate, and temporal and fusiform

gyrus. These results are in agreement with the notion that different executive processes

rely in both common and partially dissociable neural networks (F Collette et al., 2006),

and that different neural circuits relevant to various aspect of abnormal cognitive

functioning in substance dependent individuals (Verdejo-García et al., 2004). Likewise, we

found an association between cerebral metabolism and the different pattern of use

showed by the substance dependent individuals and the existence of neuropsychological

impairments in all the processes evaluated. Our results can be found in the article entitled

“Neural correlates of hot and cold executive functions in polysubstance addiction:

association between neuropsychological performance and resting-PET brain metabolism”

published in Psychiatry Research: Neuroimaging.

From these results we can derive a set of conclusions and recommendations. First

of all, the implementation of these three studies has provided further evidence of the

factors that contribute to explain the association between drug use and brain functioning.

The results found indicate that both impulsivity and drug use are associated with brain

alterations in this population. Therefore, some aspects of substance dependence and its

related neuroadaptations may be associated with personality traits. During substance

dependence, drug users may perseverate in drug-taking behaviours and the prolonged use

P á g i n a | 1 3 5

of these substances can lead to different neuroadaptations or to the exacerbation of the

ones originally associated with impulse control (Ersche et al., 2010a).

Moreover, we found significant general and specific associations of the severity of

drug use and the regional brain metabolism in the substance dependence individuals.

These associations can be interpreted as a result of the existence of common and specific

mechanisms of action of the drug studied on neurotransmitter and brain systems involved

in the addiction cycle. Studies exploring the effects of specific rehabilitation programs

based on the common and specific effects of each type of substance used can be very

useful to improve novel tailored interventions. Likewise, it is important to underscore that

both the deficient performance on the neuropsychological tests of executive function and

the severity of use of all drugs studied (except cocaine) were associated with the

functioning of the DLPFC. The relevance of this structure for substance dependence

overlaps with its relevance for the functioning of multiple executive processes. Therefore,

our results are in agreement with those of previous studies showing significant alterations

of this region in users of psychostimulants (Goldstein et al., 2004; Makris et al., 2008b),

opiates (Gerra et al., 1998), alcohol (Dao-Castellana et al., 1998; Makris et al., 2008a) or

cannabis (Eldreth et al., 2004). Importantly, the amount of cannabis used also correlated

with the regional brain metabolism in this structure, pointing to a link between use of this

substance and neuroadaptations in this region even after long-term abstinence. These

results underscore the importance of conducting further studies in cannabis users

employing brain imaging tools, as these have shown a greater sensitivity in detecting

alterations in brain functioning even in the absence of behavioral alterations (e.g.

Loscertales Barros et al., 2011).

Finally, it is important to note that substance dependent individuals showed

alterations in all the neuropsychological processes assessed. These results agree with

those of previous studies that had detected generalized alterations in the executive

P á g i n a | 1 3 6

functions in users of several substances (Di Sclafani, Tolou-Shams, Price, & Fein, 2002;

Goldstein et al., 2004; Noël et al., 2001). Furthermore, these results may have important

clinical implications for the treatment and rehabilitation of substance dependent

individuals. First, previous studies have shown that neuropsychological alterations of

executive functions are associated with reduced level of involvement and participation of

substance dependent individuals in treatment programs, and with a higher rate of

dropping out of these programs (Aharonovich et al., 2006; Passetti et al., 2008; Streeter et

al., 2008; Turner et al., 2009). On the other hand, these alterations are also related to the

severity of the problems showed by substance dependent individuals in diverse aspects of

daily functioning, including employment, social, family and legal problems (Kornreich et

al., 2002; Verdejo-García et al., 2006). Therefore, it is possible that the clinical and daily

functioning of the substance dependent individuals would benefit from the application of

specific neuropsychological training and rehabilitation strategies.

From the results obtained, we can derive the following conclusions and future

perspectives:

2. Conclusions

1. Cocaine dependent individuals have lowered gray and white matter volume in

fronto-limbic and fronto-striatal regions related to their levels of impulsivity.

2. The association between impulsivity and regional gray matter volumes shows a

differential pattern in cocaine dependent individuals vs. healthy controls:

impulsivity levels have a positive correlation with the volumes of regions

associated with reinforcement valuation, whereas this association is negative in

the healthy control group.

3. The severity of use of different drugs is associated with overlapping prefrontal and

temporal regions metabolic reductions in substance dependent individuals.

P á g i n a | 1 3 7

4. The severity of cocaine use is specifically negatively correlated with brain metabolism in

the parietal cortex, the severity of alcohol with the premotor cortex and the putamen, and

the severity of MDMA with the postcentral gyrus.

5. There are significant correlations between substance dependent individuals’

neuropsychological performance and the brain metabolism of several frontal and

subcortical regions.

6. Substance dependent individuals’ performance in cold executive test is associated with

regional brain metabolism in the DLPFC, mid-superior frontal gyrus, superior and inferior

temporal gyrus and inferior parietal cortex, whereas performance in hot executive

functioning is associated with DLPFC, mid-superior frontal gyrus, anterior and mid-

posterior cingulate, and temporal and fusiform gyrus.

3. Future perspectives

1. To examine the direction of causality of the link between executive deficits and drug

addiction. One possible approach to this issue is to perform longitudinal studies able to

examine possible executive alterations in populations at high risk of initiating drug use

(for example: adolescents, children and adolescents with ADHD or conduct disorder, etc.).

Another possible approach is to explore the relationship between mechanisms of genetic

vulnerability and neuropsychological alterations in substance dependent individuals.

2. To examine the specific brain systems involved in the abnormal functioning of executive

and emotional processes in “pure” substance dependent individuals. In this respect,

although in our study we try to control the effect of co-abuse of the different substances

used, the effect of polydrug use can not be completely ruled out (Halpern et al., 2004).

3. To investigate the effects of specific rehabilitation programs for each type of substance

dependent individual.

P á g i n a | 1 3 8

3. To develop neuroimaging paradigms sensitive to detect alterations in the brain

functioning of cannabis dependent individuals after prolonged abstinence.

4. To study the alteration of the emotional processing in cannabis dependent individuals

after prolonged abstinence.

5. To study the validity of different brain imaging techniques for the prediction of different

clinical outcomes and for prediction of relapse.

6. To develop treatment programs aimed at the restoration or compensation of the

neuropsychological functions found altered in these populations.

P á g i n a | 1 3 9

[REFERENCIAS]

P á g i n a | 1 4 0

Adams, K. M., Gilman, S., Koeppe, R. A., Kluin, K. J., Brunberg, J. A., Dede, D., Berent, S., et al.

(1993). Neuropsychological deficits are correlated with frontal hypometabolism in

positron emission tomography studies of older alcoholic patients. Alcoholism, clinical

and experimental research, 17, 205-210.

Adinoff, B., Devous, M. D., Cooper, D. B., Best, S. E., Chandler, P., Harris, T., Cervin, C. A., et al.

(2003). Resting regional cerebral blood flow and gambling task performance in

cocaine-dependent subjects and healthy comparison subjects. The American journal

of psychiatry, 160, 1892-1894.

Aguilar de Arcos, F., Verdejo-Garcia, A., Lopez Jimenez, A., Montañez Pareja, M., Gomez

Juarez, E., Arraez Sanchez, F., & Perez-Garcia, M. (2008). Cambios en la respuesta

emocional ante estímulos visuales de contenido sexual en adictos a drogas.

Adicciones, 20, 117-124.

Aguilar de Arcos, F., Verdejo-García, A., Peralta-Ramírez, M. I., Sánchez-Barrera, M., &

Pérez-García, M. (2005). Experience of emotions in substance abusers exposed to

images containing neutral, positive, and negative affective stimuli. Drug and alcohol

dependence, 78, 159-167.

Aharonovich, E., Amrhein, P. C., Bisaga, A., Nunes, E. V., & Hasin, D. S. (2008). Cognition,

Commitment Language, and Behavioral Change Among Cocaine-Dependent Patients.

Psychology of Addictive Behaviors, 22, 557-562.

Aharonovich, E., Hasin, D. S., Brooks, A. C., Liu, X., Bisaga, A., & Nunes, E. V. (2006).

Cognitive deficits predict low treatment retention in cocaine dependent patients.

Drug and alcohol dependence, 81, 313-322.

Aharonovich, E., Nunes, E., & Hasin, D. (2003). Cognitive impairment, retention and

abstinence among cocaine abusers in cognitive-behavioral treatment. Drug and

alcohol dependence, 71, 207-211.

Albein-Urios, N., Martinez-González, J. M., Lozano, O., Clark, L., & Verdejo-García, A. (2012).

Comparison of impulsivity and working memory in cocaine addiction and

pathological gambling: Implications for cocaine-induced neurotoxicity. Drug and

alcohol dependence, en prensa.

P á g i n a | 1 4 1

Alia-Klein, N., Parvaz, M. A., Woicik, P. A., Konova, A. B., Maloney, T., Shumay, E., Wang, R.,

et al. (2011). Gene x disease interaction on orbitofrontal gray matter in cocaine

addiction. Archives of general psychiatry, 68, 283-294.

Anderson, S. W., Barrash, J., Bechara, A., & Tranel, D. (2006). Impairments of emotion and

real-world complex behavior following childhood- or adult-onset damage to

ventromedial prefrontal cortex. Journal of the International Neuropsychological

Society, 12, 224-235.

Aron, A. R. (2007). The neural basis of inhibition in cognitive control. The Neuroscientist,

13, 214-228.

Aron, J. L., & Paulus, M. P. (2007). Location, location: using functional magnetic resonance

imaging to pinpoint brain differences relevant to stimulant use. Addiction, 102, 33-43.

Asensio, S., Romero, M. J., Palau, C., Sanchez, A., Senabre, I., Morales, J. L., Carcelen, R., et al.

(2010). Altered neural response of the appetitive emotional system in cocaine

addiction: an fMRI Study. Addiction biology, 15, 504-516.

Ashburner, J., & Friston, K. J. (2005). Unified segmentation. NeuroImage, 26, 839-851.

Brand, M., Rothbauer, M., Driessen, M., & Markowitsch, H. (2008). Executive functions and

risky decision-making in patients with opiate dependence. Drug and Alcohol

Dependence, 97, 64-72.

Baler, R. D., & Volkow, N. D. (2006). Drug addiction: the neurobiology of disrupted self-

control. Trends in molecular medicine, 12, 559-566.

Barratt, E. S. (1959). Anxiety and Impulsiveness Related to Psychomotor Efficiency.

Perceptual and Motor Skills, 9, 191-198.

Barrós-Loscertales, A., Bustamante, J. C., Ventura-Campos, N., Llopis, J. J., Parcet, M. A., &

Avila, C. (2011a). Lower activation in the right frontoparietal network during a

counting Stroop task in a cocaine-dependent group. Psychiatry research, 194, 111-

118.

Barrós-Loscertales, A., Garavan, H., Bustamante, J. C., Ventura-Campos, N., Llopis, J. J.,

Belloch, V., Parcet, M. A., et al. (2011b). Reduced striatal volume in cocaine-dependent

patients. NeuroImage, 56, 1021-1026.

P á g i n a | 1 4 2

Battisti, R. A., Roodenrys, S., Johnstone, S. J., Pesa, N., Hermens, D. F., & Solowij, N. (2010).

Chronic cannabis users show altered neurophysiological functioning on Stroop task

conflict resolution. Psychopharmacology, 212, 613-624.

Baumann, M. H., Wang, X., & Rothman, R. B. (2007). 3,4-Methylenedioxymethamphetamine

(MDMA) neurotoxicity in rats: a reappraisal of past and present findings.

Psychopharmacology, 189, 407-424.

Bechara, A, Damasio, A. R., Damasio, H., & Anderson, S. W. (1994). Insensitivity to future

consequences following damage to human prefrontal cortex. Cognition, 50, 7-15.

Bechara, A, Damasio, H., & Damasio, A. R. (2000). Emotion, decision making and the

orbitofrontal cortex. Cerebral cortex, 10, 295-307.

Bechara, A. (2003). Risky business: emotion, decision-making, and addiction. Journal of

gambling studies, 19, 23-51.

Bechara, A. (2005). Decision making, impulse control and loss of willpower to resist drugs:

a neurocognitive perspective. Nature neuroscience, 8, 1458-1463.

Bechara, A, Dolan, S., Denburg, N., Hindes, A., Anderson, S. W., & Nathan, P. E. (2001).

Decision-making deficits, linked to a dysfunctional ventromedial prefrontal cortex,

revealed in alcohol and stimulant abusers. Neuropsychologia, 39, 376-389.

Bechara, A., & Van Der Linden, M. (2005). Decision-making and impulse control after

frontal lobe injuries. Current opinion in neurology, 18, 734-739.

Belin, D., Mar, A. C., Dalley, J. W., Robbins, T. W., & Everitt, B. J. (2008). High impulsivity

predicts the switch to compulsive cocaine-taking. Science, 320, 1352-1355.

Bermpohl, F., Kahnt, T., Dalanay, U., Hägele, C., Sajonz, B., Wegner, T., Stoy, M., et al. (2010).

Altered representation of expected value in the orbitofrontal cortex in mania. Human

brain mapping, 31, 958-969.

Bickel, W. K., & Marsch, L. A. (2001). Toward a behavioral economic understanding of drug

dependence: delay discounting processes. Addiction, 96, 73-86.

P á g i n a | 1 4 3

Bjork, J. M., Hommer, D. W., Grant, S. J., & Danube, C. (2004). Impulsivity in abstinent

alcohol-dependent patients: relation to control subjects and type 1-/type 2-like traits.

Alcohol, 34, 133-150.

Blair, R. J. R. (2003). Facial expressions, their communicatory functions and neuro-

cognitive substrates. Philosophical transactions of the Royal Society of London. Series

B, Biological sciences, 358, 561-572.

Boellaard, R., O’Doherty, M. J., Weber, W. A., Mottaghy, F. M., Lonsdale, M. N., Stroobants, S.

G., Oyen, W. J. G., et al. (2010). FDG PET and PET/CT: EANM procedure guidelines for

tumour PET imaging: version 1.0. European journal of nuclear medicine and molecular

imaging, 37, 181-200.

Bolla, K. I., Eldreth, D. A., London, E. D., Kiehl, K. A., Mouratidis, M., Contoreggi, C., Matochik,

J. A., et al. (2003). Orbitofrontal cortex dysfunction in abstinent cocaine abusers

performing a decision-making task. NeuroImage, 19, 1085-1094.

Bolla, K., Ernst, M., Kiehl, K., Mouratidis, M., Eldreth, D., Contoreggi, C., Matochik, J., et al.

(2004). Prefrontal cortical dysfunction in abstinent cocaine abusers. The Journal of

neuropsychiatry and clinical neurosciences, 16, 456-464.

Bolla, K. I., Funderburk, F. R., & Cadet, J. L. (2000). Differential effects of cocaine and

cocaine alcohol on neurocognitive performance. Neurology, 54, 2285-2292.

Bolla, K. I., Eldreth, D. A., Matochik, J. A., & Cadet, J. L. (2005). Neural substrates of faulty

decision-making in abstinent marijuana users. NeuroImage, 26, 480-492.

Bowden-Jones, H., McPhillips, M., Rogers, R., Hutton, S., & Joyce, E. (2005). Risk-taking on

tests sensitive to ventromedial prefrontal cortex dysfunction predicts early relapse in

alcohol dependency: a pilot study. The Journal of neuropsychiatry and clinical

neurosciences, 17, 417-420.

Brady, K. T., Gray, K. M., & Tolliver, B. K. (2011). Cognitive enhancers in the treatment of

substance use disorders: clinical evidence. Pharmacology, biochemistry, and behavior,

99, 285-294.

Brand, M., Labudda, K., & Markowitsch, H. J. (2006). Neuropsychological correlates of

decision-making in ambiguous and risky situations. Neural networks, 19, 1266-1276.

P á g i n a | 1 4 4

Brett, M., Anton, J. L., Valabregue, R., & Poline, J. B. (2002). Region of interest analysis using

an SPM toolbox. NeuroImage, 16, 2.

Breiter, H. C., & Rosen, B. R. (1999). Functional magnetic resonance imaging of brain

reward circuitry in the human. Annals of the New York Academy of Sciences, 877, 523-

547.

Brooks, J. O., Rosen, A. C., Hoblyn, J. C., Woodard, S. A., Krasnykh, O., & Ketter, T. A. (2009).

Resting prefrontal hypometabolism and paralimbic hypermetabolism related to

verbal recall deficits in euthymic older adults with bipolar disorder. The American

journal of geriatric psychiatry, 17, 1022-1029.

Buchert, R., Obrocki, J., Thomasius, R., Väterlein, O., Petersen, K., Jenicke, L., Bohuslavizki,

K. H., et al. (2001). Long-term effects of “ecstasy” abuse on the human brain studied

by FDG PET. Nuclear medicine communications, 22, 889-897.

Buckholtz, J. W., Treadway, M. T., Cowan, R. L., Woodward, N. D., Li, R., Ansari, M. S.,

Baldwin, R. M., et al. (2010). Dopaminergic network differences in human impulsivity.

Science, 329, 532.

Burgess, P. W., Veitch, E., de Lacy Costello, A., & Shallice, T. (2000). The cognitive and

neuroanatomical correlates of multitasking. Neuropsychologia, 38, 848-863.

Bustamante, J. C., Barrós-Loscertales, A., Ventura-Campos, N., Sanjuán, A., Llopis, J. J.,

Parcet, M.-A., & Avila, C. (2011). Right parietal hypoactivation in a cocaine-dependent

group during a verbal working memory task. Brain research, 1375, 111-119.

Cairns, E. D., & Cammock, J. (2002). Test de emparejamiento de figuras conocidas: MFF-20.

Madrid: TEA Ediciones.

Calder, A. J., & Young, A. W. (2005). Understanding the recognition of facial identity and

facial expression. Nature reviews. Neuroscience, 6, 641-651.

Camí, J., & Farré, M. (2003). Drug addiction. The New England journal of medicine, 349, 975-

986.

Chanraud, S., Martelli, C., Delain, F., Kostogianni, N., Douaud, G., Aubin, H. J., Reynaud, M., et

al. (2007). Brain morphometry and cognitive performance in detoxified alcohol-

P á g i n a | 1 4 5

dependents with preserved psychosocial functioning. Neuropsychopharmacology, 32,

429-438.

Childress, A. R., Mozley, P. D., McElgin, W., Fitzgerald, J., Reivich, M., & O’Brien, C. P. (1999).

Limbic activation during cue-induced cocaine craving. The American journal of

psychiatry, 156, 11-18.

Childress, A. R., Ehrman, R. N., Wang, Z., Li, Y., Sciortino, N., Hakun, J., Jens, W., et al. (2008).

Prelude to passion: limbic activation by “unseen” drug and sexual cues. PloS one, 3,

e1506.

Clark, L., Robbins, T. W., Ersche, K. D., & Sahakian, B. J. (2006). Reflection impulsivity in

current and former substance users. Biological psychiatry, 60, 515-522.

Coffey, S. F., Gudleski, G. D., Saladin, M. E., & Brady, K. T. (2003). Impulsivity and rapid

discounting of delayed hypothetical rewards in cocaine-dependent individuals.

Experimental and clinical psychopharmacology, 11, 18-25.

Cohen, L. J., Gertmenian-King, E., Kunik, L., Weaver, C., London, E. D., & Galynker, I. (2005).

Personality measures in former heroin users receiving methadone or in protracted

abstinence from opiates. Acta psychiatrica Scandinavica, 112, 149-158.

Collette, F., Hogge, M., Salmon, E., & Van der Linden, M. (2006). Exploration of the neural

substrates of executive functioning by functional neuroimaging. Neuroscience, 139,

209-221.

Collette, F., & Van der Linden, M. (2002). Brain imaging of the central executive component

of working memory. Neuroscience and biobehavioral reviews, 26, 105-125.

Cowan, R. L., Lyoo, I. K., Sung, S. M., Ahn, K. H., Kim, M. J., Hwang, J., Haga, E., et al. (2003).

Reduced cortical gray matter density in human MDMA (Ecstasy) users: a voxel-based

morphometry study. Drug and alcohol dependence, 72, 225-235.

Crean, R. D., Crane, N. A., & Mason, B. J. (2011). An evidence based review of acute and

long-term effects of cannabis use on executive cognitive functions. Journal of

addiction medicine, 5, 1-8.

Cyders, M. A., Flory, K., Rainer, S., & Smith, G. T. (2009). The role of personality dispositions

to risky behavior in predicting first-year college drinking. Addiction, 104, 193-202.

P á g i n a | 1 4 6

Cyders, M. A., Smith, G. T., Spillane, N. S., Fischer, S., Annus, A. M., & Peterson, C. (2007).

Integration of impulsivity and positive mood to predict risky behavior: development

and validation of a measure of positive urgency. Psychological assessment, 19, 107-

118.

Dalley, J. W., Fryer, T. D., Brichard, L., Robinson, E. S. J., Theobald, D. E. H., Lääne, K., Peña,

Y., et al. (2007). Nucleus accumbens D2/3 receptors predict trait impulsivity and

cocaine reinforcement. Science, 315, 1267-1270.

Dao-Castellana, M. H., Samson, Y., Legault, F., Martinot, J. L., Aubin, H. J., Crouzel, C.,

Feldman, L., et al. (1998). Frontal dysfunction in neurologically normal chronic

alcoholic subjects: metabolic and neuropsychological findings. Psychological

medicine, 28, 1039-1048.

Davidson, R. J. (2002). Anxiety and affective style: Role of prefrontal cortex and amygdale.

Biological Psychiatry, 51, 68-80.

Davies, S. J. C., Pandit, S. A., Feeney, A., Stevenson, B. J., Kerwin, R. W., Nutt, D. J., Marshall, E.

J., et al. (2005). Is there cognitive impairment in clinically “healthy” abstinent alcohol

dependence? Alcohol and alcoholism, 40, 498-503.

Desmond, J. E., Chen, S. H. A., DeRosa, E., Pryor, M. R., Pfefferbaum, A., & Sullivan, E. V.

(2003). Increased frontocerebellar activation in alcoholics during verbal working

memory: an fMRI study. NeuroImage, 19, 1510-1520.

del Ser, T., García-Montalvo, J. I., Martínez Espinosa, S., Delgado Villapalos, C., & Bermejo, F.

(1997). The estimation of premorbid intelligence in Spanish people with the "Word

Accentuation Test" and its application to the diagnosis of dementia. Brain &

Cognition, 33, 343-356.

DeFilippis, N. A. (2002). Category Test: Computer Version Research Edition. Lutz, Florida:

Psychological Assessment Resources.

de Win, M. M. L., Jager, G., Booij, J., Reneman, L., Schilt, T., Lavini, C., Olabarriaga, S. D., et al.

(2008). Neurotoxic effects of ecstasy on the thalamus. The British journal of

psychiatry, 193, 289-296.

de Sola, S., Tarancón, T., Peña-Casanova, J., Espadaler, J. M., Langohr, K., Poudevida, S.,

Farré, M., et al. (2008). Auditory event-related potentials (P3) and cognitive

P á g i n a | 1 4 7

performance in recreational ecstasy polydrug users: evidence from a 12-month

longitudinal study. Psychopharmacology, 200, 425-437.

de Oliveira, L. G., Barroso, L. P., Silveira, C. M., Sanchez, Z. V. D. M., De Carvalho Ponce, J.,

Vaz, L. J., & Nappo, S. A. (2009). Neuropsychological assessment of current and past

crack cocaine users. Substance use & misuse, 44, 1941-1957.

Di Sclafani, V., Tolou-Shams, M., Price, L. J., & Fein, G. (2002). Neuropsychological

performance of individuals dependent on crack-cocaine, or crack-cocaine and

alcohol, at 6 weeks and 6 months of abstinence. Drug and alcohol dependence, 66,

161-171.

Dickman, S. J. (1990). Functional and dysfunctional impulsivity: personality and cognitive

correlates. Journal of personality and social psychology, 58, 95-102.

Dinkelacker, V., Grüter, M., Klaver, P., Grüter, T., Specht, K., Weis, S., Kennerknecht, I., et al.

(2011). Congenital prosopagnosia: multistage anatomical and functional deficits in

face processing circuitry. Journal of neurology, 258, 770-782.

Dom, G., De Wilde, B., Hulstijn, W., van den Brink, W., & Sabbe, B. (2006). Decision-making

deficits in alcohol-dependent patients with and without comorbid personality

disorder. Alcoholism, clinical and experimental research, 30, 1670-1677.

D’Esposito, M., Detre, J. A., Alsop, D. C., Shin, R. K., Atlas, S., & Grossman, M. (1995). The

neural basis of the central executive system of working memory. Nature, 378, 279-

281.

D’Esposito, M. (2007). From cognitive to neural models of working memory. Philosophical

transactions of the Royal Society of London. Series B, Biological sciences, 362, 761-772.

D’Souza, M. S., & Duvauchelle, C. L. (2008). Certain or uncertain cocaine expectations

influence accumbens dopamine responses to self-administered cocaine and non-

rewarded operant behavior. European neuropsychopharmacology, 18, 628-638.

Eckardt, M. J., Rohrbaugh, J. W., Rio, D. E., & Martin, P. R. (1990). Positron emission

tomography as a technique for studying the chronic effects of alcohol on the human

brain. Annals of medicine, 22, 341-345.

P á g i n a | 1 4 8

Eldreth, D. A., Matochik, J. A., Cadet, J. L., & Bolla, K. I. (2004). Abnormal brain activity in

prefrontal brain regions in abstinent marijuana users. NeuroImage, 23, 914-920.

Ersche, K. D., Barnes, A., Jones, P. S., Morein-Zamir, S., Robbins, T. W., & Bullmore, E. T.

(2011). Abnormal structure of frontostriatal brain systems is associated with aspects

of impulsivity and compulsivity in cocaine dependence. Brain, 134, 2013-2024.

Ersche, K. D., Bullmore, E. T., Craig, K. J., Shabbir, S. S., Abbott, S., Müller, U., Ooi, C., et al.

(2010b). Influence of compulsivity of drug abuse on dopaminergic modulation of

attentional bias in stimulant dependence. Archives of general psychiatry, 67, 632-644.

Ersche, K. D., Roiser, J. P., Robbins, T. W., & Sahakian, B. J. (2008). Chronic cocaine but not

chronic amphetamine use is associated with perseverative responding in humans.

Psychopharmacology, 197, 421-431.

Ersche, K. D., Turton, A. J., Pradhan, S., Bullmore, E. T., & Robbins, T. W. (2010a). Drug

addiction endophenotypes: impulsive versus sensation-seeking personality traits.

Biological psychiatry, 68, 770-773.

European Monitoring Centre for Drugs and Drug Addiction (2011). Annual report 2011:

The state of the drugs problem in Europe. Descargado de:

http://www.emcdda.europa.eu/publications/annual-report/2011 el 3 de Mayo de

2012.

Evenden, J. L. (1999). Varieties of impulsivity. Psychopharmacology, 146, 348-361.

Fehr, C., Hohmann, N., Gründer, G., Dielentheis, T. F., Buchholz, H. G., Chechko, N.,

Yakushev, I., et al. (2007). Tiagabine does not attenuate alcohol-induced activation of

the human reward system. Psychopharmacology, 191, 975-983.

Feil, J., Sheppard, D., Fitzgerald, P. B., Yücel, M., Lubman, D. I., & Bradshaw, J. L. (2010).

Addiction, compulsive drug seeking, and the role of frontostriatal mechanisms in

regulating inhibitory control. Neuroscience and biobehavioral reviews, 35, 248-275.

Fein, G., Klein, L., & Finn, P. (2004). Impairment on a simulated gambling task in long-term

abstinent alcoholics. Alcoholism, clinical and experimental research, 28, 1487-1491.

P á g i n a | 1 4 9

Fernández-Serrano, M. J., Perales, J. C., Moreno-López, L., Pérez-García, M., & Verdejo-

García, A. (2012). Neuropsychological profiling of impulsivity and compulsivity in

cocaine dependent individuals. Psychopharmacology, 219, 673-683.

Fernández-Serrano, M. J., Pérez-García, M., Perales, J. C., & Verdejo-García, A. (2010a).

Prevalence of executive dysfunction in cocaine, heroin and alcohol users enrolled in

therapeutic communities. European journal of pharmacology, 626, 104-112.

Fernández-Serrano, M. J., Pérez-García, M., Schmidt Río-Valle, J., & Verdejo-García, A.

(2010b). Neuropsychological consequences of alcohol and drug abuse on different

components of executive functions. Journal of psychopharmacology, 24, 1317-1332.

Fernández-Serrano, M. J., Pérez-García, M., & Verdejo-García, A. (2011). What are the

specific vs. generalized effects of drugs of abuse on neuropsychological performance?

Neuroscience and biobehavioral reviews, 35, 377-406.

Fernández-Serrano, M. J., Lozano, O., Pérez-García, M., & Verdejo-García, A. (2010c). Impact

of severity of drug use on discrete emotions recognition in polysubstance abusers.

Drug and alcohol dependence, 109, 57-64.

Fillmore, M. T., & Rush, C. R. (2002). Impaired inhibitory control of behavior in chronic

cocaine users. Drug and alcohol dependence, 66, 265-273.

Fishbein, D. H., Krupitsky, E., Flannery, B. A., Langevin, D. J., Bobashev, G., Verbitskaya, E.,

Augustine, C. B., et al. (2007). Neurocognitive characterizations of Russian heroin

addicts without a significant history of other drug use. Drug and alcohol dependence,

90, 25-38.

Fisk, J. E., & Sharp, C. A. (2004). Age-related impairment in executive functioning: updating,

inhibition, shifting, and access. Journal of clinical and experimental neuropsychology,

26, 874-890.

Foisy, M. L., Kornreich, C., Fobe, A., D’Hondt, L., Pelc, I., Hanak, C., Verbanck, P., et al. (2007).

Impaired emotional facial expression recognition in alcohol dependence: do these

deficits persist with midterm abstinence? Alcoholism, clinical and experimental

research, 31, 404-410.

P á g i n a | 1 5 0

Fontes, M. A., Bolla, K. I., Cunha, P. J., Almeida, P. P., Jungerman, F., Laranjeira, R. R.,

Bressan, R. A., et al. (2011). Cannabis use before age 15 and subsequent executive

functioning. The British journal of psychiatry, 198, 442-447.

Forman, S. D., Dougherty, G. G., Casey, B. J., Siegle, G. J., Braver, T. S., Barch, D. M., Stenger, V.

A., et al. (2004). Opiate addicts lack error-dependent activation of rostral anterior

cingulate. Biological psychiatry, 55, 531-537.

Fox, H. C., Jackson, E. D., & Sinha, R. (2009). Elevated cortisol and learning and memory

deficits in cocaine dependent individuals: relationship to relapse outcomes.

Psychoneuroendocrinology, 34, 1198-1207.

Franklin, T. R., Acton, P. D., Maldjian, J. A., Gray, J. D., Croft, J. R., Dackis, C. A., O’Brien, C. P.,

et al. (2002). Decreased gray matter concentration in the insular, orbitofrontal,

cingulate, and temporal cortices of cocaine patients. Biological psychiatry, 51, 134-

142.

Frigerio, E., Burt, D. M., Montagne, B., Murray, L. K., & Perrett, D. I. (2002). Facial affect

perception in alcoholics. Psychiatry research, 113, 161-171.

Galynker, I. I., Watras-Ganz, S., Miner, C., Rosenthal, R. N., Des Jarlais, D. C., Richman, B. L., &

London, E. (2000). Cerebral metabolism in opiate-dependent subjects: effects of

methadone maintenance. The Mount Sinai journal of medicine, 67, 381-387.

Galynker, I. I., Eisenberg, D., Matochik, J. A., Gertmenian-King, E., Cohen, L., Kimes, A. S.,

Contoreggi, C., et al. (2007). Cerebral metabolism and mood in remitted opiate

dependence. Drug and alcohol dependence, 90, 166-174.

Garavan, H., Pankiewicz, J., Bloom, A., Cho, J. K., Sperry, L., Ross, T. J., Salmeron, B. J., et al.

(2000). Cue-induced cocaine craving: neuroanatomical specificity for drug users and

drug stimuli. The American journal of psychiatry, 157, 1789-1798.

Gazdzinski, S., Durazzo, T. C., Studholme, C., Song, E., Banys, P., & Meyerhoff, D. J. (2005).

Quantitative brain MRI in alcohol dependence: preliminary evidence for effects of

concurrent chronic cigarette smoking on regional brain volumes. Alcoholism, clinical

and experimental research, 29, 1484-1495.

George, M. S., Anton, R. F., Bloomer, C., Teneback, C., Drobes, D. J., Lorberbaum, J. P., Nahas,

Z., et al. (2001). Activation of prefrontal cortex and anterior thalamus in alcoholic

P á g i n a | 1 5 1

subjects on exposure to alcohol-specific cues. Archives of general psychiatry, 58, 345-

352.

George, O., & Koob, G. F. (2010). Individual differences in prefrontal cortex function and

the transition from drug use to drug dependence. Neuroscience and biobehavioral

reviews, 35, 232-247.

Gerra, G., Calbiani, B., Zaimovic, A., Sartori, R., Ugolotti, G., Ippolito, L., Delsignore, R.,

Rustichelli, P., & Fontanesi, B. (1998). Regional cerebral blood flow and comorbid

diagnosis in abstinent opioid addicts. Psychiatry research, 83, 117-126.

Gerra, G., Zaimovic, A., Ferri, M., Zambelli, U., Timpano, M., Neri, E., Marzocchi, G. F., et al.

(2000). Long-lasting effects of (+/-)3,4-methylenedioxymethamphetamine (ecstasy)

on serotonin system function in humans. Biological psychiatry, 47, 127-136.

Giancola, P. R., & Tarter, R. E. (1999). Executive cognitive functioning and risk for

substance abuse. Psychological Science, 10, 203-205.

Gilman, S., Adams, K., Koeppe, R. A., Berent, S., Kluin, K. J., Modell, J. G., Kroll, P., et al.

(1990). Cerebellar and frontal hypometabolism in alcoholic cerebellar degeneration

studied with positron emission tomography. Annals of neurology, 28, 775-785.

Golden, C. J. (1978). Stroop color and Word tests: a manual for clinical and experimental

uses. Wood Dale, Illinois: Stoelting Co.

Goldstein, R. Z., Tomasi, D., Rajaram, S., Cottone, L. A., Zhang, L., Maloney, T., Telang, F., et

al. (2007). Role of the anterior cingulate and medial orbitofrontal cortex in

processing drug cues in cocaine addiction. Neuroscience, 144, 1153-1159.

Goldstein, R. Z., & Volkow, N. D. (2002). Drug addiction and its underlying neurobiological

basis: neuroimaging evidence for the involvement of the frontal cortex. The American

Journal of Psychiatry, 159, 1642-1652.

Goldstein, R. Z., & Volkow, N. D. (2011). Dysfunction of the prefrontal cortex in addiction:

neuroimaging findings and clinical implications. Nature reviews. Neuroscience, 12,

652-669.

Goldstein, R. Z., Alia-Klein, N., Tomasi, D., Carrillo, J. H., Maloney, T., Woicik, P. A., Wang, R.,

et al. (2009). Anterior cingulate cortex hypoactivations to an emotionally salient task

P á g i n a | 1 5 2

in cocaine addiction. Proceedings of the National Academy of Sciences of the United

States of America, 106, 9453-9458.

Goldstein, R. Z., Leskovjan, A. C., Hoff, A. L., Hitzemann, R., Bashan, F., Khalsa, S. S., Wang, G.-

J., et al. (2004). Severity of neuropsychological impairment in cocaine and alcohol

addiction: association with metabolism in the prefrontal cortex. Neuropsychologia,

42, 1447-1458.

Goudriaan, A. E., Oosterlaan, J., de Beurs, E., & van den Brink, W. (2006). Neurocognitive

functions in pathological gambling: a comparison with alcohol dependence, Tourette

syndrome and normal controls. Addiction, 101, 534-547.

Grant, D. A. & Berg, E. A. A. (1948). A behavioral analysis of degree of reinforcement and

ease of shifting to new responses in a Weigel-type card-sorting problem. Journal of

Experimental Psychology, 38, 404-411.

Grant, D. A., & Berg, E. A. (1993). Wisconsin card sorting test. California, Los Angeles:

Western Psychological Services.

Grant, I., Gonzalez, R., Carey, C. L., Natarajan, L., & Wolfson, T. (2003). Non-acute (residual)

neurocognitive effects of cannabis use: a meta-analytic study. Journal of the

International Neuropsychological Society, 9, 679-689.

Grant, S., Contoreggi, C., & London, E. D. (2000). Drug abusers show impaired performance

in a laboratory test of decision making. Neuropsychologia, 38, 1180-1187.

Green, A R, King, M. V., Shortall, S. E., & Fone, K. C. F. (2011). Lost in translation: preclinical

studies on MDMA provide information on mechanisms of action, but do not allow

accurate prediction of adverse events in humans. British journal of pharmacology, en

prensa.

Green, A. R., Mechan, A. O., Elliott, J. M., O’Shea, E., & Colado, M. I. (2003). The

pharmacology and clinical pharmacology of 3,4-methylenedioxymethamphetamine

(MDMA, “ecstasy”). Pharmacological reviews, 55, 463-508.

Gruber, S. A., Silveri, M. M., & Yurgelun-Todd, D. A. (2007). Neuropsychological

consequences of opiate use. Neuropsychology review, 17, 299-315.

P á g i n a | 1 5 3

Gruber, S. A., & Yurgelun-Todd, D. A. (2001). Neuropsychological correlatos of drug abuse.

In Kaufman, M.J. (Ed). Brain imaging in substance abuse: research, clinical and forensic

appliactions. Totowa: New Jersey: Human Press Inc.

Gu, H., Salmeron, B. J., Ross, T. J., Geng, X., Zhan, W., Stein, E. A., & Yang, Y. (2010).

Mesocorticolimbic circuits are impaired in chronic cocaine users as demonstrated by

resting-state functional connectivity. NeuroImage, 53, 593-601.

Halpern, J. H., Pope, H. G. J., Sherwood, A. R., Barry, S., Hudson, J. I., & Yurgelun-Todd, D.

(2004). Residual neuropsychological effects of illicit 3,4-

methylenedioxymethamphetamine (MDMA) in individuals with minimal exposure to

other drugs. Drug and alcohol dependence, 75, 135-147.

Hanson, K. L., Cummins, K., Tapert, S. F., & Brown, S. A. (2011). Changes in

neuropsychological functioning over 10 years following adolescent substance abuse

treatment. Psychology of addictive behaviors, 25, 127-142.

Hanson, K. L., Luciana, M., & Sullwold, K. (2008). Reward-related decision-making deficits

and elevated impulsivity among MDMA and other drug users. Drug and alcohol

dependence, 96, 99-110.

Hartstra, E., Oldenburg, J. F. E., Van Leijenhorst, L., Rombouts, S. A. R. B., & Crone, E. A.

(2010). Brain regions involved in the learning and application of reward rules in a

two-deck gambling task. Neuropsychologia, 48, 1438-1446.

Heberlein, A. S., Padon, A. A., Gillihan, S. J., Farah, M. J., & Fellows, L. K. (2008).

Ventromedial frontal lobe plays a critical role in facial emotion recognition. Journal of

cognitive neuroscience, 20, 721-733.

Heil, S. H., Johnson, M. W., Higgins, S. T., & Bickel, W. K. (2006). Delay discounting in

currently using and currently abstinent cocaine-dependent outpatients and non-

drug-using matched controls. Addictive behaviors, 31, 1290-1294.

Jang, D. P., Namkoong, K., Kim, J. J., Park, S., Kim, I. Y., Kim, S. I., Kim, Y. B., et al. (2007). The

relationship between brain morphometry and neuropsychological performance in

alcohol dependence. Neuroscience Letters, 428,21-26.

P á g i n a | 1 5 4

Jager, G., Kahn, R. S., Van Den Brink, W., Van Ree, J. M., & Ramsey, N. F. (2006). Long-term

effects of frequent cannabis use on working memory and attention: an fMRI study.

Psychopharmacology, 185, 358-368.

Jiang, G. H., Qiu, Y. W., Zhang, X. L., Han, L. J., Lv, X. F., Li, L. M., Lin, C. L., et al. (2011).

Amplitude low-frequency oscillation abnormalities in the heroin users: a resting state

fMRI study. NeuroImage, 57, 149-154.

Johnson-Greene, D., Adams, K. M., Gilman, S., & Junck, L. (2002). Relationship between

neuropsychological and emotional functioning in severe chronic alcoholism. The

Clinical neuropsychologist, 16, 300-309.

Jovanovski, D., Erb, S., & Zakzanis, K. K. (2005). Neurocognitive deficits in cocaine users: a

quantitative review of the evidence. Journal of clinical and experimental

neuropsychology, 27, 189-204.

Kalechstein, A. D., De La Garza, R., Mahoney, J. J., Fantegrossi, W. E., & Newton, T. F. (2007).

MDMA use and neurocognition: a meta-analytic review. Psychopharmacology, 189,

531-537.

Kanwisher, N., & Yovel, G. (2006). The fusiform face area: a cortical region specialized for

the perception of faces. Philosophical transactions of the Royal Society of London.

Series B, Biological sciences, 361, 2109-2128.

Kerr, A., & Zelazo, P. D. (2004). Development of “hot” executive function: the children’s

gambling task. Brain and cognition, 55, 148-157.

Kilts, C. D., Gross, R. E., Ely, T. D., & Drexler, K. P. G. (2004). The neural correlates of cue-

induced craving in cocaine-dependent women. The American journal of psychiatry,

161, 233-241.

King, G. R., Ernst, T., Deng, W., Stenger, A., Gonzales, R. M. K., Nakama, H., & Chang, L.

(2011). Altered brain activation during visuomotor integration in chronic active

cannabis users: relationship to cortisol levels. The Journal of neuroscience, 31, 17923-

17931.

Kirby, K. N., Petry, N. M., & Bickel, W. K. (1999). Heroin addicts have higher discount rates

for delayed rewards than non-drug-using controls. Journal of experimental

psychology. General, 128, 78-87.

P á g i n a | 1 5 5

Kirby, K. N., & Petry, N. M. (2004). Heroin and cocaine abusers have higher discount rates

for delayed rewards than alcoholics or non-drug-using controls. Addiction, 99, 461-

471.

Kish, S. J., Lerch, J., Furukawa, Y., Tong, J., McCluskey, T., Wilkins, D., Houle, S., et al. (2010).

Decreased cerebral cortical serotonin transporter binding in ecstasy users: a positron

emission tomography/[(11)C]DASB and structural brain imaging study. Brain, 133,

1779-1797.

Koob, G. F., & Volkow, N. D. (2010). Neurocircuitry of addiction.

Neuropsychopharmacology, 35, 217-238.

Kornreich, C., Blairy, S., Philippot, P., Hess, U., Noël, X., Streel, E., Le Bon, O., et al. (2001).

Deficits in recognition of emotional facial expression are still present in alcoholics

after mid- to long-term abstinence. Journal of studies on alcohol, 62, 533-542.

Kornreich, C., Foisy, M. L., Philippot, P., Dan, B., Tecco, J., Noël, X., Hess, U., et al. (2003).

Impaired emotional facial expression recognition in alcoholics, opiate dependence

subjects, methadone maintained subjects and mixed alcohol-opiate antecedents

subjects compared with normal controls. Psychiatry research, 119, 251-260.

Kornreich, C., Philippot, P., Foisy, M. L., Blairy, S., Raynaud, E., Dan, B., Hess, U., et al. (2002).

Impaired emotional facial expression recognition is associated with interpersonal

problems in alcoholism. Alcohol and alcoholism, 37, 394-400.

Krain, A. L., Wilson, A. M., Arbuckle, R., Castellanos, F. X., & Milham, M. P. (2006). Distinct

neural mechanisms of risk and ambiguity: a meta-analysis of decision-making.

NeuroImage, 32, 477-484.

Kreek, M. J., Bart, G., Lilly, C., LaForge, K. S., & Nielsen, D. A. (2005). Pharmacogenetics and

human molecular genetics of opiate and cocaine addictions and their treatments.

Pharmacological reviews, 57, 1-26.

Lane, S. D., Steinberg, J. L., Ma, L., Hasan, K. M., Kramer, L. A., Zuniga, E. A., Narayana, P. A.,

et al. (2010). Diffusion tensor imaging and decision making in cocaine dependence.

PloS one, 5, e11591.

P á g i n a | 1 5 6

Lawrence, A. J., Luty, J., Bogdan, N. A., Sahakian, B. J., & Clark, L. (2009). Impulsivity and

response inhibition in alcohol dependence and problem gambling.

Psychopharmacology, 207, 163-172.

Lee, B., London, E. D., Poldrack, R. A., Farahi, J., Nacca, A., Monterosso, J. R., Mumford, J. A.,

et al. (2009). Striatal dopamine d2/d3 receptor availability is reduced in

methamphetamine dependence and is linked to impulsivity. The Journal of

neuroscience, 29, 14734-14740.

Lee, T. M. C., Zhou, W. H., Luo, X. J., Yuen, K. S. L., Ruan, X. Z., & Weng, X. C. (2005). Neural

activity associated with cognitive regulation in heroin users: A fMRI study.

Neuroscience letters, 382, 211-216.

Lee, T. M., & Pau, C. W. (2002). Impulse control differences between abstinent heroin users

and matched controls. Brain Injury, 16, 885-889.

Leland, D. S., Arce, E., Feinstein, J. S., & Paulus, M. P. (2006). Young adult stimulant users’

increased striatal activation during uncertainty is related to impulsivity. NeuroImage,

33, 725-731.

Levine, B., Dawson, D., Boutet, I., Schwartz, M. L., & Stuss, D. T. (2000). Assessment of

strategic self-regulation in traumatic brain injury: its relationship to injury severity

and psychosocial outcome. Neuropsychology, 14, 491-500.

Lezak, M. D. (2004). Neuropsychological Assessment (4th edition). New York: Oxford

University Press.

Li, C. S. R., Milivojevic, V., Kemp, K., Hong, K., & Sinha, R. (2006). Performance monitoring

and stop signal inhibition in abstinent patients with cocaine dependence. Drug and

alcohol dependence, 85, 205-212.

Li, C. S. R., & Sinha, R. (2008). Inhibitory control and emotional stress regulation:

neuroimaging evidence for frontal-limbic dysfunction in psycho-stimulant addiction.

Neuroscience and biobehavioral reviews, 32, 581-597.

Logan, G. D., Cowan, W. B., & Davis, K. A. (1984). On the ability to inhibit simple and choice

reaction time responses: a model and a method. Journal of experimental psychology.

Human perception and performance, 10, 276-291.

P á g i n a | 1 5 7

Lorenzetti, V., Lubman, D. I., Whittle, S., Solowij, N., & Yücel, M. (2010). Structural MRI

findings in long-term cannabis users: what do we know? Substance use & misuse, 45,

1787-1808.

Lubman, D. I., Yücel, M., Kettle, J. W. L., Scaffidi, A., Mackenzie, T., Simmons, J. G., & Allen, N.

B. (2009). Responsiveness to drug cues and natural rewards in opiate addiction:

associations with later heroin use. Archives of general psychiatry, 66, 205-212.

Luijten, M., Veltman, D. J., van den Brink, W., Hester, R., Field, M., Smits, M., & Franken, I. H.

A. (2011). Neurobiological substrate of smoking-related attentional bias.

NeuroImage, 54, 2374-2381.

Lundqvist, T. (2010). Imaging cognitive deficits in drug abuse. Current topics in behavioral

neurosciences, 3, 247-275.

Lyoo, I. K., Streeter, C. C., Ahn, K. H., Lee, H. K., Pollack, M. H., Silveri, M. M., Nassar, L., et al.

(2004). White matter hyperintensities in subjects with cocaine and opiate

dependence and healthy comparison subjects. Psychiatry research, 131, 135-145.

López-Torrecillas, F., Godoy García, J. F., Pérez García, M., Godoy Izquierdo, D., & Sánchez-

Barrera, M. B. (2000). Variables modulating stress and coping that discriminate drug

consumers from low or nondrug consumers. Addictive behaviors, 25, 161-165.

Madden, G. J., Petry, N. M., Badger, G. J., & Bickel, W. K. (1997). Impulsive and self-control

choices in opioid-dependent patients and non-drug-using control participants: drug

and monetary rewards. Experimental and clinical psychopharmacology, 5, 256-262.

Makris, N., Gasic, G. P., Kennedy, D. N., Hodge, S. M., Kaiser, J. R., Lee, M. J., Kim, B. W., et al.

(2008). Cortical thickness abnormalities in cocaine addiction--a reflection of both

drug use and a pre-existing disposition to drug abuse? Neuron, 60, 174-188.

Makris, N., Gasic, G. P., Seidman, L. J., Goldstein, J. M., Gastfriend, D. R., Elman, I., Albaugh, M.

D., et al. (2004b). Decreased absolute amygdala volume in cocaine addicts. Neuron,

44, 729-740.

Makris, N., Oscar-Berman, M., Jaffin, S. K., Hodge, S. M., Kennedy, D. N., Caviness, V. S.,

Marinkovic, K., et al. (2008a). Decreased volume of the brain reward system in

alcoholism. Biological psychiatry, 64, 192-202.

P á g i n a | 1 5 8

Manning, V., Wanigaratne, S., Best, D., Hill, R. G., Reed, L. J., Ball, D., Marshall, J., et al. (2008).

Changes in neuropsychological functioning during alcohol detoxification. European

addiction research, 14, 226-233.

Marinkovic, K., Oscar-Berman, M., Urban, T., O’Reilly, C. E., Howard, J. A., Sawyer, K., &

Harris, G. J. (2009). Alcoholism and dampened temporal limbic activation to

emotional faces. Alcoholism, clinical and experimental research, 33, 1880-1892.

Martinez, D., & Narendran, R. (2010). Imaging neurotransmitter release by drugs of abuse.

Current topics in behavioral neurosciences, 3, 219-245.

Martín-Santos, R., Fagundo, A. B., Crippa, J. A., Atakan, Z., Bhattacharyya, S., Allen, P., Fusar-

Poli, P., et al. (2010). Neuroimaging in cannabis use: a systematic review of the

literature. Psychological medicine, 40, 383-398.

Matochik, J. A., Eldreth, D. A., Cadet, J. L., & Bolla, K. I. (2005). Altered brain tissue

composition in heavy marijuana users. Drug and alcohol dependence, 77, 23-30.

Matochik, J. A., London, E. D., Eldreth, D. A., Cadet, J. L., & Bolla, K. I. (2003). Frontal cortical

tissue composition in abstinent cocaine abusers: a magnetic resonance imaging

study. NeuroImage, 19, 1095-1102.

Matsuo, K., Nicoletti, M., Nemoto, K., Hatch, J. P., Peluso, M. A. M., Nery, F. G., & Soares, J. C.

(2009). A voxel-based morphometry study of frontal gray matter correlates of

impulsivity. Human brain mapping, 30, 1188-1195.

Medina, K. L., Hanson, K. L., Schweinsburg, A. D., Cohen-Zion, M., Nagel, B. J., & Tapert, S. F.

(2007). Neuropsychological functioning in adolescent marijuana users: subtle deficits

detectable after a month of abstinence. Journal of the International

Neuropsychological Society, 13, 807-820.

Mintzer, M. Z., Copersino, M. L., & Stitzer, M. L. (2005). Opioid abuse and cognitive

performance. Drug and alcohol dependence, 78, 225-230.

Mishina, M. (2008). Positron emission tomography for brain research. Journal of Nihon

Medical School, 75, 68-76.

Mitchell, J. M., Fields, H. L., D’Esposito, M., & Boettiger, C. A. (2005). Impulsive responding

in alcoholics. Alcoholism, clinical and experimental research, 29, 2158-2169.

P á g i n a | 1 5 9

Miyake, A., Friedman, N. P., Emerson, M. J., Witzki, A. H., Howerter, A., & Wager, T. D.

(2000). The unity and diversity of executive functions and their contributions to

complex “Frontal Lobe” tasks: a latent variable analysis. Cognitive psychology, 41, 49-

100.

Moeller, F. G., Dougherty, D. M., Barratt, E. S., Oderinde, V., Mathias, C. W., Harper, R. A., &

Swann, A. C. (2002). Increased impulsivity in cocaine dependent subjects

independent of antisocial personality disorder and aggression. Drug and alcohol

dependence, 68, 105-111.

Moeller, F. G., Dougherty, D. M., Barratt, E. S., Schmitz, J. M., Swann, A. C., & Grabowski, J.

(2001). The impact of impulsivity on cocaine use and retention in treatment. Journal

of substance abuse treatment, 21, 193-198.

Monterosso, J. R., Aron, A. R., Cordova, X., Xu, J., & London, E. D. (2005). Deficits in response

inhibition associated with chronic methamphetamine abuse. Drug and alcohol

dependence, 79, 273-277.

Montgomery, C., Seddon, A. L., Fisk, J. E., Murphy, P. N., & Jansari, A. (2012). Cannabis-

related deficits in real-world memory. Human psychopharmacology, 27, 217-225.

Morgan, M. J., Impallomeni, L. C., Pirona, A., & Rogers, R. D. (2006). Elevated impulsivity

and impaired decision-making in abstinent Ecstasy (MDMA) users compared to

polydrug and drug-naïve controls. Neuropsychopharmacology, 31, 1562-1573.

Moriyama, Y., Muramatsu, T., Kato, M., Mimura, M., & Kashima, H. (2006). Family history of

alcoholism and cognitive recovery in subacute withdrawal. Psychiatry and clinical

neurosciences, 60, 85-89.

Nelson, S. M., Dosenbach, N. U. F., Cohen, A. L., Wheeler, M. E., Schlaggar, B. L., & Petersen, S.

E. (2010). Role of the anterior insula in task-level control and focal attention. Brain

structure & function, 214, 669-680.

Nestler, E. J. (2005). The neurobiology of cocaine addiction. Science & practice perspectives

/ a publication of the National Institute on Drug Abuse, National Institutes of Health, 3,

4-10.

P á g i n a | 1 6 0

Nestor, L. J., Ghahremani, D. G., Monterosso, J., & London, E. D. (2011). Prefrontal

hypoactivation during cognitive control in early abstinent methamphetamine-

dependent subjects. Psychiatry research, 194, 287-295.

Noël, X., Van der Linden, M., Schmidt, N., Sferrazza, R., Hanak, C., Le Bon, O., De Mol, J., et al.

(2001). Supervisory attentional system in nonamnesic alcoholic men. Archives of

general psychiatry, 58, 1152-1158.

Obrocki, J., Buchert, R., Väterlein, O., Thomasius, R., Beyer, W., & Schiemann, T. (1999).

Ecstasy--long-term effects on the human central nervous system revealed by positron

emission tomography. The British journal of psychiatry, 175, 186-188.

Obrocki, J., Schmoldt, A., Buchert, R., Andresen, B., Petersen, K., & Thomasius, R. (2002).

Specific neurotoxicity of chronic use of ecstasy. Toxicology letters, 127, 285-297.

Observatorio Español sobre Drogas (2009). Informe 2009. Descargado de:

http://www.pnsd.msc.es/Categoria2/observa/pdf/oed-2009.pdf el 2 de Mayo de

2012.

Oquendo, M. A., Baca-García, E., Graver, R., Morales, M., Montalvan, V., & Mann, J. J. (2001).

Spanish adaptation of the Barratt Impulsiveness Scale (BIS-11). European Journal of

Psychiatry, 15, 147-155.

Pace-Schott, E. F., Morgan, P. T., Malison, R. T., Hart, C. L., Edgar, C., Walker, M., & Stickgold,

R. (2008). Cocaine users differ from normals on cognitive tasks which show poorer

performance during drug abstinence. The American journal of drug and alcohol abuse,

34, 109-121.

Parada, M., Corral, M., Mota, N., Crego, A., Rodríguez Holguín, S., & Cadaveira, F. (2012).

Executive functioning and alcohol binge drinking in university students. Addictive

behaviors, 37, 167-172.

Parrott, A. C., Sisk, E., & Turner, J. J. (2000). Psychobiological problems in heavy “ecstasy”

(MDMA) polydrug users. Drug and alcohol dependence, 60, 105-110.

Parrott, A. C. (2012). MDMA and serotonergic neurotoxicity: the empirical evidence for its

adverse effects in humans-no need for translation. British journal of pharmacology, en

prensa.

P á g i n a | 1 6 1

Passetti, F., Clark, L., Mehta, M. A., Joyce, E., & King, M. (2008). Neuropsychological

predictors of clinical outcome in opiate addiction. Drug and alcohol dependence, 94,

82-91.

Patton, J. H., Stanford, M. S., & Barratt, E. S. (1995). Factor structure of the Barratt

impulsiveness scale. Journal of clinical psychology, 51, 768-774.

Paulus, M. P., Tapert, S. F., & Schuckit, M. A. (2005). Neural activation patterns of

methamphetamine-dependent subjects during decision making predict relapse.

Archives of general psychiatry, 62, 761-768.

Pérez, J., & Torrubia, R. (1986). Fiabilidad y validez de la versión española de la escala de

búsqueda de sensaciones (formaV). Revista Latinoamericana de Psicología, 18, 7-22.

Pertwee, R. G., & Ross, R. A. (2002). Cannabinoid receptors and their ligands.

Prostaglandins, leukotrienes, and essential fatty acids, 66, 101-121.

Petry, N. M. (2001). Delay discounting of money and alcohol in actively using alcoholics,

currently abstinent alcoholics, and controls. Psychopharmacology, 154, 243-250.

Pezawas, L., Fischer, G., Podreka, I., Schindler, S., Brücke, T., Jagsch, R., Thurnher, M., et al.

(2002). Opioid addiction changes cerebral blood flow symmetry. Neuropsychobiology,

45, 67-73.

Pezawas, L. M., Fischer, G., Diamant, K., Schneider, C., Schindler, S. D., Thurnher, M., Ploechl,

W., et al. (1998). Cerebral CT findings in male opioid-dependent patients:

stereological, planimetric and linear measurements. Psychiatry research, 83, 139-147.

Pillay, S. S., Rogowska, J., Kanayama, G., Gruber, S., Simpson, N., Pope, H. G., & Yurgelun-

Todd, D. A. (2008). Cannabis and motor function: fMRI changes following 28 days of

discontinuation. Experimental and clinical psychopharmacology, 16, 22-32.

Pitel, A. L., Perruchet, P., Vabret, F., Desgranges, B., Eustache, F., & Beaunieux, H. (2010).

The advantage of errorless learning for the acquisition of new concepts’ labels in

alcoholics. Psychological medicine, 40, 497-502.

Pope, H. G., Gruber, A. J., Hudson, J. I., Cohane, G., Huestis, M. A., & Yurgelun-Todd, D.

(2003). Early-onset cannabis use and cognitive deficits: what is the nature of the

association? Drug and alcohol dependence, 69, 303-310.

P á g i n a | 1 6 2

Posner, M. I., Rothbart, M. K., Sheese, B. E., & Tang, Y. (2007). The anterior cingulate gyrus

and the mechanism of self-regulation. Cognitive, affective & behavioral neuroscience,

7, 391-395.

Potenza, M. N. (2001). The neurobiology of pathological gambling. Seminars in clinical

neuropsychiatry, 6, 217-226.

Prince van Leeuwen, A., Creemers, H. E., Verhulst, F. C., Ormel, J., & Huizink, A. C. (2011).

Are adolescents gambling with cannabis use? A longitudinal study of impulsivity

measures and adolescent substance use: the TRAILS study. Journal of studies on

alcohol and drugs, 72, 70-78.

Prosser, J. M., Eisenberg, D., Davey, E. E., Steinfeld, M., Cohen, L. J., London, E. D., &

Galynker, I. I. (2008). Character pathology and neuropsychological test performance

in remitted opiate dependence. Substance abuse treatment, prevention, and policy, 3,

23.

Quednow, B. B., Kühn, K. U., Hoppe, C., Westheide, J., Maier, W., Daum, I., & Wagner, M.

(2007). Elevated impulsivity and impaired decision-making cognition in heavy users

of MDMA (“Ecstasy”). Psychopharmacology, 189, 517-530.

Redish, A. D., Jensen, S., & Johnson, A. (2008). A unified framework for addiction:

vulnerabilities in the decision process. The Behavioral and brain sciences, 31, 415-

437.

Reid, A. G., Daglish, M. R. C., Kempton, M. J., Williams, T. M., Watson, B., Nutt, D. J., &

Lingford-Hughes, A. R. (2008). Reduced thalamic grey matter volume in opioid

dependence is influenced by degree of alcohol use: a voxel-based morphometry

study. Journal of psychopharmacology, 22, 7-10.

Rey, A. (1941). L'examen psychologique dans les cas d'encephalopathie traumatique.

Archives de Psychologie, 28, 286-340.

Reynolds, B. (2006). A review of delay-discounting research with humans: relations to

drug use and gambling. Behavioural pharmacology, 17, 651-667.

Reynolds, B., & Schiffbauer, R. (2004). Measuring state changes in human delay

discounting: an experiential discounting task. Behavioural processes, 67, 343-356.

P á g i n a | 1 6 3

Rogers, R. D., Everitt, B. J., Baldacchino, A., Blackshaw, A. J., Swainson, R., Wynne, K., Baker,

N. B., et al. (1999a). Dissociable deficits in the decision-making cognition of chronic

amphetamine abusers, opiate abusers, patients with focal damage to prefrontal

cortex, and tryptophan-depleted normal volunteers: evidence for monoaminergic

mechanisms. Neuropsychopharmacology, 20, 322-339.

Rogers, R. D., Owen, A. M., Middleton, H. C., Williams, E. J., Pickard, J. D., Sahakian, B. J., &

Robbins, T. W. (1999b). Choosing between small, likely rewards and large, unlikely

rewards activates inferior and orbital prefrontal cortex. The Journal of neuroscience,

19, 9029-9038.

Rogers, R. D., & Robbins, T. W. (2001). Investigating the neurocognitive deficits associated

with chronic drug misuse. Current opinion in neurobiology, 11, 250-257.

Roberts, A. C., Robbins, T. W., & Weiskrantz, L. (1998). The Prefrontal Cortex: Executive and

Cognitive Functions. New York: Oxford University Press.

Romero, M. J., Asensio, S., Palau, C., Sanchez, A., & Romero, F. J. (2010). Cocaine addiction:

diffusion tensor imaging study of the inferior frontal and anterior cingulate white

matter. Psychiatry research, 181, 57-63.

Rose, J. E., Behm, F. M., Westman, E. C., Mathew, R. J., London, E. D., Hawk, T. C., Turkington,

T. G., et al. (2003). PET studies of the influences of nicotine on neural systems in

cigarette smokers. The American journal of psychiatry, 160, 323-333.

Ruff, R. M. (1996). Ruff Figural Fluency Test: Professional Manual. Lutz, Florida:

Psychological Assessment Resources.

Salloum, J. B., Ramchandani, V. A., Bodurka, J., Rawlings, R., Momenan, R., George, D., &

Hommer, D. W. (2007). Blunted rostral anterior cingulate response during a

simplified decoding task of negative emotional facial expressions in alcoholic

patients. Alcoholism, clinical and experimental research, 31, 1490-1504.

Salthouse TA (2009) When does age-related cognitive decline begin? Neurobiology of

Aging, 30, 507-514.

Schulz, S. (2011). MDMA & cannabis: a mini-review of cognitive, behavioral, and

neurobiological effects of co-consumption. Current drug abuse reviews, 4, 81-86.

P á g i n a | 1 6 4

Sedó, M. (2005). Test de los Cinco Dígitos: Five Digit Test. Madrid: TEA Ediciones.

Sevy, S., Smith, G. S., Ma, Y., Dhawan, V., Chaly, T., Kingsley, P. B., Kumra, S., et al. (2008).

Cerebral glucose metabolism and D2/D3 receptor availability in young adults with

cannabis dependence measured with positron emission tomography.

Psychopharmacology, 197, 549-556.

Sim, M. E., Lyoo, I. K., Streeter, C. C., Covell, J., Sarid-Segal, O., Ciraulo, D. A., Kim, M. J., et al.

(2007). Cerebellar gray matter volume correlates with duration of cocaine use in

cocaine-dependent subjects. Neuropsychopharmacology, 32, 2229-2237.

Sinha, R., & Li, C. S. R. (2007). Imaging stress- and cue-induced drug and alcohol craving:

association with relapse and clinical implications. Drug and alcohol review, 26, 25-31.

Smith, G. T., Fischer, S., Cyders, M. A., Annus, A. M., Spillane, N. S., & McCarthy, D. M. (2007).

On the validity and utility of discriminating among impulsivity-like traits. Assessment,

14, 155-170.

Sofuoglu, M. (2010). Cognitive enhancement as a pharmacotherapy target for stimulant

addiction. Addiction, 105, 38-48.

Spiro, R. J., & Jehng, J. (1990). Cognitive flexibility and hipertext: Theory and technology for

the non-linear and multidimensional traversal of complex subject matter. D. Nix y R.

Spiro (Eds.), Cognition, Education, and Multimedia. Hillsdale, New Jersey: Erlbaum.

Squeglia, L. M., Sorg, S. F., Schweinsburg, A. D., Wetherill, R. R., Pulido, C., & Tapert, S. F.

(2012). Binge drinking differentially affects adolescent male and female brain

morphometry. Psychopharmacology, 220, 529-539.

Stout, J. C., Busemeyer, J. R., Lin, A., Grant, S. J., & Bonson, K. R. (2004). Cognitive modeling

analysis of decision-making processes in cocaine abusers. Psychonomic bulletin &

review, 11, 742-747.

Streeter, C. C., Terhune, D. B., Whitfield, T. H., Gruber, S., Sarid-Segal, O., Silveri, M. M.,

Tzilos, G., et al. (2008). Performance on the Stroop predicts treatment compliance in

cocaine-dependent individuals. Neuropsychopharmacology, 33, 827-836.

Stuss, D. T., & Alexander, M. P. (2000). Executive functions and the frontal lobes: a

conceptual view. Psychological Research, 63, 289-298.

P á g i n a | 1 6 5

Stuss, D. T., & Knight, R. T. (2002). Principles of frontal lobe functioning. New York: Oxford

University Press.

Sullivan, E. V., Deshmukh, A., De Rosa, E., Rosenbloom, M. J., & Pfefferbaum, A. (2005).

Striatal and forebrain nuclei volumes: contribution to motor function and working

memory deficits in alcoholism. Biological psychiatry, 57, 768-776.

Tanabe, J., Tregellas, J. R., Dalwani, M., Thompson, L., Owens, E., Crowley, T., & Banich, M.

(2009). Medial orbitofrontal cortex gray matter is reduced in abstinent substance-

dependent individuals. Biological psychiatry, 65, 160-164.

Tapert, S. F., Brown, G. G., Baratta, M. V., & Brown, S. A. (2004). fMRI BOLD response to

alcohol stimuli in alcohol dependent young women. Addictive behaviors, 29, 33-50.

Tapert, S. F., Schweinsburg, A. D., Drummond, S. P. A., Paulus, M. P., Brown, S. A., Yang, T. T.,

& Frank, L. R. (2007). Functional MRI of inhibitory processing in abstinent adolescent

marijuana users. Psychopharmacology, 194, 173-183.

Tedstone, D., & Coyle, K. (2004). Cognitive impairments in sober alcoholics: performance

on selective and divided attention tasks. Drug and alcohol dependence, 75, 277-286.

Thomas, L. A., Hall, J. M., Skup, M., Jenkins, S. E., Pine, D. S., & Leibenluft, E. (2011). A

developmental neuroimaging investigation of the change paradigm. Developmental

science, 14, 148-161.

Tomasi, D., Goldstein, R. Z., Telang, F., Maloney, T., Alia-Klein, N., Caparelli, E. C., & Volkow,

N. D. (2007). Widespread disruption in brain activation patterns to a working

memory task during cocaine abstinence. Brain research, 1171, 83-92.

Torrubia, R., Ávila, C., Moltó, J., & Caseras, X. (2001). The Sensitivity to Punishment and

Sensitivity to Reward Questionnaire (SPSRQ) as a measure of Gray’s anxiety and

impulsivity dimensions. Personality and Individual Differences, 31, 837-862.

Townshend, J. M., & Duka, T. (2003). Mixed emotions: alcoholics’ impairments in the

recognition of specific emotional facial expressions. Neuropsychologia, 41, 773-782.

Tucker, K. A., Potenza, M. N., Beauvais, J. E., Browndyke, J. N., Gottschalk, P. C., & Kosten, T.

R. (2004). Perfusion abnormalities and decision making in cocaine dependence.

Biological psychiatry, 56, 527-530.

P á g i n a | 1 6 6

Turner, T. H., LaRowe, S., Horner, M. D., Herron, J., & Malcolm, R. (2009). Measures of

cognitive functioning as predictors of treatment outcome for cocaine dependence.

Journal of substance abuse treatment, 37, 328-334.

Tzourio-Mazoyer, N., Landeau, B., Papathanassiou, D., Crivello, F., Etard, O., Delcroix, N.,

Mazoyer, B., et al. (2002). Automated anatomical labeling of activations in SPM using

a macroscopic anatomical parcellation of the MNI MRI single-subject brain.

NeuroImage, 15, 273-289.

United Nations Office on Drugs and Crime (2011). United Nations World Drug Report.

Descargado de: http://www.unodc.org/unodc/en/data-and-analysis/WDR-

2011.html el 2 de Mayo de 2012.

van Holst, R. J., & Schilt, T. (2011). Drug-related decrease in neuropsychological functions

of abstinent drug users. Current drug abuse reviews, 4, 42-56.

van Holst, R. J., de Ruiter, M. B., van den Brink, W., Veltman, D. J., & Goudriaan, A. E. (2012).

A voxel-based morphometry study comparing problem gamblers, alcohol abusers,

and healthy controls. Drug and alcohol dependence, en prensa.

Verdejo, A., Toribio, I., Orozco, C., Puente, K. L., & Pérez-García, M. (2005).

Neuropsychological functioning in methadone maintenance patients versus abstinent

heroin abusers. Drug and alcohol dependence, 78, 283-288.

Verdejo-García, A., Benbrook, A., Funderburk, F., David, P., Cadet, J.-L., & Bolla, K. I. (2007a).

The differential relationship between cocaine use and marijuana use on decision-

making performance over repeat testing with the Iowa Gambling Task. Drug and

alcohol dependence, 90, 2-11.

Verdejo-García, A, Betanzos-Espinosa, P., Lozano, O. M., Vergara-Moragues, E., González-

Saiz, F., Fernández-Calderón, F., Bilbao-Acedos, I., et al. (2012). Self-regulation and

treatment retention in cocaine dependent individuals: a longitudinal study. Drug and

alcohol dependence, 122, 142-148.

Verdejo-García, A., & Bechara, A. (2009). A somatic marker theory of addiction.

Neuropharmacology, 56, 48-62.

Verdejo-García, A., Bechara, A., Recknor, E. C., & Pérez-García, M. (2006). Executive

dysfunction in substance dependent individuals during drug use and abstinence: an

P á g i n a | 1 6 7

examination of the behavioral, cognitive and emotional correlates of addiction.

Journal of the International Neuropsychological Society, 12, 405-415.

Verdejo-García, A., Bechara, A., Recknor, E. C., & Pérez-García, M. (2007c). Negative

emotion-driven impulsivity predicts substance dependence problems. Drug and

alcohol dependence, 91, 213-219.

Verdejo-García, A. J., López-Torrecillas, F., Aguilar de Arcos, F., & Pérez-García, M. (2005).

Differential effects of MDMA, cocaine, and cannabis use severity on distinctive

components of the executive functions in polysubstance users: a multiple regression

analysis. Addictive behaviors, 30, 89-101.

Verdejo-García, A. J., Perales, J. C., & Pérez-García, M. (2007b). Cognitive impulsivity in

cocaine and heroin polysubstance abusers. Addictive behaviors, 32, 950-966.

Verdejo-García, A., Lawrence, A. J., & Clark, L. (2008). Impulsivity as a vulnerability marker

for substance-use disorders: review of findings from high-risk research, problem

gamblers and genetic association studies. Neuroscience and biobehavioral reviews, 32,

777-810.

Verdejo-García, A., Lozano, O., Moya, M., Alcázar, M. A., & Pérez-García, M. (2010).

Psychometric properties of a Spanish version of the UPPS-P impulsive behavior scale:

reliability, validity and association with trait and cognitive impulsivity. Journal of

personality assessment, 92, 70-77.

Verdejo-García, A., López-Torrecillas, F., Giménez, C. O., & Pérez-García, M. (2004). Clinical

implications and methodological challenges in the study of the neuropsychological

correlates of cannabis, stimulant, and opioid abuse. Neuropsychology review, 14, 1-41.

Verdejo-García, A., & Pérez-García, M. (2007). Profile of executive deficits in cocaine and

heroin polysubstance users: common and differential effects on separate executive

components. Psychopharmacology, 190, 517-530.

Verdejo-García, A., Rivas-Pérez, C., Vilar-López, R., & Pérez-García, M. (2007d). Strategic

self-regulation, decision-making and emotion processing in poly-substance abusers

in their first year of abstinence. Drug and alcohol dependence, 86, 139-146.

Volkow, N. D., Chang, L., Wang, G. J., Fowler, J. S., Ding, Y. S., Sedler, M., Logan, J., et al.

(2001). Low level of brain dopamine D2 receptors in methamphetamine abusers:

P á g i n a | 1 6 8

association with metabolism in the orbitofrontal cortex. The American journal of

psychiatry, 158, 2015-2021.

Volkow, N. D., Fowler, J. S., Wang, G. J., Hitzemann, R., Logan, J., Schlyer, D. J., Dewey, S. L., et

al. (1993). Decreased dopamine D2 receptor availability is associated with reduced

frontal metabolism in cocaine abusers. Synapse, 14, 169-177.

Volkow, N. D., Gillespie, H., Mullani, N., Tancredi, L., Grant, C., Valentine, A., & Hollister, L.

(1996). Brain glucose metabolism in chronic marijuana users at baseline and during

marijuana intoxication. Psychiatry research, 67, 29-38.

Volkow, N. D., Hitzemann, R., Wang, G. J., Fowler, J. S., Wolf, A. P., Dewey, S. L., &

Handlesman, L. (1992). Long-term frontal brain metabolic changes in cocaine

abusers. Synapse, 11, 184-190.

Volkow, N. D., Ma, Y., Zhu, W., Fowler, J. S., Li, J., Rao, M., Mueller, K., et al. (2008). Moderate

doses of alcohol disrupt the functional organization of the human brain. Psychiatry

research, 162, 205-213.

Volkow, N. D., Tomasi, D., Wang, G. J., Fowler, J. S., Telang, F., Goldstein, R. Z., Alia-Klein, N.,

et al. (2011). Reduced metabolism in brain “control networks” following cocaine-cues

exposure in female cocaine abusers. PloS one, 6, e16573.

Volkow, N. D., Fowler, J. S., Wang, G. J., Telang, F., Logan, J., Jayne, M., Ma, Y., et al. (2010).

Cognitive control of drug craving inhibits brain reward regions in cocaine abusers.

NeuroImage, 49, 2536-2543.

Volkow, N. D., Wang, G. J., Telang, F., Fowler, J. S., Logan, J., Childress, A. R., Jayne, M., et al.

(2006). Cocaine cues and dopamine in dorsal striatum: mechanism of craving in

cocaine addiction. The Journal of neuroscience, 26, 6583-6588.

Volz, K. G., Schubotz, R. I., & von Cramon, D. Y. (2003). Predicting events of varying

probability: uncertainty investigated by fMRI. NeuroImage, 19, 271-280.

Wang, G. J., Volkow, N. D., Fowler, J. S., Cervany, P., Hitzemann, R. J., Pappas, N. R., Wong, C.

T., et al. (1999). Regional brain metabolic activation during craving elicited by recall

of previous drug experiences. Life sciences, 64, 775-784.

P á g i n a | 1 6 9

Wang, Z. X., Zhang, J. X., Wu, Q.-L., Liu, N., Hu, X.-P., Chan, R. C. K., & Xiao, Z. W. (2010).

Alterations in the processing of non-drug-related affective stimuli in abstinent heroin

addicts. NeuroImage, 49, 971-976.

Watter, S., Geffen, G. M., & Geffen, L. B. (2001). The n-back as a dual-task: P300 morphology

under divided attention. Psychophysiology, 38, 998-1003.

Weijers, H. G., Wiesbeck, G. A., & Böning, J. (2001). Reflection–impulsivity, personality and

performance: a psychometric and validity study of the Matching Familiar Figures

Test in detoxified alcoholics. Personality and Individual Differences, 31, 731-754.

Wechsler, D. (1997). Wechsler memory scale (3rd Edition). San Antonio, Texas: The

Psychological Corporation.

Whiteside, S. P., & Lynam, D. R. (2001). The five factor model and impulsivity: using a

structural model of personality to understand impulsivity. Personality and Individual

Differences, 30, 669-689.

Whiteside, S. P., & Lynam, D. R. (2003). Understanding the role of impulsivity and

externalizing psychopathology in alcohol abuse: application of the UPPS impulsive

behavior scale. Experimental and Clinical Psychopharmacology, 11, 210-217.

Wilson, W., Mathew, R., Turkington, T., Hawk, T., Coleman, R. E., & Provenzale, J. (2000).

Brain morphological changes and early marijuana use: a magnetic resonance and

positron emission tomography study. Journal of addictive diseases, 19, 1-22.

Woicik, P. A., Moeller, S. J., Alia-Klein, N., Maloney, T., Lukasik, T. M., Yeliosof, O., Wang, G.-J.,

et al. (2009). The neuropsychology of cocaine addiction: recent cocaine use masks

impairment. Neuropsychopharmacology, 34, 1112-1122.

Yaouhi, K., Bertran, F., Clochon, P., Mézenge, F., Denise, P., Foret, J., Eustache, F., et al.

(2009). A combined neuropsychological and brain imaging study of obstructive sleep

apnea. Journal of sleep research, 18, 36-48.

Yip, J. T., & Lee, T.M. (2006). Selective impairment of sadness and disgust recognition in

abstinent ecstasy users. Neuropsychologia, 44, 959-965.

P á g i n a | 1 7 0

Young, A. W., Perrett, D. I., Calder, A. J., Sprengelmeyer, R., & Ekman, P. (2002). Facial

expressions of emotion: stimuli and tests (FEEST). Bury St. Edmunds: Thames Valley

Test Company.

Yuan, Y., Zhu, Z., Shi, J., Zou, Z., Yuan, F., Liu, Y., Lee, T. M. C., et al. (2009). Gray matter

density negatively correlates with duration of heroin use in young lifetime heroin-

dependent individuals. Brain and cognition, 71, 223-228.

Yücel, M., Solowij, N., Respondek, C., Whittle, S., Fornito, A., Pantelis, C., & Lubman, D. I.

(2008). Regional brain abnormalities associated with long-term heavy cannabis use.

Archives of general psychiatry, 65, 694-701.

Zakzanis, K. K. (2001). Statistics to tell the truth, the whole truth, and nothing but the

truth: formulae, illustrative numerical examples, and heuristic interpretation of effect

size analyses for neuropsychological researchers. Archives of clinical neuropsychology,

16, 653-667.

Zakzanis, K. K., Campbell, Z., & Jovanovski, D. (2007). The neuropsychology of ecstasy

(MDMA) use: a quantitative review. Human psychopharmacology, 22, 427-435.

Zinn, S., Stein, R., & Swartzwelder, H. S. (2004). Executive functioning early in abstinence

from alcohol. Alcoholism, clinical and experimental research, 28, 1338-1346.

Zubieta, J. K., Heitzeg, M. M., Xu, Y., Koeppe, R. A., Ni, L., Guthrie, S., & Domino, E. F. (2005).

Regional cerebral blood flow responses to smoking in tobacco smokers after

overnight abstinence. The American journal of psychiatry, 162, 567-577.

Zuckerman, M., Kuhlman, D., Joireman, J., Teta, P., & Kraft, M. (1993). A comparison of the

three structural models for the personality: the big three, the big five and the

alternative five. Journal of Personality and Social Psychology, 65, 747-768.

P á g i n a | 1 7 1

[ANEXOS]

[ANEXO I]

G

D

Ti

LEa

b

c

d

e

a

ARRAA

KCIGWVU

1

lfHcuitec

p

0d

ARTICLE IN PRESS Model

AD-4380; No. of Pages 7

Drug and Alcohol Dependence xxx (2012) xxx– xxx

Contents lists available at SciVerse ScienceDirect

Drug and Alcohol Dependence

jo u rn al hom epage: www.elsev ier .com/ locate /drugalcdep

rait impulsivity and prefrontal gray matter reductions in cocaine dependentndividuals

aura Moreno-Lópeza, Andrés Catenab, María José Fernández-Serranoc, Elena Delgado-Ricoa,mmanuel A. Stamatakisd, Miguel Pérez-Garcíaa,e, Antonio Verdejo-Garcíaa,e,∗

Department of Clinical Psychology, School of Psychology, University of Granada, SpainDepartment of Experimental Psychology and Behavioral Physiology, School of Psychology, University of Granada, SpainDepartment of Psychology, University of Jaen, SpainDivision of Anaesthesia, School of Clinical Medicine, University of Cambridge, UKInstitute of Neuroscience F. Olóriz, University of Granada, Spain

r t i c l e i n f o

rticle history:eceived 22 August 2011eceived in revised form 9 February 2012ccepted 10 February 2012vailable online xxx

eywords:ocaine

mpulsivityray matterhite matter

oxel based morphometryPPS

a b s t r a c t

Background: Impulsivity is thought to play a key role in cocaine addiction onset and progression; therefore,we hypothesized that different facets of impulsive personality may be significantly associated with brainstructural abnormalities in cocaine-dependent individuals.Methods: Thirty-eight cocaine-dependent individuals and 38 non-drug using controls completed theUPPS-P scale (measuring five different facets of impulsivity: sensation seeking, lack of premeditation,lack of perseverance, and positive and negative urgency) and were scanned on a 3T MRI scanner. Weused whole-brain voxel-based morphometry analyses (VBM) to detect differences in gray matter (GM)and white matter (WM) volumes between cocaine users and controls, and to measure differences in theway that impulsivity relates to GM and WM volumes in cocaine users vs. controls.Results: Cocaine-dependent individuals had lower GM volumes in a number of sections of the orbitofrontalcortex, right inferior frontal gyrus, right insula, left amygdala and parahippocampal gyrus, temporalgyrus, and bilateral caudate. They also had lower WM volumes in the left inferior and medial frontal

gyrus, superior temporal gyrus, right anterior cingulate cortex, insula and caudate. There was a positivecorrelation between trait impulsivity and GM volume in the left inferior/middle frontal gyrus of cocaine-dependent individuals, a pattern directly opposed to the association in controls. Conversely, in cocaineusers lack of premeditation was negatively correlated with GM volume in the insula and the putamen.Conclusions: Trait impulsivity may influence cocaine dependence by impacting its neurobiological under-pinnings in frontostriatal systems.

. Introduction

Cocaine addiction is a worldwide major public health prob-em for which current prevention and treatment options are notully satisfactory (EMCDDA, 2009; Substance Abuse and Mentalealth Services Administration, 2010). Neuroimaging studies inocaine-dependent individuals have revealed significant brain vol-me reductions, most notably in a priori selected regions of

nterest, such as the striatum, the amygdala or the prefrontal cor-

Please cite this article in press as: Moreno-López, L., et al., Trait impulsivity anDrug Alcohol Depend. (2012), doi:10.1016/j.drugalcdep.2012.02.012

ex (Barrós-Loscertales et al., 2011; Makris et al., 2004; Matochikt al., 2003; Tanabe et al., 2009). These regions are thought toontribute to critical aspects of the addictive cycle, including

∗ Corresponding author at: Facultad de Psicología, Universidad de Granada Cam-us de Cartuja S/N, 18071 Granada, Spain. Tel.: +34 958242948; fax: +34 958243749.

E-mail address: [email protected] (A. Verdejo-García).

376-8716/$ – see front matter © 2012 Elsevier Ireland Ltd. All rights reserved.oi:10.1016/j.drugalcdep.2012.02.012

© 2012 Elsevier Ireland Ltd. All rights reserved.

reinforcement learning, craving, and inhibitory control (Koob andVolkow, 2010). Nonetheless, the findings from nonbiased auto-mated techniques, such as Voxel Based Morphometry (VBM),have yielded inconsistent results. Despite previous positivefindings showing significant gray matter (GM) reductions incocaine-dependent individuals, including the medial prefrontalcortex, superior temporal cortex, insula, thalamus and cerebel-lum (Franklin et al., 2002; Sim et al., 2007), a recent VBM studyfailed to detect any structural change (GM or white matter (WM)reductions) in cocaine users compared to non-drug using controls(Narayana et al., 2010). Perhaps more critically, most of these stud-ies have failed to find any correlation between estimates of druguse patterns (e.g. amount, duration or age at onset) and GM and

d prefrontal gray matter reductions in cocaine dependent individuals.

WM reductions (Franklin et al., 2002; Makris et al., 2004; Matochiket al., 2003; but see Ersche et al., 2011). This apparent lack ofassociation between cocaine exposure and brain attrition raisesthe possibility that certain personality traits, such as impulsivity,

ING Model

D

2 Alcoho

mi

mtetrthStisvittastftcsii

yccduiuldragftaeas

2

2

npdprt(httatccdd(

ARTICLEAD-4380; No. of Pages 7

L. Moreno-López et al. / Drug and

ay relate to GM and WM abnormalities in cocaine-dependentndividuals.

Impulsivity is viewed as a multifaceted trait that varies nor-ally across the population, but that in high levels may predispose

o a range of dysfunctional behaviors, including addiction (Cyderst al., 2007; Verdejo-García et al., 2008). Animal studies have shownhat increased impulsivity is associated with reduced dopamineeceptors availability and cocaine use escalation and progressiono dependence (Belin et al., 2008; Dalley et al., 2007). In humans,igh levels of trait impulsivity (indexed with the Barratt Impulsivitycale) are associated with lower midbrain dopamine autorecep-or binding, and greater amphetamine-induced dopamine releasen the striatum (Buckholtz et al., 2010). In addition, trait impul-ivity is negatively correlated with orbitofrontal cortex (OFC) GMolumes in healthy individuals (Matsuo et al., 2009). Nonetheless,n addicted individuals, continued use of stimulants is thoughto further exacerbate impulsive traits (Ersche et al., 2010), ando possibly modify its neural underpinnings. In fact, the associ-tion between trait impulsivity levels and striatal dysfunction istronger in methamphetamine-dependent individuals comparedo healthy controls (Lee et al., 2009). Conversely, Ersche et al. (2011)ailed to find significant correlations between particular aspects ofrait impulsivity (impulsive reward-seeking) and GM reductions inocaine-dependent subjects. Therefore, more studies are needed topecifically explore if, as expected, different facets of trait impulsiv-ty are specifically linked to GM volumes in frontostriatal regionsn cocaine-dependent individuals.

The aims of this study were (i) to use whole-brain VBM anal-ses to examine possible GM and WM reductions in a sample ofurrently abstinent (>1 month) cocaine-dependent individuals, asompared to a non-drug using control group; and (ii) to measureifferences in the way impulsivity relates to GM and WM vol-mes in cocaine users vs. controls. Because impulsive personality

n the normal population is negatively associated with GM vol-me in the OFC (Matsuo et al., 2009), which is also impacted by

ifetime cocaine use (Alia-Klein et al., 2011), we expected cocaine-ependent individuals to have reduced GM volumes in the OFC andichly interconnected regions, such as insula, amygdala, striatum,nd WM adjacent to these regions (Bechara, 2005). Because pro-ression of cocaine use is thought to provoke neuroadaptationsrom the ventral striatum to the dorsal striatum, and from the OFCo more extensive prefrontal regions like the anterior cingulatend dorsolateral prefrontal cortex (Koob and Volkow, 2010), wexpected impulsivity to be uniquely associated with more lateralspects of the prefrontal cortex and more posterior aspects of thetriatum in cocaine-dependent individuals.

. Materials and methods

.1. Participants

Thirty-eight cocaine-dependent individuals, mean age = 29.58, SD = 6.53, and 38on-drug using controls, mean age = 31.08, SD = 5.14, participated in this study. Allarticipants were male due to the low prevalence of women entering drug treatmenturing the recruitment period. Cocaine users were recruited in an inpatient thera-eutic community (“Proyecto Hombre”), in the city of Granada, Spain. All of themeported cocaine as their main drug of choice and the one for which they requestedreatment. Clinical interviews based on Diagnostic and Statistical Manual version IVDSM-IV) criteria confirmed cocaine dependence diagnosis; nonetheless, they alsoad regular use of tobacco, alcohol, cannabis and MDMA (see Table 1). To enterhe study, cocaine users had to be abstinent for at least 30 days (for any drug butobacco), as confirmed by weekly urine tests; in this way, we could rule out acutend residual effects of previously used drugs on brain structure, with the excep-ion of nicotine (84.2% of cocaine-dependent individuals and 44.7% of controls were

Please cite this article in press as: Moreno-López, L., et al., Trait impulsivity anDrug Alcohol Depend. (2012), doi:10.1016/j.drugalcdep.2012.02.012

urrent smokers). None of the cocaine patients were currently following pharma-ological substitution treatments. Potential participants who had previously beeniagnosed with any disorder from DSM-IV Axes I and II (other than substance depen-ence), or had neurological or systemic diseases affecting central nervous systemCNS) functioning were excluded.

PRESSl Dependence xxx (2012) xxx– xxx

Non-drug using controls were recruited through adverts distributed by a localemployment agency, and therefore they were also matched to cocaine participantsin terms of unemployment status. Selection criteria for control participants were:(i) absence of current or past substance use, excluding past or current social drink-ing (less than ten standard alcohol units per week); (ii) absence of documentedmajor psychiatric disorders; (iii) absence of documented head injury or neurologicaldisorder, and (iv) not being taking medication with effects on the CNS.

For both groups, evidence of stroke or space-occupying lesions observed onconventional clinical MR images, any contraindications to MRI scanning (includingclaustrophobia and implanted ferromagnetic objects), and history of loss of con-sciousness (LOC) for longer than 30 min or LOC with any neurological consequencewere exclusionary.

2.2. Instruments and assessment procedures

The study was approved by the Ethics Committee for Research in Humans of theUniversity of Granada. All participants signed an informed consent form certifyingtheir voluntary participation. Controls, but not patients, received a D 40 compensa-tion for participating in the study. Assessments were conducted across two sessionsseparated by less than one week. During the first session we administered the Inter-view for Research on Addictive Behavior, the Hamilton Rating Scales for Depressionand Anxiety (HAM-D and HAM-A), and the UPPS-P Impulsive Behavior Scale, alongwith a battery of cognitive tests of which results will be reported separately. Thesecond session involved the MRI scanning. The MRI scans lasted approximately6 min.

2.2.1. Patterns of drug use. Data regarding lifetime amount and duration of use ofthe different drugs was self-reported by participants and collected using the Inter-view for Research on Addictive Behavior (Verdejo-García et al., 2005). This interviewprovides an estimation of monthly use of each substance during regular use (amountper month) and total duration of use of each substance (in years). The descriptivescores for these variables in the sample are presented in Table 1.

2.2.2. Trait impulsivity. We used the Spanish version of the UPPS-P Scale (Verdejo-García et al., 2010; Whiteside and Lynam, 2001). This is a 59-item self-reportinventory designed to measure five distinct personality pathways to impulsivebehavior: sensation seeking, (lack of) perseverance, (lack of) premeditation, neg-ative urgency, and positive urgency. The first 4 dimensions were included in theoriginal version of the UPPS scale (Whiteside and Lynam, 2001); the fifth dimensionhas been included based on recent work by (Cyders et al., 2007; Smith et al., 2007).Sensation seeking (12 items) incorporates two aspects: (1) a tendency to enjoy andpursue activities that are exciting, and (2) an openness to trying new experiencesthat may or may not be dangerous; (lack of) perseverance (10 items) refers to theindividual’s ability to remain focused on a task that may be boring or difficult; (lackof) premeditation (11 items) refers to the tendency to think and reflect on the con-sequences of an act before engaging in that act; and finally urgency (12 items) refersto the tendency to experience strong impulses under conditions of negative affect(negative urgency, 12 items) or positive affect (positive urgency, 14 items). Each itemon the UPPS is rated on a 4-point scale ranging from 1 (strongly agree) to 4 (stronglydisagree). We obtained the total scores of each of these five UPPS-P dimensions foranalysis (Table 2).

2.3. MRI acquisition

Participants were scanned on a 3T whole body MRI scanner (Phillips Achieva)operating with 8 channels phased-array head coil for reception. For each participant,a T1-weighted 3D volume was acquired using a T1-weighted 3D-turbo-gradient-echo sequence (3D-TFE), in sagittal orientation with 0.94 × 0.94 × 1.0 mm resolution(160 slides, FOV = 240 × 240 mm2, matrix 256 × 256 × 160) with repetition time of8 ms, echo time 4 ms, inversion delay = 1022.6264 ms, flip angle of 8◦ , band with191 Hz/pixel. The sequence was optimal for reducing motion sensitivity, suscepti-bility artifacts and field inhomogeneities.

2.4. Image analysis

Before automatic preprocessing, the images were checked for artifacts andmanually aligned to the AC-PC line. Data were processed and analyzed usingSPM8 (http://www.fil.ion.ucl.ac.uk/spm) and VBM8 toolbox (http://dbm.neuro.uni-jena.de/vbm.html), for which we used the default parameters. Thus, within a unifiedsegmentation model (Ashburner and Friston, 2005), images were corrected for bias-field inhomogeneities, registered using linear (12 parameters affine) and non-lineartransformations (warped), and tissue was clustered into GM, WM, and cerebrospinalfluid (CSF). Segments were further refined by using adaptive maximum a posteri-ori estimations, which account for partial volume effects, and by applying a hidden

d prefrontal gray matter reductions in cocaine dependent individuals.

Markov random field model, as implemented in VBM8. Importantly, to preservelocal GM/WM values, we multiplied the segments by the Jacobian determinantsof the deformation field to create modulated images. Segments were smoothed byan 8 mm full-width at half-maximum (FWHM) using an isotropic Gaussian kernel.Afterwards, we conducted analysis on modulated GM and WM segments.

ARTICLE IN PRESSG Model

DAD-4380; No. of Pages 7

L. Moreno-López et al. / Drug and Alcohol Dependence xxx (2012) xxx– xxx 3

Table 1Descriptive information about patterns of drug use in cocaine-dependent individuals.

Substance Ever used (%) Unit Amount per month Duration of use Abstinence

Cocaine 100 # grams 18.92 ± 29.46 4.05 ± 3.07 36.76 ± 22.60Cannabis 76.31 # joints 109.90 ± 111.40 2.35 ± 3.17 147.45 ± 203.66MDMA 50 # tablets 9.29 ± 10.12 2.27 ± 1.90 227.68 ± 225.78Alcohol 100 # SAU 107.58 ± 114.03 7.38 ± 5.77 33.11 ± 21.77Tobacco 86.8 # cigarettes 488.48 ± 307.40 9.19 ± 7 7.8 ± 15.82

Note: Mean ± standard deviation of the mean. Duration of use is expressed in years. Abstinence duration is expressed in number of weeks.

Table 2Descriptive scores – means and standard deviations (in parentheses) of cocaine users and non-drug using controls in the five UPPS-P dimensions.

UPPS-P dimensions Cocaine-dependent individuals Non-drug using controls Test

Negative urgency 31.45 (6.79) 23.16 (6.99) (t46 = 5.14, p = 0.000)Positive urgency 32.91 (8.79) 21.66 (7.63) (t46 = 5.86, p = 0.000)

2(i(tttos

2ntcacwvfg

2etUPic(

3

3

aceaC[Sb

3

ts(p

Lack of premeditation 23.83 (5.95)

Lack of perseverance 21.26 (4.42)

Sensation seeking 30.63 (8.73)

.4.1. Global effects of patterns of drug use. Total gray and white matter volumesTGMV, TWMV) were computed by adding up the voxel values of GM/WM segments,ncluding cerebellum and sub-cortical structures. Next, the total brain volumeTBV = TGMV + TWMV) was used for computing the TGMV/TBV ratios, which servedo evaluate the global effects of patterns of drug use on brain tissue volumes usinghe backward step-wise procedure in the general linear model, including as predic-ors age, years of education, cocaine amount per month during regular use, durationf cocaine use, and age of onset of cocaine use. t-Tests were used to estimate theignificance of the regression parameters. Significance threshold was set at p < 0.05.

.4.2. Regional GM and WM differences between cocaine-dependent individuals andon-drug using controls. The general linear model implemented in SPM8 was usedo conduct voxel-wise comparisons between cocaine-dependent individuals andontrols, and to determine the relationship between consume severity factors (i.e.mount, duration and age of onset) and regional GM and WM volume variations inocaine-dependent individuals. In both analyses the total volume of GM, or WM,as modeled as a linear confound in order to account for residual global volume

ariability. The significance threshold was set at p < 0.05 after family-wise correctionor multiple comparisons (pFWE < 0.05). Significant peaks from the t-test maps areiven in MNI space.

.4.3. Regional relationships between GM and WM and measures of impulsivity andstimates of drug use. We performed SPM8 multiple regression analysis, in which weested the significances of the Group (cocaine-dependent individuals vs. controls) xPPS-P dimensions interactions across the entire brain. We only included the UPPS-

dimensions showing significant differences between the groups. The differencesn the regression coefficients as a function of group were tested using a signifi-ance threshold of p < 0.05 after family-wise correction for multiple comparisonspFWE < 0.05).

. Results

.1. Participants’ characteristics

Both groups were matched on gender, ethnicity and language,nd had statistically equivalent distributions for age. Controls,ompared to cocaine-dependent individuals, had more years ofducation [(M = 17.58, SD = 4.56) vs. (M = 12.03, SD = 3.62)]; butnalysis including education as a covariate did not alter the results.ocaine-dependent individuals had greater scores for depression(M = 5.08, SD = 3.98) vs. (M = 1.32, SD = 2.15)] and anxiety [(M = 7.05,D = 6.40) vs. (M = 1.29, SD = 2.29)], but on average both groups wereelow cut-offs for clinical significance.

.2. Trait impulsivity

Cocaine-dependent individuals had higher scores than con-

Please cite this article in press as: Moreno-López, L., et al., Trait impulsivity anDrug Alcohol Depend. (2012), doi:10.1016/j.drugalcdep.2012.02.012

rols across the five dimensions of the UPPS-P Scale. t-Testshowed these differences were significant for lack of premeditationt = 2.46, p = 0.016, Cohen’s d = 0.6), lack of perseverance (t = 2.56,

= 0.013, Cohen’s d = 0.6), negative urgency (t = 5.14, p = 0.000,

21 (3.68) (t46 = 2.46, p = 0.016)18.84 (3.63) (t46 = 2.56, p = 0.013)29.89 (7.37) (t46 = 0,34, p = 0.73)

Cohen’s d = 1.2), and positive urgency (t = 5.86, p = 0.000, Cohen’sd = 1.4). We found no significant differences for sensation seeking,such that this dimension was not further used in the subsequentanalysis.

3.3. Imaging analysis

3.3.1. Global effects. For cocaine-dependent individuals, theTGMV/TBV ratio variability was accounted for by monthly cocaineintake during regular use, t(34) = 2.19, p = 0.035, and age of onset,t(34) = −2.33, p = 0.026; the effect of duration of cocaine use alsoshowed a trend to significance, t(34) = −1.80, p = 0.080. Thesevariables explained about a 32% of the variability in TGMV/TBVratios, F(3,34) = 5.21, p = 0.005. Noteworthy, higher amounts ofcocaine (r = 0.41, p = 0.01) and younger age of onset of cocaine use(r = −.34, p = 0.04) were associated with larger TGMV/TBV. The firstresult is associated with a significant decrease in WM volumes(TWMV-amount of cocaine: r = 0.43, p = 0.006), but the secondshowed a trend in relation to reductions in GM (TGMV-age of onsetof cocaine abuse: r = −0.18, NS).

3.3.2. Regional GM and WM differences between cocaine-dependentindividuals and non-drug using controls. The cocaine group hadsignificantly lower GM and WM volumes than controls. Cocaine-dependent individuals showed lower GM volumes in a numberof sections of the OFC (left superior and medial frontal gyrus),right inferior frontal gyrus, right insula, left amygdala and parahip-pocampal gyrus, left inferior and middle temporal gyrus, andbilateral caudate (Table 3 and Fig. 1). Likewise, cocaine dependent-individuals showed lower WM volumes in the left inferior andmedial frontal gyrus, superior temporal gyrus, right anterior cin-gulate cortex, insula and caudate (Table 4 and Fig. 2).

3.3.3. Regional relationships between GM and WM and measuresof impulsivity and estimates of drug use. We found a number ofregions in which GM volumes showed differential associations withthe impulsivity dimensions of lack of premeditation and negativeurgency; in all cases, there was a significant Group x Impulsiv-ity interaction (see Table 5). The associations between impulsivitylevels and GM volumes were significantly different for cocaine-dependent individuals vs. controls in the left inferior frontal gyrus,right insula and left putamen (in relation to lack of premeditation),

d prefrontal gray matter reductions in cocaine dependent individuals.

and left middle frontal gyrus and sub-gyral (in relation to negativeurgency). Left inferior frontal gyrus was positively associated withlack of premeditation in cocaine users, whereas it was negativelycorrelated with this dimension in controls. Conversely, right insula

ARTICLE IN PRESSG Model

DAD-4380; No. of Pages 7

4 L. Moreno-López et al. / Drug and Alcohol Dependence xxx (2012) xxx– xxx

Table 3Summary of the results obtained by the gray matter voxel-wise analysis of volume reductions in cocaine-dependent individuals and non-drug using controls.

Anatomical region k Anatomical labels for peaks BA T p x y z

FrontalLobe

3342 L Medial Frontal Gyrus 11, 10 5.81 0.004 −9 27 −13Cingulate Gyrus 32, 25, 24 5.75 0.005

Frontal Lobe 2802 R Inferior Frontal Gyrus 47, 11, 34 6.39 0.000 28 21 −15Caudate R Caudate Head 5.53 0.010 9 19 −15Caudate 1118 L Caudate Head 5.52 0.010 6 −1 3Limbic Lobe/Parahippocampal Gyrus 972 L Amygdala/Hippocampus 6.42 0.000 −33 −7 −19Sub-Lobar 581 R Insula 13 5.81 0.004 52 −19 23Temporal Lobe 3095 L Middle/Inferior Temporal Gyrus 21, 22, 20 5.98 0.002 −56 −14 −18

Fig. 1. Brain regions showing significant gray matter decreases in cocaine-dependent individuals with respect to non-drug using controls.

Table 4Summary of the results obtained by the white matter voxel-wise analysis of volume reductions in cocaine-dependent individuals and non-drug using controls.

Anatomical Region k Anatomical Labels for peaks T p x y z

FrontalLobe

1181L Inferior Frontal Gyrus

6.07 0.001 −26 18 −17Sub-GyralFrontal Lobe 623 L Medial Frontal Gyrus 5.46 0.003 −20 57 −3Temporal Lobe 389 L Superior Temporal Gyrus 5.24 0.007 −48 −25 −3FrontalLobe

1651R Anterior Cingulate

6.81 0.001 7 33 −15R Medial Frontal Gyrus

Sub-Lobar 1531Caudate

5.86 0.001 12 12 −3Extra-NuclearLentiform Nucleus

TemporalLobe

674Insula

5.82 0.001 46 −26 12R Superior Temporal GyrusR Transverse Temporal Gyrus

Fig. 2. Brain regions showing significant white matter decreases in cocaine-dependent individuals with respect to non-drug using controls.

Table 5Regression slopes showing different associations between impulsivity and gray matter volumes between cocaine-dependent individuals (CDI) and non-drug-using controls(NDC) (pFWE = 0.05).

Anatomical region k Anatomical Label BA p (ˇ) Slopes Volume x y z

Lack of premeditation

Sub-lobar179 R Insula BA 47/BA

130.02 CDI−/NDC+ CDI < NDC 30 10 −3

358 L Putamen 0.05 CDI−/NDC+ CDI < NDC −21 10 −4Fontal Lobe 249 L Inferior Frontal Gyrus BA 10 0.005 CDI+/NDC− CDI < NDC −37 41 1

Please cite this article in press as: Moreno-López, L., et al., Trait impulsivity anDrug Alcohol Depend. (2012), doi:10.1016/j.drugalcdep.2012.02.012

Negative urgencyFrontalLobe

228 L Middle Frontal Gyrus BA10/46

381 R Sub-Gyral BA 8

d prefrontal gray matter reductions in cocaine dependent individuals.

0.038 CDI+/NDC− CDI < NDC −39 37 190.024 CDI+/NDC− CDI < NDC 25 24 36

ING Model

D

Alcoho

aiisamza

ag

mc

4

parriiiowmuccwiiamotta

uitt(Macreumst2Gpq

ttfism

ARTICLEAD-4380; No. of Pages 7

L. Moreno-López et al. / Drug and

nd left putamen were negatively correlated with lack of premed-tation in cocaine users, but did not correlate with this dimensionn controls. With regard to negative urgency, the sub-gyral showedignificant positive correlations in the cocaine group, whereas thessociation was negative in controls. Finally, in the case of the leftiddle frontal gyrus, none of the slopes significantly differed from

ero, but the trend was positive in cocaine-dependent individuals,nd negative in controls.

We did not find significant associations between WM volumesnd the different dimensions of impulsivity in the cocaine or controlroups.

Finally, we did not find significant associations between theain patterns of drug use (amount, duration and age of onset of

ocaine abuse) and GM and WM volumes.

. Discussion

Our results showed that cocaine-dependent individuals, com-ared to controls, have lower GM volumes in the OFC cortex,nterior cingulate, inferior frontal gyrus, insula, amygdala, tempo-al gyrus, and caudate. We also found WM reductions in adjacentegions of the anterior cingulate, inferior/middle frontal gyrus,nsula and putamen. As expected, we found significant differencesn the way that impulsivity dimensions correlates with GM volumesn cocaine users vs. controls: cocaine patients had elevated levelsf trait impulsivity, and these scores were differentially associatedith GM volumes in the left inferior frontal gyrus, insula and puta-en (lack of premeditation), and left middle frontal gyrus (negative

rgency). Inferior and medial superior frontal clusters positivelyorrelated with impulsivity in cocaine users, and negatively inontrols; conversely, insula and putamen correlated negativelyith impulsivity in cocaine patients and showed no correlations

n controls. These results point to a distinctive link between traitmpulsivity and regional GM among cocaine-dependent individu-ls, which specifically relates to frontostriatal systems regions. Theeasures of patterns of cocaine use correlated with general indices

f GM/TBV ratios but had no significant influence on GM volumes athe stricter voxel level. Overall, our findings are in agreement withhe notion that GM reductions in cocaine-dependent individualsre at least partly mediated by trait impulsivity.

Our findings of GM reductions in cocaine-dependent individ-als are largely consistent with previous neuroimaging studies

n revealing structural abnormalities in brain regions relevanto reinforcement learning (amygdala, OFC) and inhibitory con-rol (anterior cingulate, inferior frontal gyrus, insula and caudate)Ersche et al., 2011; Franklin et al., 2002; Makris et al., 2004;

atochik et al., 2003). Recent findings from fMRI connectivitynalyses also provide support to the notion that neural networksonnecting the medial frontal cortex with paralimbic and tempo-al regions are significantly less functional in cocaine users (Gut al., 2010). Biological abnormalities in these regions are thought tonderlie psychological alterations leading to overestimation of theotivational relevance of drug-related reinforcement (the impul-

ive system) and to breakdown of inhibitory processes necessaryo achieve long-term abstinence (the reflective system) (Bechara,005; Goldstein and Volkow, 2002; Redish et al., 2008; Verdejo-arcía and Bechara, 2009). These alterations are proposed to anchorervasive drug taking in the face of increasing aversive conse-uences, one of the key features of addiction (DSM-IV).

A key unresolved issue is that if these structural abnormali-ies are reflective of cocaine induced brain attrition, or are related

Please cite this article in press as: Moreno-López, L., et al., Trait impulsivity anDrug Alcohol Depend. (2012), doi:10.1016/j.drugalcdep.2012.02.012

o personality traits, or reflect a combination of both factors. Ourndings, in agreement with growing neuroscience evidence, areupportive of the impact of personality traits. However, these traitsay interact with cocaine use in producing GM alterations, as

PRESSl Dependence xxx (2012) xxx– xxx 5

suggested by the global TGMV analysis. Recent studies indicatethat, in healthy individuals, impulsivity is negatively correlatedwith midbrain and striatal D2/D3 receptors availability (Buckholtzet al., 2010; Lee et al., 2009), but this association is strongerin methamphetamine dependent individuals (Lee et al., 2009)and correlates with amphetamine-induced dopamine release inthe striatum (Buckholtz et al., 2010). Striatal D2 availability isassociated with metabolic rate in the prefrontal cortex of psychos-timulant users (Volkow et al., 2001). Therefore, different strandsof evidence indicate that frontostriatal systems mediate impul-sive behavior and thereby promote psychostimulant addiction.Nonetheless, we acknowledge that these issues can only be testedby longitudinal imaging studies in individuals at high risk of devel-oping psychostimulant dependence, or by gene association imagingstudies (Kreek et al., 2005; Verdejo-García et al., 2008).

In accordance with the notion that impulsivity is a multifacetedconstruct (Smith et al., 2007; Whiteside and Lynam, 2001), ourresults showed differential links between impulsivity dimensionsand GM volumes. Lack of premeditation was positively associatedwith left inferior frontal gyrus GM in cocaine users but not con-trols. This region is importantly involved in the representationof expected values (Bermpohl et al., 2010); such that persistentstimulation by drug-expected rewards could have induced GMenlargements linked to an increased tendency to rapidly engagein non-adequately forethought behavior. On the other hand, bothinsula and putamen have been involved in drug-related neuroadap-tations related to attentional bias and craving (Childress et al., 2008;Luijten et al., 2011; Volkow et al., 2006), such that reduced volumesmay predict increased prepotency of drug-related action tenden-cies. On the other hand, negative urgency was positively associatedwith GM in subgyral BA 8 in cocaine users, negatively in controls.This region has been previously associated with the experience ofuncertainty (Volz et al., 2003), which is a key aspect of the associa-tive learning processes mediating expected drug effects (D’Souzaand Duvauchelle, 2008). In fact, adjacent regions of the medialfrontal gyrus (BA 10) display increased activation during uncertainoutcome selection in psychostimulant users (Leland et al., 2006).Therefore, in this case the rationale could be similar to that positedfor inferior frontal gyrus; repeated exposure to uncertainty-loadedscenarios may be associated with increased subgyral volumesand higher risk to trigger impulsive acts when under emotionallyuneasy conditions. Remarkably, in both dimensions (lack of pre-meditation and urgency) there was a positive correlation betweentrait impulsivity and GM volume in the inferior/middle frontalgyrus of cocaine-dependent individuals, a pattern directly opposedto the association in controls. Given that GM in these regions isreduced in cocaine users, one may speculatively argue that highimpulsivity may render some cocaine users more similar to con-trols. Nonetheless, given the role of these regions in emotionalvaluation, greater GM in these areas may hold specific emotionalmotives in cocaine users: high impulsive individuals, with greaterGM in inferior/middle frontal gyrus, might be – for example – highlysensitive to craving (Garavan et al., 2000). More research is war-ranted to resolve this issue.

Strengths of this study include the adequate sample size, thesound clinical characterization and the agreement between resultsfrom the different analytical approaches (between-group differ-ences and regressions testing the link between impulsivity andGM). On the other hand, we should also acknowledge a numberof limitations. The cross-sectional design precludes us from draw-ing conclusions about the causality of GM deficits; one way outof addressing this issue would be by using longitudinal designs,

d prefrontal gray matter reductions in cocaine dependent individuals.

but these studies are costly and convey important ethical con-cerns and methodological complexities. Furthermore, we cannotcompletely rule out the influence of depression and anxiety lev-els, although they were on average below the cut-off for clinical

ING Model

D

6 Alcoho

sepuccdct

R

MsaiifSQ

C

ihaEt

C

prysp

R

A

AB

B

B

B

B

C

C

D

ARTICLEAD-4380; No. of Pages 7

L. Moreno-López et al. / Drug and

ignificance and they had very low impact on regression mod-ls when included. Finally, although we discuss our results asertaining to cocaine dependence, polysubstance abuse is almostbiquitous among cocaine-dependent individuals. However, theurrent sample was carefully selected for relative specificity ofocaine related problems and low use of other drugs. The illegalrugs more frequently and intensely co-abused in the sample wereannabis and MDMA, but estimates of the use of these drugs failedo predict regional GM.

ole of funding source

This study has been funded by grants SEJ2006-08278, Spanishinistry of Education and Science and P07-HUM-03089, Andalu-

ia Council of Science and Innovation (PI: Miguel Pérez-García)nd grant COPERNICO, Plan Nacional sobre Drogas: Spanish Min-stry of Health (PI: Antonio Verdejo-García). Laura Moreno-Lópezs supported by a FPU predoctoral research grant (AP2007-03583)rom the Spanish Ministry of Science and Innovation. Emmanuel A.tamatakis is supported by the Stephen Erskine Fellowship fromueens’ College, Cambridge.

ontributors

Laura Moreno-López and Antonio Verdejo-García have workedn the conception and design of the manuscript. Andrés Catenaas carried out the analysis and interpretation of the data as wells the revision of the manuscript. María José Fernández-Serrano,lena Delgado-Rico and Miguel Pérez-García have contributed tohe revision of the manuscript.

onflict of interest

The authors declare that, except for income received from theirrimary employer, no financial support or compensation has beeneceived from any individual or corporate entity over the past threeears for research or professional service and there are no per-onal financial holdings that could be perceived as constituting aotential conflict of interest.

eferences

lia-Klein, N., Parvaz, M.A., Woicik, P.A., Konova, A.B., Maloney, T., Shumay, E.,Wang, R., Telang, F., Giegon, Al., Wang, G.J., Fowler, J.S., Tomasi, D., Volkow, N.D.,Goldstein, R.Z., 2011. Gene × disease interaction on orbitofrontal gray matter incocaine addiction. Arch. Gen. Psychiatry 68, 283–294.

shburner, J., Friston, K.J., 2005. Unified segmentation. Neuroimage 26, 839–851.arrós-Loscertales, A., Garavan, H., Bustamante, J.C., Ventura-Campos, N., Llopis, J.J.,

Belloch, V., Parcet, M.A., Avila, C., 2011. Reduced striatal volume in cocaine-dependent patients. Neuroimage 56, 1021–1026.

echara, A., 2005. Decision making, impulse control and loss of willpower to resistdrugs: a neurocognitive perspective. Nat. Neurosci. 8, 1458–1463.

elin, D., Mar, A.C., Dalley, J.W., Robbins, T.W., Everitt, B.J., 2008. Highimpulsivity predicts the switch to compulsive cocaine-taking. Science 320,1352–1355.

ermpohl, F., Kahnt, T., Dalanay, U., Hägele, C., Sajonz, B., Wegner, T., Stoy, M., Adli, M.,Kruger, S., Wrase, J., Strohle, A., Bauer, M., Heinz, A., 2010. Altered representationof expected value in the orbitofrontal cortex in mania. Hum. Brain Mapp. 31,958–969.

uckholtz, J.W., Treadway, M.T., Cowan, R.L., Woodward, N.D., Li, R., Ansari, M.S.,Baldwin, R.M., Schwartzman, A.N., Shelby, E.S., Smith, C.E., Kessler, R.M., Zaid,D.H., 2010. Dopaminergic network differences in human impulsivity. Science329, 532.

hildress, A.R., Ehrman, R.N., Wang, Z., Li, Y., Sciortino, N., Hakun, J., Jens, W., Suh, J.,Listerud, J., Marquez, K., Franklin, T., Langleben, D., Detre, J., O’Brien, C.P., 2008.Prelude to passion: limbic activation by “unseen” drug and sexual cues. PLoSONE 3, e1506.

yders, M.A., Smith, G.T., Spillane, N.S., Fischer, S., Annus, A.M., Peterson, C.,

Please cite this article in press as: Moreno-López, L., et al., Trait impulsivity anDrug Alcohol Depend. (2012), doi:10.1016/j.drugalcdep.2012.02.012

2007. Integration of impulsivity and positive mood to predict risky behavior:development and validation of a measure of positive urgency. Psychol. Assess.19, 107–118.

alley, J.W., Fryer, T.D., Brichard, L., Robinson, E.S., Theobald, D.E., Lääne, K., Pena,Y., Murphy, E.R., Shah, Y., Probst, K., Abakumova, I., Aigbirhio, F.I., Richards,

PRESSl Dependence xxx (2012) xxx– xxx

H.K., Hong, Y., Baron, J.C., Everitt, B.J., Robbins, T.W., 2007. Nucleus accumbensD2/3 receptors predict trait impulsivity and cocaine reinforcement. Science 315,1267–1270.

D’Souza, M.S., Duvauchelle, C.L., 2008. Certain or uncertain cocaine expectationsinfluence accumbens dopamine responses to self-administered cocaine andnon-rewarded operant behavior. Eur. Neuropsychopharmacol. 18, 628–638.

EMCDDA European Monitoring Centre for Drugs and Drug Addiction, 2009. AnnualReport 2009: The State of the Drugs Problem in Europe. Publications Office ofthe European Union, Luxembourg.

Ersche, K.D., Barnes, A., Simon Jones, P., Morein-Zamir, S., Robbins, T.W., Bullmore,E.T., 2011. Abnormal structure of frontostriatal brain systems is associated withaspects of impulsivity and compulsivity in cocaine dependence. Brain 134,2013–2024.

Ersche, K.D., Bullmore, E.T., Craig, K.J., Shabbir, S.S., Abbott, S., Müller, U., Ooi, C., Suck-ling, J., Barnes, A., Sahakian, B.J., Merlo-Pich, E.V., Robbins, T.W., 2010. Influenceof compulsivity of drug abuse on dopaminergic modulation of attentional biasin stimulant dependence. Arch. Gen. Psychiatry 67, 632–644.

Franklin, T.R., Acton, P.D., Maldjian, J.A., Gray, J.D., Croft, J.R., Dackis, C.A., O’Brien,C.P., Childress, A.R., 2002. Decreased gray matter concentration in the insular,orbitofrontal, cingulate, and temporal cortices of cocaine patients. Biol. Psychi-atry 51, 134–142.

Garavan, H., Pankiewicz, J., Bloom, A., Cho, J.K., Sperry, L., Ross, T.J., Salmeron, B.J.,Risinger, R., Kelley, D., Stein, E.A., 2000. Cue-induced cocaine craving: neu-roanatomical specificity for drug users and drug stimuli. Am. J. Psychiatry 157,1789–1798.

Goldstein, R.Z., Volkow, N.D., 2002. Drug addiction and its underlying neurobiologi-cal basis: neuroimaging evidence for the involvement of the frontal cortex. Am.J. Psychiatry 159, 1642–1652.

Gu, H., Salmeron, B.J., Ross, T.J., Geng, X., Zhan, W., Stein, E.A., Yang, Y., 2010. Meso-corticolimbic circuits are impaired in chronic cocaine users as demonstrated byresting-state functional connectivity. Neuroimage 53, 593–601.

Koob, G.F., Volkow, N.D., 2010. Neurocircuitry of addiction. Neuropsychopharma-cology 35, 217–238.

Kreek, M.J., Bart, G., Lilly, C., LaForge, K.S., Nielsen, D.A., 2005. Pharmacogenetics andhuman molecular genetics of opiate and cocaine addictions and their treatments.Pharmacol. Rev. 57, 1–26.

Lee, B., London, E.D., Poldrack, R.A., Farahi, J., Nacca, A., Monterosso, J.R., Mumford,J.A., Bokarius, A.V., Dahlborn, M., Mukherjee, J., Bilder, R.M., Brody, A.L., Man-delkern, M.A., 2009. Striatal dopamine d2/d3 receptor availability is reducedin methamphetamine dependence and is linked to impulsivity. J. Neurosci. 29,14734–14740.

Leland, D.S., Arce, E., Feinstein, J.S., Paulus, M.P., 2006. Young adult stimulant users’increased striatal activation during uncertainty is related to impulsivity. Neu-roimage 33, 725–731.

Luijten, M., Veltman, D.J., van den Brink, W., Hester, R., Field, M., Smits, M., Franken,I.H., 2011. Neurobiological substrate of smoking-related attentional bias. Neu-roimage 54, 2374–2381.

Makris, N., Gasic, G.P., Seidman, L.J., Goldstein, J.M., Gastfriend, D.R., Elman, I.,Albaugh, M.D., Hodge, S.M., Ziegler, D.A., Sheahan, F.S., Caviness Jr., V.S., Tsuang,M.T., Kennedy, D.N., Hyman, S.E., Rosen, B.R., Breiter, H.C., 2004. Decreased abso-lute amygdala volume in cocaine addicts. Neuron 44, 729–740.

Matochik, J.A., London, E.D., Eldreth, D.A., Cadet, J.L., Bolla, K.I., 2003. Frontal corticaltissue composition in abstinent cocaine abusers: a magnetic resonance imagingstudy. Neuroimage 19, 1095–1102.

Matsuo, K., Nicoletti, M., Nemoto, K., Hatch, J.P., Peluso, M.A., Nery, F.G., Soares, J.C.,2009. A voxel-based morphometry study of frontal gray matter correlates ofimpulsivity. Hum. Brain Mapp. 30, 1188–1195.

Narayana, P.A., Datta, S., Tao, G., Steinberg, J.L., Moeller, F.G., 2010. Effect of cocaine onstructural changes in brain: MRI volumetry using tensor-based morphometry.Drug Alcohol Depend. 111, 191–199.

Redish, A.D., Jensen, S., Johnson, A., 2008. A unified framework for addiction: vul-nerabilities in the decision process. Behav. Brain Sci. 31, 415–437.

Sim, M.E., Lyoo, I.K., Streeter, C.C., Covell, J., Sarid-Segal, O., Ciraulo, D.A., Kim, M.J.,Kaufman, M.J., Yurgelun-Todd, D.A., Renshaw, P.F., 2007. Cerebellar gray mattervolume correlates with duration of cocaine use in cocaine-dependent subjects.Neuropsychopharmacology 32, 2229–2237.

Smith, G.T., Fischer, S., Cyders, M.A., Annus, A.M., Spillane, N.S., 2007. On the valid-ity and utility of discriminating among impulsivity-like traits. Assessment 14,155–170.

Substance Abuse and Mental Health Services Administration, 2010. Results from the2009 National Survey on Drug Use and Health: Mental Health Findings (Officeof Applied Studies, NSDUH Series H-39, HHS Publication No. SMA 10-4609),Rockville, MD.

Tanabe, J., Tregellas, J.R., Dalwani, M., Thompson, L., Owens, E., Crowley, T., et al.,2009. Medial orbitofrontal cortex gray matter is reduced in abstinent substance-dependent individuals. Biol. Psychiatry 65, 160–164.

Verdejo-García, A., Bechara, A., 2009. A somatic marker theory of addiction. Neu-ropharmacology 56, 48–62.

Verdejo-García, A., Lawrence, A.J., Clark, L., 2008. Impulsivity as a vulnerabilitymarker for substance-use disorders: review of findings from high-risk research,problem gamblers and genetic association studies. Neurosci. Biobehav. Rev. 32,

d prefrontal gray matter reductions in cocaine dependent individuals.

777–810.Verdejo-García, A., Lozano, O., Moya, M., Alcázar, M.A., Pérez-García, M., 2010. Psy-

chometric properties of a Spanish version of the UPPS-P impulsive behaviorscale: reliability, validity and association with trait and cognitive impulsivity. J.Pers. Assess. 92, 1532–7752.

ING Model

D

Alcoho

V

V

V

ARTICLEAD-4380; No. of Pages 7

L. Moreno-López et al. / Drug and

erdejo-García, A.J., López-Torrecillas, F., Aguilar de Arcos, F., Pérez-García, M., 2005.Differential effects of MDMA, cocaine, and cannabis use severity on distinc-tive components of the executive functions in polysubstance users: a multipleregression analysis. Addict. Behav. 30, 89–101.

Please cite this article in press as: Moreno-López, L., et al., Trait impulsivity anDrug Alcohol Depend. (2012), doi:10.1016/j.drugalcdep.2012.02.012

olkow, N.D., Wang, G.J., Telang, F., Fowler, J.S., Logan, J., Childress, A.R., Jayne, M., Ma,Y., Wong, C., 2006. Cocaine cues and dopamine in dorsal striatum: mechanismof craving in cocaine addiction. J. Neurosci. 26, 6583–6588.

olkow, N.D., Chang, L., Wang, G.J., Fowler, J.S., Ding, Y.S., Sedler, M., Logan, J.,Franceschi, D., Gatley, J., Hitzemann, R., Gifford, A., Wong, C., Pappas, N., 2001.

PRESSl Dependence xxx (2012) xxx– xxx 7

Low level of brain dopamine D2 receptors in methamphetamine abusers: asso-ciation with metabolism in the orbitofrontal cortex. Am. J. Psychiatry 158,2015–2021.

Volz, K.G., Schubotz, R.I., von Cramon, D.Y., 2003. Predicting events of

d prefrontal gray matter reductions in cocaine dependent individuals.

varying probability: uncertainty investigated by fMRI. Neuroimage 19,271–280.

Whiteside, S.P., Lynam, D.R., 2001. The five factor model and impulsivity: using astructural model of personality to understand impulsivity. Pers. Individ. Dif. 30,669–689.

[ANEXO II]

Neural Correlates of the Severity of Cocaine, Heroin, Alcohol, MDMA and Cannabis

Use in Polysubstance Abusers: A Resting-PET Brain Metabolism Study

Laura Moreno-Lópeza, Emmanuel A. Stamatakisb, Maria José Fernández-Serranoc, Manuel

Gómez-Ríod, Antonio Rodríguez-Fernándezd, Miguel Pérez-Garcíaa,e,f and Antonio

Verdejo-Garcíaa,f

a Department of Personality, Evaluation and Psychological Treatment, University of

Granada, Campus de Cartuja s/n, 18071, Spain.

b Division of Anaesthesia, School of Clinical Medicine, University of Cambridge, U.K.

c Departament of Psychology, University of Jaén, Campus Las Lagunillas s/n, 23071, Spain

d Service of Nuclear Medicine of the Hospital Virgen de las Nieves of Granada, Avda. de

las Fuerzas Armadas 2, 18014, Spain.

e Centro de Investigación Biomédica en Red de Salud Mental, CIBERSAM. University of

Granada, Spain.

f Institute of Neurosciences Federico Olóriz, University of Granada, Avda. de Madrid, s/n,

Facultad de Medicina, 18013, Spain.

Correspondence should be sent to:

Laura Moreno-López

Departamento de Personalidad, Evaluación y Tratamiento Psicológico.

Universidad de Granada. Campus de Cartuja s/n, 18071, Granada, Spain.

Phone: +34 958 242 948; Fax: +34 958 243 749;

E-mail: [email protected]

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

Abstract

Introduction: Functional imaging studies of addiction following protracted

abstinence have not been systematically conducted to look at the associations between

severity of use of different drugs and brain dysfunction. Findings from such studies may be

relevant to implement specific interventions for treatment. The aim of this study was to

examine the association between resting-state regional brain metabolism (measured with

18F-fluorodeoxyglucose Positron Emission Tomography (FDG-PET) and the severity of

use of cocaine, heroin, alcohol, MDMA and cannabis in a sample of polysubstance users

with prolonged abstinence from all drugs used. Methods: Our sample consisted of 49

polysubstance users enrolled in residential treatment. We conducted correlation analyses

between estimates of use of cocaine, heroin, alcohol, MDMA and cannabis and brain

metabolism (BM) (using Statistical Parametric Mapping voxel-based (VB) whole-brain

analyses). In all correlation analyses conducted for each of the drugs we controlled for the

co-abuse of the other drugs used. Results: The analysis showed significant negative

correlations between severity of heroin, alcohol, MDMA and cannabis use and BM in the

dorsolateral prefrontal cortex (DLPFC) and temporal cortex. Alcohol use was further

associated with lower metabolism in frontal premotor cortex and putamen, and stimulants

use with parietal cortex. Conclusions: Duration of use of different drugs negatively

correlated with overlapping regions in the DLPFC, whereas severity of cocaine, heroin and

alcohol use selectively impact parietal, temporal, and frontal-premotor/basal ganglia regions

respectively. The knowledge of these associations could be useful in the clinical practice

since different brain alterations have been associated with different patterns of execution

that may affect to the rehabilitation of these patients.

Key words: Cocaine, Heroin, Alcohol, MDMA, Cannabis, Drug severity, Positron

Emission Tomography.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

1. Introduction

Drug addiction has been associated with neuroadaptations in brain systems involved

in motivation, memory and executive control [1]. Functional neuroimaging studies have

revealed that the use of different classes of drugs is associated with dysfunctions in a range

of overlapping brain regions including ventromedial and dorsolateral prefrontal cortex

(DLPFC), anterior cingulate cortex, inferior frontal gyrus, insula, amygdala, basal ganglia

and cerebellum [2]. These dysfunctions may explain the overlapping cognitive deficits

observed in users of different drugs, including working memory, inhibitory control,

flexibility or decision-making deficits [3]. In addition, some specific effects have been

described for the association between cocaine use and the parietal cortex [4,5,6], heroin use

and the temporal cortex [7], MDMA use and occipital, hippocampus and thalamic regions

[8,9], and cannabis use and the premotor cortex [10,11]. These associations are congruent

with relatively specific neuropsychological deficits pertaining to attention and cognitive

control in cocaine users, long-term memory in opiate users, visuospatial memory in MDMA

users, and psychomotor function in cannabis users [3,7]. In spite of the available evidence,

in vivo studies have not been systematically conducted to look at associations between

severity of use of different drugs and brain dysfunction. In fact, several neuroimaging

studies have failed to detect such link or have provided controversial results (reviews in 12).

These negative findings and controversies might be linked to the impact of relevant

confounding variables. One of these is the fact that neuroimaging studies have frequently

tested drug users currently using drugs or having brief periods of abstinence (24-48 h) [13,

14]; under these conditions, the presumed link between lifetime drug use and brain

dysfunction could be masked by several other factors, including recent drug use or

psychological symptoms associated with withdrawal and short-term abstinence [2]. Another

key variable for consideration is the concurrent use of multiple types of substances, which

can introduce significant confounds in the interpretation of the data.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

The current study aimed to specifically address the association between lifetime

estimates of use of different types of drugs and brain functioning (as measured by 18F-

fluorodeoxyglucose Positron Emission Tomography (FDG-PET), statistically controlling

for the concurrent use of other drugs in a sample of polysubstance users with prolonged

abstinence from all drugs. Specifically, we examined the link between lifetime estimates of

amount and duration of use of cocaine, heroin, alcohol, MDMA and cannabis and brain

metabolism (BM). Based on previous findings we hypothesized that exposure to all

substances would be significantly associated with overlapping reductions of metabolism in

frontal, limbic, striatal and cerebellar regions, whereas cocaine and heroin use would

specifically correlate with parietal and temporal cortices respectively.

2. Methods

2.1 Participants

Forty-nine substance dependent individuals (SDI), forty-one men and eight women

with a mean age of 32.67 years, were recruited during their stay in an inpatient treatment

program at the centre �Proyecto Hombre� in Granada (Spain). This centre provides a

controlled setting for the treatment of substance use disorders. Selection criteria for

participants in this study were (i) meeting the DSM-IV criteria for substance dependence;

(ii) absence of documented comorbid mood or personality disorders, as assessed by clinical

reports; (iii) absence of documented head injury or neurological disorders; and (iv) to have

minimum abstinence duration from all drugs consumed (except for tobacco) of 15 days

before testing. In fact, SDI participants had overall much longer abstinence duration (mean

of 32.94, SD=11.25 weeks); for that reason it was possible to rule out the presence of

withdrawal symptoms or brain function alterations associated with the acute or short term

effects of the drugs. None of the participants were experiencing withdrawal symptoms as

assessed by routine medical examination or were enrolled in opioid substitution treatments

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

with methadone or other pharmacological treatments (e.g., benzodiazepines) during the

course of the assessment interview/PET evaluations.

2.2 Testing protocols and procedures

This study was approved by the Human Subjects Committee of the University of

Granada. One-step Syva urine drug screens for amphetamines, benzodiazepines, cannabis,

cocaine and opiates were conducted to confirm abstinence at time of testing. Participants

were evaluated individually in two different sessions. During the first session, the

participants were provided with information about the study and were encouraged to ask

questions to ascertain they fully understood the rational for the study and were clear on the

details for participation. The participants then provided written consent and were assessed

with the Interview for Research on Addictive Behaviours [15] in the �Proyecto Hombre�

installations�. During the second session, each patient was accompanied by a therapist to

the nuclear medicine service of the �Virgen de las Nieves� hospital, where the PET

neuroimaging session took place. This second session usually took place one week after the

first session was completed.

2.3 Tools

2.3.1 Drug Use Information

The Interview for Research on Addictive Behaviour [15] was used to examine the

severity of drug use. This interview evaluates, by means of a brief interview, the quantity

(average dosing), frequency (number of drug taking episodes per month), and duration

(years of duration) of the use of different substances that can produce physical or

psychological dependence, including cocaine, heroin, alcohol, MDMA and cannabis which

were the main drugs of choice in the present study. For every substance the participant had

actually used, the following information was requested: (i) the average dose of each target

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

drug taken in each episode of use (number of grams for cocaine and heroin, number of units

for alcohol, considering that 25 cl. of hard liquor (e.g., scotch) equals 2 units, while 25 cl. of

wine or beer equals 1 unit, number of pills for MDMA and number of cigarettes for

cannabis); (ii) the frequency of these consumption episodes per month (daily, between one

and three times upon a week, once a week, between one and three times upon a month, or

once a month); and (iii) the number of years that elapsed since the onset of use. From this

information, we obtained two indices for each of the drugs used: (i) Amount (average dose x

frequency), an index of monthly use of the drug; and (ii) Duration of drug use, measuring

number of years of exposure to the drug. The main sociodemographic and clinical features

from the sample are displayed in Table 1.

- Please insert Table 1 here-

2.3.2 PET Image Acquisition

PET scans were acquired with an ECAT/931 scanner (Siemens CTI ECAT

EXACT), at the service of nuclear medicine centre of the hospital �Virgen de las Nieves� in

Granada (Spain). For each individual, we obtained 20 minute emission scans 30 minutes

after the injection of one dose of 200 MBq of FDG administrated only after levels of

glycemia had been checked (they must be below 120 mg/dl) [16]. The subjects were

scanned with their eyes open and ears unplugged in a dimly illuminated quiet room. Raw

data were processed using an iterative reconstruction technique (OSEM method: 10

iterations, 32 subsets). Images were reoriented in transaxial, coronal and saggital planes.

For the analyses reported here the reconstructed images used had 47 axial slices each with

an in-plane resolution of 2.57x2.57 and a slice thickness of 3.38mm.

2.4 Data analysis

2.4.1 Preprocessing of PET images

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

PET images were converted from DICOM to NIFTI format and were spatially

normalized to the SPM PET template (Wellcome Department of Cognitive Neurology,

London, UK) using linear (12 affine) and non-linear normalization. An average image of all

the normalized images was obtained and this acted as a study specific PET template. In this

manner we constructed a template in the same modality, from the same scanner as the

images we wanted to normalize. The raw PET images were then spatially normalized with

linear affine and nonlinear parameters to this study-specific template. Visual inspection

revealed this optimized spatial normalization technique produced satisfactory spatial

normalization for the PET image of every volunteer.

2.4.2 Statistical Analysis

The spatially normalized PET images were smoothed with an 8mm3 isotropic

Gaussian filter and were statistically modeled using the General Linear Model in SPM5.

Linear regressions were used to carry out several analyses in order to identify brain areas

with a significant correlation between uptake values and measures of the amount and

duration of cocaine, heroin, alcohol, MDMA and cannabis use. Not all subjects used all the

drugs but since virtually all participants had regularly abused two or more substances, the

analyses were conducted on the whole group and results were interpreted in relation to

polysubstance use. In each of the analysis, we controlled for the effects of the other

substances concurrently used (e.g., analysis of severity of cocaine use controlled for the

effects of co-abuse of heroin, alcohol, MDMA and cannabis), age and years of education.

We report local maximum peaks that survive a voxel threshold of p<0.001 (uncorrected for

multiple comparisons, cluster size �100 voxels). We chose to report our findings at this

threshold based on numerous similar studies of PET imaging in drug addiction [17,18].

3. Results

The main results are summarized in Table 2.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

3.1 Amount of use

Voxel-based analyses showed negative correlations between lifetime amount of

cocaine, alcohol and cannabis use and several brain regions. The amount of cocaine used

was negatively correlated with a cluster encompassing the left inferior parietal lobule

extending to the left postcentral gyrus (Figure 1.1). The amount of alcohol used was

negatively associated with three clusters: one cluster included the left middle and superior

temporal cortex; a second cluster included the bilateral superior frontal cortex extending to

the left DLPFC and the right supplementary motor area; and a third cluster included the

right DLPFC extending to the superior frontal gyrus (Figure 1.2). The amount of cannabis

use showed a negative correlation with two clusters: one cluster included the left inferior

frontal gyrus (pars triangularis) extending to the DLPFC, and a second cluster included the

right DLPFC extending to the superior frontal cortex (Figure 1.3). We did not find

significant correlations between the amount of heroin or MDMA use and BM measures.

3.2 Duration of use

Voxel-based analyses showed negative correlations between duration of heroin,

alcohol and MDMA use and several brain regions. Duration of heroin use showed a

negative correlation with three clusters, which encompassed the right inferior, middle and

superior temporal cortex extending to the right supramarginal and inferior frontal gyrus

(Figure 1.4). Duration of alcohol use was negatively correlated with four clusters. One

cluster included the right inferior temporal cortex extending to the middle aspect of this

area. A second cluster included the right putamen and pallidum. A third cluster

encompassed the right precentral and postcentral gyrus and the fourth cluster included the

right fusiform and parahippocampal gyrus (Figure 1.5). Finally, we found negative

correlations between the duration of MDMA use and BM in three clusters: the first cluster

included the left postcentral and inferior parietal gyrus; the second cluster included the right

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

inferior frontal gyrus (pars triangularis) extending to the DLFPC; and the third cluster

included the right superior temporal pole extending to the middle part of this region (Figure

1.6).

� Please insert Table 2 here �

4. Discussion

Our findings showed that the estimates of severity of use of heroin, alcohol, MDMA

and cannabis were negatively associated with DLPFC and temporal cortex regional

metabolism in polysubstance dependent individuals with prolonged abstinence from all

these drugs. Alcohol use was further associated with lower metabolism in frontal premotor

cortex and putamen. Furthermore, severity of stimulants use (cocaine and MDMA) was

uniquely associated with inferior parietal/postcentral cortex metabolism. These associations

were established after a prolonged period of abstinence, and after controlling for multiple

confounding variables, thus surpassing the limitations of previous functional neuroimaging

studies that were not specifically designed to study prolonged abstinence, and were

confounded by important variables, such as withdrawal related symptoms or concurrent use

of multiple substances. The associations found are in agreement with our initial predictions,

and with the neuropsychological deficits typically aligned with the chronic use of these

substances; these include reduced competency in executive functions (related to all classes

of drugs), episodic memory (mainly observed on heroin, alcohol, MDMA and cannabis

users), motor control (alcohol and cannabis users), and visuospatial attention (stimulants

users) (see meta-analyses in 3,8,19,20).

As expected, the estimates of amount or duration of different classes of drugs were

negatively associated with overlapping regions within the DLPFC (BA 9, 45, 46). This

region has been linked with several of the key neuroadaptations associated with drug

addiction, including drug conditioning, loss of self-control, and stimulus-driven compulsive

behavior [2]. This finding is also in agreement with results from several neuropsychological

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

studies showing cognitive deficits on working memory, planning or inhibition processes,

which are associated with DLPFC functioning, across users of several drugs [3]. We also

found negative associations between estimates of duration of heroin, alcohol, and MDMA

use and regional metabolism on partially overlapping sections of the temporal cortex. These

associates may substantiate the links between the intensity and duration of involvement

with these drugs and the degree of memory attrition, which has been observed in

longitudinal studies with users of these substances [21,22]. It is worth mentioning that

duration of heroin use was shorter to that of cocaine and alcohol use; however, previous

brain structural findings indicate that duration of heroin use is a critical factor leading to

brain damage, even in users with periods of use below five years [23]. Surprisingly, we

failed to find significant associations between the severity of the substances used and

overlapping limbic, striatal and cerebellar regions but this result could be explained from

the fact that these regions play a greater role during current use and short-term abstinence

[24]. In addition, we found relatively specific negative correlations between amount of

cocaine use and regional metabolism in the inferior parietal cortex, and between alcohol

duration and regions involved in motor programming and control (frontal precentral and

putamen). The link between cocaine use and parietal dysfunction has been observed on

previous functional imaging studies during cognitive challenges of attention and executive

control [6,25,26]. The link between alcohol use and precentral and basal ganglia

dysfunction is in agreement with structural imaging findings showing that function of these

regions is significantly reduced in alcohol users compared to other forms of addiction [27].

Additional evidence comes from volumetric studies which associated volumetric measures

of these regions and deficient motor skills in alcoholics [28]. These more specific findings

could be used to design tailored interventions aimed to address specific aspects of frontal-

subcortical executive functions in different profiles of substance abusers.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

An important implication of this study is the fact that the link between severity of

drug use and BM is still observable in substance abusers with prolonged periods of

abstinence (an average of eight months in this sample). Therefore, quantity and duration of

drug use may still be affect brain functioning after several months of successful cessation of

drug use. This is particularly relevant in view of recent findings indicating that certain

patterns of brain dysfunction can reliably predict mid-term and long-term relapse.

Permanent brain alterations may also have important repercussions for the rehabilitation of

SDI, since these alterations can affect the ability of SDI to assimilate the contents and

participate in the activities of the rehabilitation programs [29,30]. In fact, preliminary

evidence suggests that cognitive rehabilitation techniques, such as errorless learning may be

successful in compensating chronic cognitive deficits in substance abusers [31]. Thus, the

brain networks impacted by severity of drug abuse can play a central role in clinical

outcome and relapse and should become priority targets for therapeutic interventions.

Several possible limitations of our study should be considered and addressed by

future research. Our findings were not obtained in �pure� users of each of the substances

but in mostly polysubstance users. Although this limitation is inherent to the clinical

literature on addiction (i.e., �pure� users of one single drug are quite rare among substance

abusers enrolled in treatment programs), we attempted to control for the co-abuse of other

drugs using statistics. A related confounding variable was the use of nicotine, which is

ubiquitous among drug users but was not specifically assessed or controlled for.

Furthermore, due to the use of a correlational design, this study cannot yield conclusions

about cause-effect relationships between the use of drug and resting BM. However, given

the agreement of our results with both resting state PET, activation PET/fMRI and addiction

neuropsychology studies, we could hypothesize that, at least in part, the alterations observed

in this population could be due to the quantity and duration related effects of the

consumption of the substances under consideration in this study.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

One last point to consider is that our findings could be due to premorbid brain

alterations or the results of the interaction between the premorbid alterations and the

neurotoxic effects of drug use. This question should be addressed through longitudinal

designs. A related limitation is that no comparison with a healthy control database was

performed. As we argued before, a healthy control group was not needed to test the main

predictions of our study; however, the lack of controls has stopped us from making any

assumption about the direction of these findings in relation to the healthy population.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

References

1. Koob GF, Volkow ND (2010) Neurocircuitry of addiction. Neuropsychopharmacology

35: 217-238.

2. Goldstein RZ, Volkow ND (2011) Dysfunction of the prefrontal cortex in addiction:

neuroimaging findings and clinical implications. Nat Rev Neurosci 12: 652-669.

3. Fernández-Serrano MJ, Pérez-García M, Verdejo-García A (2011) What are the specific

vs. generalized effects of drugs of abuse on neuropsychological performance? Neurosci

Biobehav Rev 35: 377-406.

4. Aron JL, Paulus MP (2007) Location, location: using functional magnetic resonance

imaging to pinpoint brain differences relevant to stimulant use. Addiction 1: 33-43.

5. Lane SD, Steinberg JL, Ma L, Hasan KM, Kramer LA, et al. (2010) Diffusion tensor

imaging and decision making in cocaine dependence. PLoS One 5: e11591.

6. Tomasi D, Goldstein RZ, Telang F, Maloney T, Alia-Klein N, et al. (2007) Widespread

disruption in brain activation patterns to a working memory task during cocaine abstinence.

Brain Res 1171: 83-92.

7. Gruber SA, Silveri MM, Yurgelun-Todd DA (2007) Neuropsychological consequences

of opiate use. Neuropsychol Rev 17: 299-315.

8. Kish SJ, Lerch J, Furukawa Y, Tong J, McCluskey T, et al. (2010) Decreased cerebral

cortical serotonin transporter binding in ecstasy users: a positron emission

tomography/[(11)C]DASB and structural brain imaging study. Brain 133: 1779-1797.

9. de Win MM, Jager G, Booij J, Reneman L, Schilt T, et al. (2008) Neurotoxic effects of

ecstasy on the thalamus. Br J Psychiatry 193: 289-296.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

10. King GR, Ernst T, Deng W, Stenger A, Gonzales RM, et al. (2011) Altered brain

activation during visuomotor integration in chronic active cannabis users: relationship to

cortisol levels. J Neurosci 31: 17923-17931.

11. Pillay SS, Rogowska J, Kanayama G, Gruber S, Simpson N, et al. (2008) Cannabis and

motor function: fMRI changes following 28 days of discontinuation. Exp Clin

Psychopharmacol 16: 22-32.

12. Martín-Santos R, Fagundo AB, Crippa JA, Atakan Z, Bhattacharyya S, et al. (2010)

Neuroimaging in cannabis use: a systematic review of the literature. Psychol Med 40: 383-

98.

13. Goldstein RZ, Alia-Klein N, Tomasi D, Carrillo JH, Maloney T, et al. (2009) Anterior

cingulate cortex hypoactivations to an emotionally salient task in cocaine addiction. Proc

Natl Acad Sci U S A 106: 9453-9458.

14. Zubieta JK, Heitzeg MM, Xu Y, Koeppe RA, Ni L, et al. (2005) Regional cerebral

blood flow responses to smoking in tobacco smokers after overnight abstinence. Am J

Psychiatry 162: 567-577.

15. Verdejo-García AJ, López-Torrecillas F, Aguilar de Arcos F, Pérez-García M (2005)

Differential effects of MDMA, cocaine, and cannabis use severity on distinctive

components of the executive functions in polysubstance users: a multiple regression

analysis. Addict Behav 30: 89-101.

16. Boellaard R, O�Doherty MJ, Weber WA, Mottaghy FM, Lonsdale MN, et al. (2010)

FDG PET and PET/CT: EANM procedure guidelines for tumour PET imaging: version 1.0.

Eur J Nucl Med Mol Imaging 37: 181-200.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

17. Fehr C, Hohmann N, Gründer G, Dielentheis TF, Buchholz HG, et al. (2007) Tiagabine

does not attenuate alcohol-induced activation of the human reward system.

Psychopharmacology (Berl) 191: 975-983.

18. Volkow ND, Fowler JS, Wang GJ, Telang F, Logan J, et al. (2010) Cognitive control of

drug craving inhibits brain reward regions in cocaine abusers. Neuroimage 49: 2536-2543.

19. Jovanovski D, Erb S, Zakzanis KK (2005) Neurocognitive deficits in cocaine users: a

quantitative review of the evidence. J Clin Exp Neuropsychol 27: 189-204.

20. Zakzanis KK, Campbell Z, Jovanovski D (2007) The neuropsychology of ecstasy

(MDMA) use: a quantitative review. Hum Psychopharmacol 22: 427-435.

21. Hanson KL, Cummins K, Tapert SF, Brown SA (2011) Changes in neuropsychological

functioning over 10 years following adolescent substance abuse treatment. Psychol Addict

Behav 25: 127-142

22. de Sola S, Tarancón T, Peña-Casanova J, Espadaler JM, Langohr K, et al. (2008)

Auditory event-related potentials (P3) and cognitive performance in recreational ecstasy

polydrug users: evidence from a 12-month longitudinal study. Psychopharmacology (Berl)

200: 425-437

23. Yuan Y, Zhu Z, Shi J, Zou Z, Yuan F, et al. (2009) Gray matter density negatively

correlates with duration of heroin use in young lifetime heroin-dependent individuals. Brain

Cogn 71: 223-228.

24. Breiter HC, Rosen BR (1999) Functional magnetic resonance imaging of brain reward

circuitry in the human. Ann N Y Acad Sci 877: 523-547.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

25. Barrós-Loscertales A, Bustamante JC, Ventura-Campos N, Llopis JJ, Parcet MA, et al.

(2011) Lower activation in the right frontoparietal network during a counting Stroop task in

a cocaine-dependent group. Psychiatry Res 194: 111-118.

26. Bustamante JC, Barrós-Loscertales A, Ventura-Campos N, Sanjuán A, Llopis JJ, et al.

(2011) Right parietal hypoactivation in a cocaine-dependent group during a verbal working

memory task. Brain Res 1375: 111-119.

27. van Holst RJ, de Ruiter MB, van den Brink W, Veltman DJ, Goudriaan AE (2012) A

voxel-based morphometry study comparing problem gamblers, alcohol abusers, and healthy

controls. Drug Alcohol Depend. In press.

28. Sullivan EV, Deshmukh A, De Rosa E, Rosenbloom MJ, Pfefferbaum A (2005) Striatal

and forebrain nuclei volumes: contribution to motor function and working memory deficits

in alcoholism. Biol Psychiatry 57: 768-776.

29. Johnson-Greene D, Adams KM, Gilman S, Junck L (2002) Relationship between

neuropsychological and emotional functioning in severe chronic alcoholism. Clin

Neuropsychol 16: 300-309.

30. Zinn S, Stein R, Swartzwelder HS (2004) Executive functioning early in abstinence

from alcohol. Alcohol Clin Exp Res 28: 1338-1346.

31. Pitel AL, Perruchet P, Vabret F, Desgranges B, Eustache F, et al. (2009) The advantage

of errorless learning for the acquisition of new concepts' labels in alcoholics. Psychol Med

40: 497-502.

Tzourio-Mazoyer N, Landeau B, Papathanassiou D, Crivello F, Etard O, et al. (2002)

Automated anatomical labeling of activations in SPM using a macroscopic anatomical

parcellation of the MNI MRI single-subject brain. Neuroimage 15: 273-289.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

Figure Legends

Figure 1. Areas of reduced metabolism predicted by the amount or duration of drug

use: (1) amount of cocaine, (2) amount of alcohol, (3) amount of cannabis (4) duration of

heroin (5) duration of alcohol and (6) duration of MDMA. Areas of reduced metabolism are

superimposed on a T1 weighted MRI image in MRICRON. MNI coordinates are shown

underneath each panel.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

Tables

Table 1 Sociodemographic data and clinical features of the participants

Subjects Characteristics

Gender

Male 41

Female 8

Mean S.D. Maximum Minimum

Age (years) 32.67 8.03 53 21

Education (years) 9.73 2.34 17 5

Amount of drug usea

Substance Mean S.D. Maximum Minimum

Cocaine (g) (46) 48.69 44.83 180 0

Heroin (g) (16) 9.12 20.63 120 0

Alcohol (units) (45) 571.24 489.34 1800 0

MDMA (pills) (23) 13.41 24.31 120 0

Cannabis (cigarettes) (37) 148.57 190.64 750 0

Duration of drug use (years)

Substance Mean S.D. Maximum Minimum

Cocaine 7.95 5.95 23 0

Heroin 1.76 3.69 17 0

Alcohol 10.85 7.67 27 0

MDMA 1.40 2.28 8 0

Cannabis 8.31 8.13 29 0

Abstinence (weeks) 32.94 11.25 60 12

a Amount of monthly use of the drug calculated using the average dose (per episode) ×

frequency (per month).

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

Table 2 Relation between severity of drug use and PET uptake.

Substanc Voxels-Based BA KE T MNI Amount

Cocaine Parietal Inf L 40 176 3.98 -58 -38 54

Alcohol Temporal Mid L 21 267 4.04 -66 -2 -10

Frontal Sup L 9 308 3.96 -20 30 36

Frontal Sup R 6 365 3.73 22 2 60

Frontal Mid R 46 372 3.82 26 22 34

Cannabis

Frontal Inf Tri L 45 355 4.40 -44 22 30

Frontal Mid R 46 125 4.02 24 48 28

Duration

Heroin Frontal Mid R 46 876 3.77 44 36 34

Temporal Inf R 20 257 4.58 62 -16 -32

Temporal Sup R 22 1090 4.19 66 -44 20

Alcohol Temporal Inf R 20 238 4.22 64 -16 -30

Putamen R 472 3.71 20 -2 8

Precentral R 6 274 3.67 50 2 34

Fusiform R 20 299 3.65 30 -6 -40

MDMA Postcentral L 40 131 3.73 -28 -40 48

Frontal Inf Tri R 45 139 3.67 44 34 20

Temporal Pole Sup R 20 561 3.67 32 12 -30

Voxel-based analyses at p<0.001 (uncorrected, cluster size > 100 voxels). Labels were

obtained using the aal toolbox (Tzourio-Mazoyer et al., 2002). Broadmann areas were

obtained using MRICron (Rorden and Brett, 2000). KE indicate the number of voxels

included in the cluster. Stereotaxic coordinates are those listed in SPM5. Inf, inferior;

Sup, superior; L, left; R, right; Supp, supplementary; Mid, middle; Orb, orbitalis.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

[ANEXO III]

1Q2

2

3

4Q1

5

678910

11

1213141516171819202122232425

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

Psychiatry Research: Neuroimaging xxx (2012) xxx–xxx

PSYN-09914; No of Pages 8

Contents lists available at SciVerse ScienceDirect

Psychiatry Research: Neuroimaging

j ourna l homepage: www.e lsev ie r .com/ locate /psychresns

RO

OF

Neural correlates of hot and cold executive functions in polysubstance addiction:Association between neuropsychological performance and resting-PETbrain metabolism

Laura Moreno-López a,⁎, Emmanuel Andreas Stamatakis b, María José Fernández-Serrano a,Manuel Gómez-Río c, Antonio Rodríguez-Fernández c, Miguel Pérez-García a,d,e, Antonio Verdejo-García a,d

a Department of Personality, Evaluation and Psychological Treatment, University of Granada, Spainb Division of Anaesthesia, School of Clinical Medicine, University of Cambridge, UKc Service of Nuclear Medicine of the Hospital Virgen de las Nieves of Granada, Spaind Institute of Neurosciences Federico Olóriz, University of Granada, Spaine Centro de Investigacion Biomédica en Red de Salud Mental, CIBERSAM, University of Granada, Spain

⁎ Corresponding author at: Departamento de Persmiento Psicológico, Universidad de, Granada, Campus dSpain. Tel.: +34 958 242 948; fax: +34 958 243 749.

E-mail address: [email protected] (L. Moreno-López).

0925-4927/$ – see front matter © 2012 Elsevier Irelanddoi:10.1016/j.pscychresns.2012.01.006

Please cite this article as: Moreno-López, L.ation between neuropsychological perform

Pa b s t r a c t

a r t i c l e i n f o

26

27

28

29

30

31

32

33

34

35

Article history:Received 7 April 2010Received in revised form 20 October 2011Accepted 16 January 2012Available online xxxx

Keywords:Brain substratesDecision-makingExecutive functioningPositron emission tomography

36

37

38

39

40

41

RECTEDThe study of substance abuse related neuropsychological deficits and brain alterations may provide a better

understanding of the neuroadaptations associated with addiction. In this study we investigated the associa-tion between performance on neuropsychological tests of cold and hot executive functions and regional brainmetabolism (measured with Positron Emission Tomography – PET) in a sample of 49 substance dependentindividuals (SDI). Neuropsychological performance in the SDI group was compared to that of a non-drugusing control group of 30 participants, and associated with two sets of PET-derived dependent measures:one based on regions of interest (examining mean uptake on selected regions), and a second based onvoxel uptake measures (using Statistical Parametric Mapping voxel-based whole-brain analyses). Behavioralanalyses showed that SDI had poorer performance than controls across executive function and emotion pro-cessing measures. Regression models showed that SDI's performance in cold executive tests was associatedwith regional metabolism in the dorsolateral prefrontal cortex (DLPFC), mid-superior frontal gyrus, superiorand inferior temporal gyrus and inferior parietal cortex, whereas performance in hot executive functions wasassociated with DLPFC, mid-superior frontal gyrus, anterior and mid-posterior cingulate, and temporal andfusiform gyrus. These results are discussed in terms of their relevance for the understanding of cognitive dys-function and neuroadaptations linked to addiction.

© 2012 Elsevier Ireland Ltd. All rights reserved.

4243

R

59

60

61

62

63

64

65

66

67

68

69

70

71

72

UNCO1. Introduction

Contemporary neurobiological models consider addiction to be abrain disorder that involves long-term neuroadaptations leading topersistent drug use despite increasing negative consequences(Goldstein and Volkow, 2002; Verdejo-García et al., 2004; Baler andVolkow, 2006). These neuroadaptations are thought to affect neuralsystems involved in motivation, emotion, learning, memory and ex-ecutive functioning (Ersche et al., 2008; Verdejo-García and Bechara,2009). From this perspective, it is crucial to understand both the long-term neuropsychological consequences of drug abuse and their neu-ral substrates.

There is compelling evidence that substance dependent individ-uals (SDI) present widespread deficits in neuropsychological

73

74

75

76

77

onalidad, Evaluación y Trata-e Cartuja S/N, 18071, Granada,

Ltd. All rights reserved.

, et al., Neural correlates of hance..., Psychiatry Research:

functioning, and that these deficits are especially prominent in so-called executive functions (Fernández-Serrano et al., 2011). Executivefunctions are defined as a set of higher-order abilities involved in theemployment, monitoring, and regulation of goal-directed behaviors.Recent evidence suggests that these functions can be classified intotwo broad domains: (i) cold executive functions, involved in the pro-cessing of relatively abstract, context-free information, which arelinked to more dorsal and lateral regions of the prefrontal cortex,and (ii) hot executive functions, which come to play when emotion-ally laden information is concerned, and are linked to the functioningof the orbitofrontal cortex (OFC) (Kerr and Zelazo, 2004).

Although executive dysfunction in SDI has been extensively de-scribed, the study of the neurobiological substrates of executive dys-function in the context of drug addiction is still a growing researcharea (Goldstein and Volkow, 2002; Lundqvist, 2010). A number of func-tional magnetic resonance imaging (fMRI) studies have shown signifi-cant associations between performance on cognitive tasks indexingcold aspects of executive functions and abnormal activation of brainsystems in SDI. Polysubstance abusers have shown abnormally

ot and cold executive functions in polysubstance addiction: Associ-Neuroimaging (2012), doi:10.1016/j.pscychresns.2012.01.006

Original text:
Inserted Text
"givenname"
Original text:
Inserted Text
"givenname"
Original text:
Inserted Text
"givenname"
Original text:
Inserted Text
"givenname"
Original text:
Inserted Text
"givenname"
Original text:
Inserted Text
"givenname"
Original text:
Inserted Text
"givenname"
Original text:
Inserted Text
"surname"
Original text:
Inserted Text
"surname"
Original text:
Inserted Text
"surname"
Original text:
Inserted Text
"surname"
Original text:
Inserted Text
"surname"
Original text:
Inserted Text
"surname"
Original text:
Inserted Text
"surname"
Original text:
Inserted Text
"SDI's "

TE

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125126127128129130131132133134135136137138139140141142143144

145146147148149150151152153154155156157

158

159160161162163164165166167

168

169170171172173174175

176177178179180181182183184185186187188189

190191192193194195196197198199

Table 1 t1:1

Sociodemographic data and clinical features of the participants.t1:2t1:3SDI (n=49) HC (n=30)

t1:4Variable N % N %

t1:5Gendert1:6Male 41 84 24 80t1:7Female 8 16 6 20t1:8Mean S.D. Mean S.D.t1:9Age (years) 32.67* 8.03 26.4* 8.03t1:10Education (years) 9.73* 2.34 11.63* 2.04

* p≤0.001. t1:11

2 L. Moreno-López et al. / Psychiatry Research: Neuroimaging xxx (2012) xxx–xxx

UNCO

RREC

increased or decreased activation of the dorsolateral prefrontal cortex(DLPFC), orbitofrontal cortex (OFC), temporo-parietal regions and thecerebellum during performance on working memory (Desmond et al.,2003; Jager et al., 2006; Tomasi et al., 2007), response inhibition (Dao-Castellana et al., 1998; Bolla et al., 2004; Eldreth et al., 2004; Formanet al., 2004; Lee et al., 2005), and cognitive flexibility tasks (Gilman etal., 1990; Goldstein et al., 2004). Nevertheless, very few studies have in-vestigated the neural correlates of indices of hot executive functions inSDI. These abilities have been associatedwith the OFC and the cingulatecortex in cognitive neuroscience studies (Anderson et al., 2006;Heberlein et al., 2008; Hartstra et al., 2010), and they have been foundto be dysfunctional in polysubstance abusers even aftermid-term absti-nence (circa sixmonths) (Fernández-Serrano et al., 2011). For example,two neuropsychological studies have revealed deficits in goal-drivenself-regulation in psychostimulant users using the Revised Strategy Ap-plication Test (R-SAT) (Halpern et al., 2004; Verdejo-García et al., 2007),but no studies to date have investigated the neural substrates of self-regulation in addiction. Another key factor for addiction initiation andmaintenance is adaptive decision-making, which is typically indexedusing the Iowa Gambling Task (IGT) (Bechara et al., 1994). Performanceon this task has been associatedwith hyperperfusion in the anterior cin-gulate cortex (ACC), middle and superior frontal gyrus and DLPFC insmall samples of shortly abstinent cocaine-dependent individuals(Adinoff et al., 2003; Tucker et al., 2004). However, because of the keyrole of decision-making in mid- and long-term drug relapse (Paulus etal., 2005; Passetti et al., 2008), there is a need to extend these findingsin larger samples ofmid-term abstinent SDI. Similarly, the neural corre-lates of emotional processing and perception must be considered inlight of their association with hot aspects of executive functions in SDI(Verdejo-García et al., 2007), and their predictive role in drug relapse(Lubman et al., 2009). Only two previous neuroimaging studies have fo-cused on this domain, and they have found decreased ACC and amygda-la activations in alcohol users exposed to facial emotional expressions(Salloum et al., 2007; Marinkovic et al., 2009).

In this study we aimed to investigate the neuropsychological per-formance of a sample of SDI and healthy controls (HC) on tests tap-ping on cold (i.e., updating, inhibition and flexibility) and hot (i.e.,self-regulation, decision-making and emotion perception) aspects ofexecutive functions, and the association between the neuropsycho-logical performance of SDI on those tests and PET-indexed regionalmetabolism. We hypothesize that neuropsychological performancewill be poorer in SDI than HC and that it will be associated withDLPFC, ACC and temporal regions for tests of cold executive functions,and with OFC, ACC and limbic regions for tests of hot executivefunction.

2. Methods

2.1. Participants

Forty-nine SDI were recruited from an inpatient treatment program at the centre“Proyecto Hombre” in Granada (Spain). The treatment has a mean duration of 6 monthsand integrates key elements of cognitive behavioral psychology and systemic approachesto provide complete social rehabilitation and integration. Selection criteria for participantsin this studywere (1) meeting the DSM-IV criteria for substance dependence at the onsetof the treatment; (2) absence of documented comorbid mood or personality disorders, asassessed by clinical reports; (3) absence of documented head injury or neurological disor-ders, as assessed by clinical reports and PET images; (4) to be literate enough to performany literacy tests, as assessed by the TAP reading skills test (Del Ser et al., 1997) and (5) tohave aminimum abstinence duration of 15 days before testing. In fact, the SDI included inthe sample had overall much longer abstinence duration (M=32.94, S.D.=11.25 weeks),allowing us to rule out acute or short term effects of drug intake. Since virtually all partic-ipants had regularly abused two or more substances, the analyses were conducted on thewhole group and results were interpreted in relation to polysubstance use. Urine toxicol-ogy screening (one-step Syva rapid tests) for the different drugs used were conductedroutinely in the SDI, sowe can rule out the use of these substances during the entire periodof abstinence. None of the participants were experiencing withdrawal symptoms beforeor during neuropsychological/PET testing as assessed by routine medical examination.Likewise, none of the participants were enrolled in opioid substitution treatments withmethadone or other pharmacological treatments during the course of the

Please cite this article as: Moreno-López, L., et al., Neural correlates of hation between neuropsychological performance..., Psychiatry Research:

neuropsychological/PET testing. A group of drug-free healthy individuals was recruitedas a comparison group for neuropsychological performance, but these participants werenot PET-scanned. HC were recruited through local advertisements and snowball commu-nication among adult people from the community. Selection criteria for these groupwere:(1) absence of current or past substance abuse, excluding past cannabis use and past orcurrent smoking or social drinking (in cases of current social drinking, exclusion criteriaincluded acute use of alcohol during the last 24 h before evaluation), (2) absence of docu-mentedmajor psychiatric disorders, (3) absence of documented head injury or neurolog-ical disorder and (4) not being on any medication that could affect the normalneuropsychological functioning. This group only had exposure to alcohol use, having aregular mean use of 8.85 standard units per week (S.D.=20.7) and a mean duration ofuse of 6.7 years (S.D.=7.3). The main sociodemographic and clinical features from thesamples are displayed in Table 1.

OO

F

2.2. Testing protocols and procedures

This study was approved by the Human Subjects Committee of the University ofGranada. All participants signed an informed consent form before participating in thestudy. The testing protocol, including the neuropsychological assessment and the PETscanning were administered into two sessions that lasted approximately 70 mineach. The control group only participated in the first session. During the first session,participants completed the Interview for Research on Addictive Behaviors and per-formed the neuropsychological tests protocol. During the second session (up to7 days later), SDI were scanned at the nuclear medicine service of the “Virgen de lasNieves” hospital.

D P

R2.3. Instruments

2.3.1. Patterns of drug useThe information about the severity of drug addiction was collected by the Interview

for Research on Addictive Behavior (IRAB; López-Torrecillas et al., 2001) (see supple-mentary material). This is a brief interview that collects information about the quantity(average dose), frequency (consumption episodes by month), and duration (years ofduration) of the use of a series of drugs of abuse, including cocaine, heroin, and alcohol.The main drug use information of the SDI group is displayed in Table 2.

2.3.2. Neuropsychological testsWe administered a battery of tests designed to assess several domains related to

cold and hot executive functions, including working memory, response inhibition, cog-nitive flexibility, self-regulation, decision-making and emotion processing.

The tests used were Letter Number Sequencing (LNS) (Wechsler Adult IntelligenceScale – WAIS-III, Wechsler, 1997), Stroop (Golden, 1978), Category Test (DeFilippis,2002), Revised Strategy Application Test (R-SAT) (Levine et al., 2000), Iowa GamblingTask (IGT) (Bechara et al., 1994) and Ekman Faces Test (EFT) (Young et al., 2002). Alltests were administered in a fixed order and according to standard instructions. Thedescription of the instruments is provided in the supplementary material.

Due to technical problems during data acquisition, we could not obtain the perfor-mance indices for some of the tests in the SDI group. Therefore, the number of SDI par-ticipants included in the analyses of LNS was 40, 39 for Stroop, 47 for Category Test, 49for R-SAT and IGT, and 48 for EFT.

2.3.3. PET image acquisitionPET scans were acquired with the ECAT/931 scanner (Siemens CTI ECAT EXACT), at

the nuclear medicine facility of the “Virgen de las Nieves” hospital in Granada, (Spain).For each individual, we obtained 20 minute emission scans 30 min after injection ofFDG resting in quiet conditions. The subjects were scanned with their eyes open andears unplugged in a dimly illuminated quiet room. Raw data were processed usingan iterative reconstruction (OSEM method: 10 iterations, 32 subsets). Images werereoriented in transaxial, coronal and saggital planes. For the analyses reported herethe reconstructed images used had 47 axial slices each with an in-plane resolution of2.57×2.57 and a slice thickness of 3.38 mm.

ot and cold executive functions in polysubstance addiction: Associ-Neuroimaging (2012), doi:10.1016/j.pscychresns.2012.01.006

Original text:
Inserted Text
"x"
Original text:
Inserted Text
"SD"
Original text:
Inserted Text
"SD"
Original text:
Inserted Text
"SD"
Original text:
Inserted Text
"70 "
Original text:
Inserted Text
"minutes "
Original text:
Inserted Text
"- "
Original text:
Inserted Text
" minutes"
Original text:
Inserted Text
"n ="
Original text:
Inserted Text
"SD"
Original text:
Inserted Text
"p ≤"

T200

201202203204205206207208209210211

212213214215216

217

218219220221222223

224225226227228229230

231232233234235236237238239240241242243244245246

247248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

Table 2t2:1

Amount and duration of drug use in the substance dependence individuals.t2:2t2:3 Quantity of use

t2:4 Substance Averagea S.D. Maximum Minimum

t2:5 Cocaine (46) 48.69 44.83 180 0t2:6 Cannabis (37) 148.57 190.64 750 0t2:7 Heroin (16) 9.12 20.63 120 0t2:8 MDMA (18) 13.41 24.31 120 0t2:9 Alcohol (45) 571.24 489.34 1800 0

t2:10 Duration of use (years)

t2:11 Substance Duration S.D. Maximum Minimum Age ofonset

t2:12 Cocaine 7.95 5.95 23 0 23.26t2:13 Cannabis 8.31 8.13 29 0 20.91t2:14 Heroin 1.76 3.69 17 0 24.75t2:15 MDMA 1.40 2.28 8 0 23.66t2:16 Alcohol 10.85 7.67 27 0 20.87t2:17 Abstinenceb 32.94 11.25 60 12

a Average dose (per episode)×Frequency (per month). A dose is units in alcohol,grams in cocaine and heroin, fags in cannabis and pills in MDMA.t2:18

b Abstinence in weeks. The number in brackets is the number of substance depen-dence individuals that consumed the substance.t2:19

Table 3 t3:1

Descriptive scores on the different indices of executive functioning.t3:2t3:3Variable Group p-valuea F-value d-value

t3:4SDI HC

t3:5LNSb 9.75±2.27 15.13±2.33 0.000* 32.12 −2.34t3:6Stroopc −1.76±5.76 4.15±7.24 0.000* 7 −0.92t3:7Category testd 65.67±26 31.80±25.33 0.000* 11.48 1.32t3:8R-SATe 83.78±17.14 93.98±5.71 0.013* 3.83 −0.73t3:9IGTf −1.98±21.06 37.20±26.16 0.000* 17.90 −1.70t3:10Ekman faces testg 47.50±4.49 53±4.61 0.000* 10.03 −1.21

a pb0.05. t3:11b Letter and number sequence total score. t3:12c Stroop interference score. t3:13d Category test total errors. t3:14e Revised strategy application test percentage brief items. t3:15f Iowa gambling test total score. t3:16g Ekman faces test total identifications. t3:17

3L. Moreno-López et al. / Psychiatry Research: Neuroimaging xxx (2012) xxx–xxx

UNCO

RREC

2.4. Data analysis

2.4.1. Preprocessing of PET imagesThe PET images were converted from DICOM to NIFTI format, were spatially nor-

malized to the SPM5 PET template (Wellcome Department of Cognitive Neurology,London, UK) using linear (12 affine) and non-linear normalization. An average of allthe normalized images was obtained and this acted as a study specific PET template.In this manner we constructed a template in the samemodality, from the same scanneras the images we wanted to normalize. The raw PET images were then spatially nor-malized with linear affine and nonlinear parameters to this study-specific template. Vi-sual inspection revealed this optimized spatial normalization technique producedsatisfactory spatial normalization for the PET image of every volunteer. Finally, the im-ages were smoothed with an 8 mm3 isotropic Gaussian filter.

2.4.2. ROI analysisHaving achieved satisfactory spatial normalisation we used the Marsbar tool

(http://marsbar.sourceforge.net- Brett et al., 2002) to extract 90 regions of interest(ROIs) in the entire brain. We used the AAL atlas as a reference for these ROIs(Tzourio-Mazoyer et al., 2002).

2.4.3. Statistical analyses

2.4.3.1. Behavioral analysis. ANCOVAmodels were conducted to test neuropsychologicalperformance differences between the groups while controlling for age and educationin SPSS 15.0 for Windows (SPSS Inc., Chicago IL). Cohen's d estimations of effect sizeswere calculated according to formulations proposed in (Zakzanis, 2001). The associa-tion between length of abstinence and neuropsychological performance was exploredusing bivariate correlations in SPSS 15.0.

2.4.3.2. ROI analysis. Relative glucose uptake on standardized uptake value (SUV) wereused to obtain the mean uptake values for each ROI. Next, these values were correlatedwith the indices of neuropsychological performance. In a second analysis, those ROIsthat showed significant associations with neuropsychological indices were also corre-lated with estimates of severity of drug use. In both cases, we used partial correlationsand age as confounding variable in SPSS 15.0. Results are showed at p≤0.01 andp≤0.05 respectively (uncorrected for multiple comparisons).

2.4.3.3. Voxel based analysis. Linear regressions were used to identify clusters with a sig-nificant correlation between uptake values and indices of neuropsychological perfor-mance using age as confounding factor. Scans were corrected for differences in globalactivity by including proportional scaling in the regression model. We report localmaximum peaks that survive a voxel threshold of pb0.001 (uncorrected for multiplecomparisons, cluster size>100 voxels). Although this threshold offers only modestprotection again risk of type I error, it has been chosen in multiples PET studiesunder similar assumptions (i.e. Brooks et al., 2009; Yaouhi et al., 2009). The relation-ship between duration of abstinence and resting glucose metabolism was exploredby using linear regression models but no associations survived at a threshold ofpb0.001, and therefore it was not further considered. We did not use treatment seek-ing as a variable in our analyses because all our volunteers were on treatment, andmore specifically on the same modality of treatment (i.e., therapeutic community). Fi-nally, we carried out a second analysis in which the mean uptake of the regions wefound to be associated with neuropsychological performance in whole brain voxel-based analysis were correlated with estimates of drug use using age as confounding

Please cite this article as: Moreno-López, L., et al., Neural correlates of hation between neuropsychological performance..., Psychiatry Research:

ED P

RO

OF

variable in SPSS 15.0. Results are showed at p≤0.01 and p≤0.05 respectively (uncor-rected for multiple comparisons).

3. Results

3.1. Neuropsychological performance

Results showed that SDI had significantly poorer performance thanHC on all of the neuropsychological tests administered (Table 3). Calcu-lation of effect sizes showed large effect sizes (d>0.8) for LNS (workingmemory), Stroop (response inhibition), Category Test (cognitive flexi-bility), IGT (decision-making) and EFT (emotion processing), andmedi-um effect sizes (d>0.5) for R-SAT (self-regulation) (Table 3). We onlyfound a significant negative correlation between length of abstinenceand performance on LNS (r=−0.303).

3.2. ROI analyses

3.2.1. Association between neuropsychological performance and ROIsThese results are shown in Table 4. We found significant correla-

tions between number of hits on LNS (working memory) and cerebralmetabolism (CM) at the right middle temporal pole (r=−0.367), be-tween R-SAT proportion of brief items (self-regulation) at the right cal-carine (r=0.391) and posterior cingulum bilaterally (left r=0.440)(right r=0.454), and between IGT net score (decision-making) and theright middle (r=0.390) and superior frontal cortex (r=0.376). Therewere not significant correlations between the remaining tests and theROIs at p≤0.01.

3.2.2. Association between the ROIs associated with neuropsychologicalperformance and estimates of drug use

We failed to find significant associations between measures of se-verity of drug use and cerebral metabolism at a pb0.01 threshold. Byrelaxing the alpha level to pb0.05, we only found a positive correla-tion between amount of cocaine use and brain metabolism in theright middle temporal pole (r=0.301).

3.3. Voxel-based analyses

The main results from these analyses are summarized in Table 4.The results presented in text and table correspond to an alpha levelof pb0.001 (uncorrected, cluster size>100 voxels), but for demon-stration purposes the figures display statistical parametric maps thre-sholded at uncorrected voxel level of pb0.01.

3.3.1. Cold executive functions –Whole brain voxel-based correlates in SDIFor working memory, there was a positive correlation between

LNS and one cluster with peaks at the superior temporal gyrus(51–29 3) and inferior temporal gyrus (51–72 0); this cluster

ot and cold executive functions in polysubstance addiction: Associ-Neuroimaging (2012), doi:10.1016/j.pscychresns.2012.01.006

Original text:
Inserted Text
"a Average dose (per episode) × Frequency (per month). A dose is units in alcohol, grams in cocaine and heroin, fags in cannabis and pills in MDMA. b Abstinence in weeks. The number in brackets is the number of substance dependence individuals that consumed the substance."
Original text:
Inserted Text
"b"
Original text:
Inserted Text
"c"
Original text:
Inserted Text
"d"
Original text:
Inserted Text
"e"
Original text:
Inserted Text
"f"
Original text:
Inserted Text
"g"
Original text:
Inserted Text
"´"
Original text:
Inserted Text
"p ≤"
Original text:
Inserted Text
"p <"
Original text:
Inserted Text
"100 "
Original text:
Inserted Text
"p <"
Original text:
Inserted Text
"p ≤"
Original text:
Inserted Text
"p ≤"
Original text:
Inserted Text
"p ≤"
Original text:
Inserted Text
"p <"
Original text:
Inserted Text
"p <"
Original text:
Inserted Text
"p <"
Original text:
Inserted Text
" voxel"
Original text:
Inserted Text
"p <"
Original text:
Inserted Text
"Valuea"
Original text:
Inserted Text
"Value"
Original text:
Inserted Text
"Value"

RO

OF

287

288

289

290

291

292

293

294

295

296

Table 4t4:1

Relationship between neuropsychological performance and PET uptake using both ROI and voxel based methods.t4:2t4:3 Instrument ROIs Voxel-based analysis BA KE t p MNI Talairach

t4:4 LNS R middletemporal pole (+) R superior temporal gyrus (+) BA22 621 5.23 0.000 52 −30 2 51 −29 3t4:5 R inferior temporal gyrus (+) BA37 261 3.77 0.000 52 −74 −4 51 −72 0t4:6 L paracentral lobule (−) BA5 502 4.67 0.000 −20 −38 52 −20 −34 50t4:7 L superior frontal gyrus (−) BA8 449 3.95 0.000 −20 28 52 −20 30 46t4:8 Stroop L pallidus (−) 191 3.82 0.000 −16 −12 −4 −16 −12 −3t4:9 L middle frontal gyrus (−) BA10 160 3.72 0.000 −34 64 10 −34 62 6t4:10 Category test L superior frontal gyrus (+) BA10 127 4.25 0.000 −26 68 4 −26 66 0t4:11 R middle frontal gyrus (−) BA6 268 3.59 0.000 50 10 48 50 12 44t4:12 L inferior parietal lobule (−) BA40 198 3.74 0.000 −46 −48 38 −46 −45 37t4:13 R-SAT R calcarine (+)

R Posterior cingulate (+)L Posterior cingulate (+)

R posterior cingulate (+) BA30 5811 4.12 0.000 18 −68 10 18 −65 12t4:14 R anterior cingulate (+) BA32 702 3.75 0.000 14 32 16 14 32 13t4:15 R precentral gyrus (+) BA6 416 3.96 0.000 22 −20 64 22 −16 60t4:16 R inferior temporal gyrus (+) BA20 152 3.58 0.000 40 2 −48 40 −0 −40t4:17 L middle temporal gyrus (−) BA21 464 4.48 0.000 −56 10 −20 −55 9 −17t4:18 IGT R middle frontal (+)

R superior frontal (+)R middle frontal gyrus (+) BA6 945 4.91 0.000 26 −6 60 26 −3 55

t4:19 L middle frontal gyrus (+) BA6 377 3.71 0.000 −28 2 54 −28 4 50t4:20 R middle frontal gyrus (+) BA6 222 3.94 0.000 54 −12 56 53 14 51t4:21 L superior frontal gyrus (+) BA10 212 3.67 0.000 −18 68 14 −18 67 10t4:22 L superior frontal gyrus (+) BA8 141 3.72 0.000 −20 44 52 −20 45 46t4:23 Ekman Faces Test Lingual gyrus (−) BA18 308 4.10 0.000 0 −102 −14 0 −99 −7t4:24 Fusiform gyrus (−) BA19 191 3.90 0.000 −36 −78 −18 −36 −76 −11

L, left; R, right; (+) positive association; (−) negative association. Correlation ROI analyses are reported at p≤0.01 (uncorrected). Voxel-based analyses at pb0.001 (uncorrected,cluster size>100 voxels). Talairach coordinates were obtained by applying the mni2tal function in MATLABR2006a (http://imaging.mrc-cbu.cam.ac.uk/imaging/MniTalairach) tothe MNI coordinates obtained from SPM5. The corresponding anatomical nameswere obtained using the TalairachClient application – version 2.4.2. – (http://www.talairach.org/client.html)for peak and nearest grey matter.t4:25

4 L. Moreno-López et al. / Psychiatry Research: Neuroimaging xxx (2012) xxx–xxx

included right superior and inferior lateral temporal regions extend-ing medially. LNS negatively correlated with a cluster in the left post-central region extending to parietal cortex, precuneus andparacentral lobule (peak at paracentral lobule, −20 −34 50), and asecond cluster which encompassed regions in the left frontal cortex

UNCO

RRECTA) Letter Number sequencing negative (b

L -20 52

B) Stroop negative correlations (blue).

L -34 -16

C) Category Test negative (blue) and pos

L -45 -26

Fig. 1. Images showing the areas where cold execut

Please cite this article as: Moreno-López, L., et al., Neural correlates of hation between neuropsychological performance..., Psychiatry Research:

D Pextending to the left DLPFC (peak at superior frontal gyrus, −20 30

46) (Fig. 1 – Panel A).For response inhibition, there was a negative correlation between

the Stroop interference score and two different clusters. The first clus-ter correspond to the left thalamus, hippocampus and pallidum (peak

E

lue) and positive (yellow) correlations.

R

R

itive (yellow) correlations.

45 R

ive functions correlated with brain metabolism.

ot and cold executive functions in polysubstance addiction: Associ-Neuroimaging (2012), doi:10.1016/j.pscychresns.2012.01.006

Original text:
Inserted Text
"a p < 0.05. b Letter and Number sequence total score. c Stroop interference score. d Category test total errors. e Revised strategy application test percentage brief items. f Iowa gambling test total score. g Ekman faces test total identifications."
Original text:
Inserted Text
"-"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
"p ≤"
Original text:
Inserted Text
"p <"
Original text:
Inserted Text
" voxel"

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

5L. Moreno-López et al. / Psychiatry Research: Neuroimaging xxx (2012) xxx–xxx

at globus pallidus, −16 −12 −3), and the second cluster encom-passed the DLPFC bilaterally extending to the superior frontal cortex(peak at middle frontal gyrus, −34 62 6) (Fig. 1 – Panel B).

For cognitive flexibility, results showed a positive correlation be-tween Category Test total number of errors and a cluster which in-cluded the DLPFC and superior frontal gyrus (peak at superiorfrontal gyrus, −26 66 0). There were also negative correlations be-tween task performance and two other clusters. The first was at theDLPFC extending to the precentral gyrus with peak at middle frontalgyrus (50 12 44), and the second in the parietal cortex extending tothe supramarginal and angular area (peak at inferior parietal lobule,−46 −45 37) (Fig. 1 –Panel C).

In summary, workingmemory and cognitive flexibility performancenegatively correlated with cerebral metabolism in the left DLPFC andleft parietal regions, whereas response inhibition performance nega-tively correlated with bilateral DLPFC and basal ganglia. We found pos-itive correlations between working memory performance and cerebralmetabolism in the temporal gyrus, and between cognitive flexibilityand cerebral metabolism in the superior frontal gyrus.

3.3.2. Hot executive functions –whole brain voxel-based correlates in SDIFor the R-SAT (self-regulation) results showed significant positive

correlations with four different clusters. One of them had a peak atthe posterior cingulate (18–65 12), including bilateral middle andposterior cingulate cortex and right precuneus. A second cluster hada statistical peak at the anterior cingulate (14 32 13) extending tothe insula. A third cluster encompassed regions of the precentralgyrus extending to superior frontal cortex and to the postcentralgyrus and supplementary motor area (peak at precentral gyrus22–16 60). The fourth cluster had a statistical peak at the inferiortemporal gyrus (40–0 −40) extending to the temporal pole and fusi-form. We also found a negative correlation between R-SAT

UNCO

RRECTA) Revised Strategy Application Test negativ

L -56 8

B) Iowa Gambling Task positive correlations

L -26 8

C) Ekman Faces Test negative correlations (b

L -37 -32

Fig. 2. Images showing the areas where hot executi

Please cite this article as: Moreno-López, L., et al., Neural correlates of hation between neuropsychological performance..., Psychiatry Research:

ED P

RO

OF

performance and a cluster localized in the temporal cortex which in-cluded the middle and superior temporal gyrus and the temporal pole(peak at the middle temporal gyrus −55 9–17) (Fig. 2 –Panel A).

For decision-making, there was a positive correlation between IGTscores and three clusters. The first cluster included theDLPFC extendingto posterior areas (peak at the right middle frontal gyrus 26–3 55), asecond cluster which included the DLPFC bilaterally (peak at leftmiddlefrontal gyrus −28 4 50 and right middle frontal gyrus 53 14 51) andtwo clusters with peak at the superior frontal gyrus (−18 67 10; −2045 46) (Fig. 2 –Panel B).

For emotion processing, there was a negative correlation betweenEFT emotion recognition hits and two main clusters. One includingthe lingual gyrus (peak at lingual gyrus, 0–99 −7) and a second clus-ter including the cerebellum and fusiform and lingual gyrus (peak atthe fusiform gyrus −36 −76 −11) (Fig. 2 – Panel C).

In summary, we found positive correlations between self-regulationperformance and cerebral metabolism within different sections of thecingulate cortex, superior frontal cortex and inferior temporal gyrus,and between decision-making performance and DLPFC/superior frontalmetabolism. Conversely, we found negative correlations between self-regulation performance and cerebral metabolism in the superior tem-poral gyrus, and between emotion recognition performance and cere-bral metabolism in the fusiform, cerebellum and lingual gyrus.

3.3.3. The relationship between voxel-based derived regions associatedwith neuropsychological performance and estimates of drug use

We only found a significant positive correlation between amount ofheroin used and metabolism in the left middle temporal gyrus(r=0.377) at pb0.01. By relaxing the alpha level to pb0.05, we foundpositive correlations between duration of cannabis (r=0.330) and co-caine used (r=0.296) and the right precentral gyrus, and negative cor-relations between the amount of cannabis used and the left middle

e (blue) and positive (yellow) correlations.

35 R

(yellow).

35 R

lue).

R

ve functions correlated with brain metabolism.

ot and cold executive functions in polysubstance addiction: Associ-Neuroimaging (2012), doi:10.1016/j.pscychresns.2012.01.006

Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
"Whole "
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
" -"
Original text:
Inserted Text
"p <"
Original text:
Inserted Text
"p <"

T

359

360

361

362

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

488

6 L. Moreno-López et al. / Psychiatry Research: Neuroimaging xxx (2012) xxx–xxx

UNCO

RREC

frontal gyrus (r=−0.370) and left inferior parietal lobe (r=−0.307),the amount of heroin used and the right inferior temporal gyrus (r=−0.308), and the amount of alcohol used and the right middle frontalgyrus (r=−0.297).

4. Discussion

This study investigated the association between the neuropsycho-logical performance of SDI on measures of cold and hot executivefunctions (as measured outside the scanner) and PET-indexed region-al brain glucose metabolism. SDI were also compared to non-drugusing controls with respect to behavioral performance in these tasks.

We found that SDI had significantly poorer performance than HCindividuals on all of the neuropsychological tests administered. Like-wise, the neuropsychological performance in the SDI was just affectedby the duration of abstinence in the LNS task, allowing us to rule outthe effects of this variable on all other tasks. On the other hand, in ac-cordance with our initial hypotheses, we found that both cold and hotexecutive tests were significantly associated with common brain re-gions, including the DLPFC, mid-superior frontal gyrus and superior,inferior and middle temporal gyrus. All these regions are thought toplay important roles in the executive control of goal-driven behavior(Bechara and Van der Linden, 2005), and in neuroadaptations relatedto addiction (Koob and Volkow, 2010); therefore, our findings high-light the relevance of this set of regions for a wide range of cognitivefunctions that are significantly impaired in SDI. In terms of differentialassociations, cold executive tests mainly correlated with frontal, tem-poral and parietal regions, whereas hot executive functions correlatedwith midline mid-superior frontal and cingulate regions in both ROIand whole brain voxel-based analyses. These results are in agreementwith the notion that different executive processes rely in both com-mon and partially dissociable neural networks (Collette et al., 2006),and that different neural circuits are relevant to various aspects of ab-normal cognitive functioning in SDI (Verdejo-García et al., 2004).

Specifically, our results showed that verbal working memory(indexed by LNS) is mainly associated with DLPFC and temporo-parietal regions (Fig. 1 –Panel A), which is consistent with previousfindings suggesting these regions are the key neural substrates formemory updating and manipulation (Collette and Van der Linden,2002; D'Esposito, 2007). For response inhibition, results showed asso-ciations between Stroop performance and DLPFC and basal gangliametabolism (Fig. 1 –Panel B). This finding is in agreement with evi-dence from cognitive and neuroimaging studies showing thatfronto-striatal alterations related to inhibitory control are ubiquitousacross different forms of addiction, and are associated with difficultiesin ceasing drug use (Feil et al., 2010). For cognitive flexibility results,we showed associations between the number of errors and DLPFCand parietal cortex resting metabolism (Fig. 1 –Panel C), replicatingprevious findings using this probe and other switching-perseveration tests (e.g., the Wisconsin Card Sorting Test) in alco-holics (Adams et al., 1993) and healthy subjects (Thomas et al., 2011).

With regard to hot executive functions, results showed that R-SATperformance was associated with resting metabolism in the mid-posterior and ACC, insula and temporal regions (Fig. 2 –Panel A).Both insula and ACC are involved in task-level control and self-regulation of behavior (Posner et al., 2007; Nelson et al., 2010), thekey cognitive processes needed to develop and implement goal-driven adjustment of performance (as indexed by this task). ACCand insula dysfunctions have been associated with stress-inducedcraving, poor inhibitory control, impaired insight and less motivationto change in SDI (Sinha and Li, 2007; Li and Sinha, 2008; Goldstein etal., 2009; Romero et al., 2010); this range of symptoms may relate tocore deficits of self-regulation (i.e., difficulties to set and adjust be-havior according to long-term goals) in addicted individuals. Further-more, both anterior and posterior cingulate regions have beenpreviously associated with abnormal performance on multi-tasking

Please cite this article as: Moreno-López, L., et al., Neural correlates of hation between neuropsychological performance..., Psychiatry Research:

ED P

RO

OF

measures, specifically with deficits in task learning and remembering,rule-breaking and task switching (Burgess et al., 2000).

According to previous studies using similar methodologies, IGT per-formance in mid-term abstinent SDI is associated with resting metabo-lism in both OFC and DLPFC (Adinoff et al., 2003; Tucker et al., 2004;Krain et al., 2006) (Fig. 2 – Panel B). This finding is indicative of the no-tion that both cold processes, such as working memory or switchingskills relying on DLPFC functioning, and hot processes, such asmonitor-ing of reward value, relying on OFC functioning play a role in decision-making performance in SDI (Bolla et al., 2003). Finally, although wefailed to find significant associations between ventral fronto-striatal re-gions and emotional perception, we found robust associations betweenscores in this task and two areas (lingual and fusiform gyrus) (Fig. 2 –

Panel C) that have been consistently associated with face processingin numerous studies (Kanwisher and Yovel, 2006; Dinkelacker et al.,2010). The association between emotion processing and the ventrome-dial frontal stream could be better grasped by fMRI studies monitoringbrain activation during actual emotional arousal.

Finally, we found significant associations between patterns of druguse and the regions associated with executive performance in bothROI andwhole brain voxel-based analysis. Althoughmostly exploratory,some of these effects are informative about the link between the use ofparticular substances and the brain regions supporting different execu-tive functions. For example, higher metabolism in the middle temporalgyrus was associated with both R-SAT performance and severity of her-oin use, in agreement with data from resting fMRI showing left middletemporal gyrus dysfunction in chronic heroin users (Jiang et al., 2011).Higher metabolism in the right precentral gyrus was associated withboth severity of cocaine use and cognitive flexibility, such that we mayspeculatively suggest a role of this region in cocaine-inducedmotor per-severation deficits (Ersche et al., 2008). Lower metabolism in the supe-rior frontal gyrus was associated with both severity of alcohol use andperformance on cold executive tasks andon decision-making; this resultis fitting with the findings of Goldstein et al. (2004) who found that se-verity of alcohol use was the main predictor of a composite index of ex-ecutive impairment, whichwas predicted by regional metabolism in theanterior cingulate. Finally, lower metabolism in the adjacent middlefrontal gyrus was linked to severity of cannabis use and performanceon response inhibition and decision-making, fittingwith evidence of im-pairment of these functions in chronic cannabis users (Verdejo-Garcia etal., 2007; Battisti et al., 2010). These associations are in agreement withthe notion that amount and duration of drug use impact brain function-ing in key regions for executive control, supporting the relevance of thecorrelations between executive performance and brain metabolism inthis clinical group of SDI.

Several possible limitations of our study should be considered andaddressed by future research. In the first place, the methodology usedin this study does not allow us to monitor brain activity as a direct re-sponse to neuropsychological challenges; instead it shows deficits inresting glucose metabolism that were related to test performance.Therefore, this study cannot attribute NP function to an inability of thesubstance dependent group to activate specific regions of the brain.FMRI or PET designs that involve scanning during actual test perfor-mancemay be better suited to test the neuronal substrates of executivedysfunction in SDI. On the other hand, this design allowed us to explorethe brain correlates of highly complex neuropsychological paradigms,such as the R-SAT that need to be simplified to fit them in fMRI designs.Furthermore, PET was altogether not obtained in the control group andtherefore it is not clear whether performance differences on the NPtasks were associated with actual abnormalities (i.e. compared tohealth) in glucose metabolism at rest. Future studies should examineif there are differences in these associations between a clinical groupand a control group. A fourth limitationwas the lack of estimates of nic-otine use/dependence, considering that previous studies have observedsignificant alterations in PET-indexed cerebral metabolism associatedwith nicotine use (i.e. Rose et al., 2003; Zubieta et al., 2005) and should

ot and cold executive functions in polysubstance addiction: Associ-Neuroimaging (2012), doi:10.1016/j.pscychresns.2012.01.006

Original text:
Inserted Text
" ."
Original text:
Inserted Text
"’"

T

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

513

514

515516517518519520521522523524525526527528529530531532533534535536537538539540541542543544545546547548549550551552553554555556557558

559560561562563564565566567568569570571572573574575576577578579580581582583584585586587588589590591592593594595596597598599600601602603604605606607608609610611612613614615616617618619620621622623624625626627628629630631632633634635636637638639640641642643644

7L. Moreno-López et al. / Psychiatry Research: Neuroimaging xxx (2012) xxx–xxx

UNCO

RREC

be assessed in any PET study of drug use. The voxel based analysesproved to be overall more sensitive than the ROI method, this is anexpected outcome because the voxel based method is used to carryout statistical tests in smaller brain units (voxels) than the ROI analysis(whole gyri). Finally, due to its correlational design, this study cannotyield conclusions about cause–effect relationships in the area of theneuropsychology of drug addiction.

Taken together, our results show that SDI have neuropsychologi-cal deficits in both cold and hot executive functions. These deficitsare associated with brain metabolism in both common and partiallydissociable neural networks encompassing frontal, parietal, temporaland basal ganglia regions.

Acknowledgements

This study has been supported with funds from the SEJ2006-08278 project from the Spanish Ministry of Education and Science,the P07-HUM-03089 project from the Council of Science and Innova-tion. Laura Moreno López is supported by the FPU grant (AP2007-03583) from the Spanish Ministry of Science and Innovation. AntonioVerdejo-García was supported by a MCINN Jose Castillejo grant in theDepartment of Experimental Psychology, University of Cambridge.Emmanuel A. Stamatakis is supported by the Stephen Erskine Fellow-ship from Queens' College, Cambridge.

Appendix A. Supplementary data

Supplementary data to this article can be found online at doi:10.1016/j.pscychresns.2012.01.006.

References

Adams, K.M., Gilman, S., Koeppe, R.A., Kluin, K.J., Brunberg, J.A., Dede, D., Berent, S.,Kroll, P.D., 1993. Neuropsychological deficits are correlated with frontal hypometa-bolism in positron emission tomography studies of older alcoholic patients. Alco-holism, Clinical and Experimental Research 17, 205–210.

Adinoff, B., Devous, M.D., Cooper, D.B., Best, S.E., Chandler, P., Harris, T., Cervin, C.A.,Cullum, C.M., 2003. Resting regional cerebral blood flow and gambling task perfor-mance in cocaine-dependent subjects and healthy comparison subjects. The Amer-ican Journal of Psychiatry 160, 1892–1894.

Anderson, S.W., Barrash, J., Bechara, A., Tranel, D., 2006. Impairments of emotion andreal-world complex behavior following childhood- or adult-onset damage to ven-tromedial prefrontal cortex. Journal of the International Neuropsychological Socie-ty 12, 224–235.

Baler, R.D., Volkow, N.D., 2006. Drug addiction: the neurobiology of disrupted self-control. Trends in Molecular Medicine 12, 559–566.

Battisti, R.A., Roodenrys, S., Johnstone, S.J., Pesa, N., Hermens, D.F., Solowij, N., 2010.Chronic cannabis users show altered neurophysiological functioning on Strooptask conflict resolution. Psychopharmacology 212, 613–624.

Bechara, A., Van Der Linden, M., 2005. Decision-making and impulse control after fron-tal lobe injuries. Current Opinion in Neurology 18, 734–739.

Bechara, A., Damasio, A.R., Damasio, H., Anderson, S.W., 1994. Insensitivity to future con-sequences following damage to the human prefrontal cortex. Cognition 50, 7–15.

Bolla, K.I., Eldreth, D.A., London, E.D., Kiehl, K.A., Mouratidis, M., Contoreggi, C.,Matochik, J.A., Kurian, V., Cadet, J.L., Kimes, A.S., Funderburk, F.R., Ernst, M., 2003.Orbitofrontal cortex dysfunction in abstinent cocaine abusers performing adecision-making task. NeuroImage 19, 1085–1094.

Bolla, K., Ernst, M., Kiehl, K., Mouratidis, M., Eldreth, D., Contoreggi, C., Matochik, J.,Kurian, V., Cadet, J., Kimes, A., Funderburk, F., London, E., 2004. Prefrontal corticaldysfunction in abstinent cocaine abusers. The Journal of Neuropsychiatry and Clin-ical Neurosciences 16, 456–464.

Brett, M., Anton, J.L., Valabregue, R., Poline, J.B., 2002. Region of interest analysis usingan SPM toolbox. NeuroImage 16, 2.

Brooks III, J.O., Rosen, A.C., Hoblyn, J.C., Woodard, S.A., Krasnykh, O., Ketter, T.A., 2009.Resting prefrontal hypometabolism and paralimbic hypermetabolism related toverbal recall deficits in euthymic older adults with bipolar disorder. The AmericanJournal of Geriatric Psychiatry 17, 1022–1029.

Burgess, P.W., Veitch, E., de Lacy Costello, A., Shallice, T., 2000. The cognitive and neu-roanatomical correlates of multitasking. Neuropsychologia 38, 848–863.

Collette, F., Van der Linden, M., 2002. Brain imaging of the central executive componentof working memory. Neuroscience and Biobehavioral Reviews 26, 105–125.

Collette, F., Hogge, M., Salmon, E., Van der Linden, M., 2006. Exploration of the neuralsubstrates of executive functioning by functional neuroimaging. Neuroscience139, 209–221.

Dao-Castellana, M.H., Samson, Y., Legault, F., Martinot, J.L., Aubin, H.J., Crouzel, C.,Feldman, L., Barrucand, D., Rancurel, G., Féline, A., Syrota, A., 1998. Frontal dysfunc-

Please cite this article as: Moreno-López, L., et al., Neural correlates of hation between neuropsychological performance..., Psychiatry Research:

ED P

RO

OF

tion in neurologically normal chronic alcoholic subjects: metabolic and neuropsy-chological findings. Psychological Medicine 28, 1039–1048.

DeFilippis, N.A., 2002. The Category Test, Computer Version, 2nd ed. Psychological As-sessment Resources, Odessa, Florida.

Del Ser, T., García-Montalvo, J.I., Martínez Espinosa, S., Delgado Villapalos, C., Bermejo,F., 1997. The estimation of premorbid intelligence in Spanish people with the“Word Accentuation Test” and its application to the diagnosis of dementia. Brainand Cognition 33, 343–356.

Desmond, J.E., Chen, S.H., DeRosa, E., Pryor, M.R., Pfefferbaum, A., Sullivan, E.V., 2003.Increased frontocerebellar activation in alcoholics during verbal working memory:an fMRI study. NeuroImage 19, 1510–1520.

D'Esposito, M., 2007. From cognitive to neural models of working memory. Philosoph-ical Transactions of the Royal Society B: Biological Sciences 29, 761–772.

Dinkelacker, V., Grüter, M., Klaver, P., Grüter, T., Specht, K., Weis, S., Kennerknecht, I., Elger,C.E., Fernandez, G., 2010. Congenital prosopagnosia: multistage anatomical and func-tional deficits in face processing circuitry. Journal of Neurology 258, 770–782.

Eldreth, D.A., Matochik, J.A., Cadet, J.L., Bolla, K.I., 2004. Abnormal brain activity in pre-frontal brain regions in abstinent marijuana users. NeuroImage 23, 914–920.

Ersche, K.D., Roiser, J.P., Robbins, T.W., Sahakian, B.J., 2008. Chronic cocaine but notchronic amphetamine use is associated with perseverative responding in humans.Psychopharmacology 197, 421–431.

Feil, J., Sheppard, D., Fitzgerald, P.B., Yücel, M., Lubman, D.I., Bradshaw, J.L., 2010. Addic-tion, compulsive drug seeking, and the role of frontostriatal mechanisms in regu-lating inhibitory control. Neuroscience and Biobehavioral Reviews 35, 248–275.

Fernández-Serrano, M.J., Pérez-García, M., Verdejo-García, A., 2011. What are the spe-cific vs. generalized effects of drugs of abuse on neuropsychological performance?Neuroscience and Biobehavioral Reviews 35, 377–406.

Forman, S.D., Dougherty, G.G., Casey, B.J., Siegle, G.J., Braver, T.S., Barch, D.S., Stenger, V.A.,Wick-Hull, C., Pisarov, L.A., Lorensen, E., 2004. Opiate addicts lack error-dependentactivation of rostral anterior cingulate. Biological Psychiatry 55, 531–537.

Gilman, S., Adams, K., Koeppe, R.A., Berent, S., Kluin, K.J., Modell, J.G., Kroll, P., Brunberg,J.A., 1990. Cerebellar and frontal hypometabolism in alcoholic cerebellar degenera-tion studied with positron emission tomography. Annals of Neurology 28, 775–785.

Golden, C.J., 1978. Stroop Color and Word Tets. A manual for clinical and experimentaluses. Stoelting Co, Wood Dale, Illinois.

Goldstein, R.Z., Volkow, N.D., 2002. Drug addiction and its underlying neurobiologicalbasis: neuroimaging evidence for the involvement of the frontal cortex. The Amer-ican Journal of Psychiatry 159, 1642–1652.

Goldstein, R.Z., Leskowjan, A.C., Hoff, A.L., Hitzemann, R., Bashan, F., Khalsa, S., Wang,G.J., Fowler, J.S., Volkow, N.D., 2004. Severity of neuropsychological impairmentin cocaine and alcohol addiction: association with metabolism in the prefrontalcortex. Neuropsychologia 42, 1447–1458.

Goldstein, R.Z., Craig, A.D., Bechara, A., Garavan, H., Childress, A.R., Paulus, M.P.,Volkow, N.D., 2009. The neurocircuitry of impaired insight in drug addiction.Trends in Cognitive Sciences 9, 372–380.

Halpern, J.H., Pope Jr., H.G., Sherwood, A.R., Barry, S., Hudson, J.I., Yurgelun-Todd, D.,2004. Residual neuropsychological effects of illicit 3,4-methylenedioxymetham-phetamine (MDMA) in individuals with minimal exposure to other drugs. Drugand Alcohol Dependence 75, 135–147.

Hartstra, E., Oldenburg, J.F., Van Leijenhorst, L., Rombouts, S.A., Crone, E.A., 2010. Brainregions involved in the learning and application of reward rules in a two-deckgambling task. Neuropsychologia 48, 1438–1446.

Heberlein, A.S., Padon, A.A., Gillihan, S.J., Farah, M.J., Fellows, L.K., 2008. Ventromedialfrontal lobe plays a critical role in facial emotion recognition. Journal of CognitiveNeuroscience 20, 721–733.

Jager, G., Kahn, R.S., Van Den Brink, W., Van Ree, J.M., Ramsey, N.F., 2006. Long-term ef-fects of frequent cannabis use on working memory and attention: an fMRI study.Psychopharmacology 185, 358–368.

Jiang, G.H., Qiu, Y.W., Zhang, X.L., Han, L.J., Lv, X.F., Li, L.M., Lin, C.L., Zhuo, F.Z., Hu, S.Y.,Tian, J.Z., 2011. Amplitude low-frequency oscillation abnormalities in the heroinusers: a resting state fMRI study. NeuroImage 57, 149–154.

Kanwisher, N., Yovel, G., 2006. The fusiform face area: a cortical region specialized forthe perception of faces. Philosophical Transactions of the Royal Society of London.Series B, Biological Sciences 361, 2109–2128.

Kerr, A., Zelazo, P.D., 2004. Development of “hot” executive function: the children'sgambling task. Brain and Cognition 55, 148–157.

Koob, G.F., Volkow, N.D., 2010. Neurocircuitry of addiction. Neuropsychopharmacology35, 217–238.

Krain, A.L., Wilson, A.M., Arbuckle, R., Castellanos, F.X., Milham, M.P., 2006. Distinctneural mechanisms of risk and ambiguity: a meta-analysis of decision-making.NeuroImage 32, 477–484.

Lee, T.M., Zhou, W.H., Luo, X.J., Yuen, K.S., Ruan, X.Z., Weng, X.C., 2005. Neural activityassociated with cognitive regulation in heroin users: a fMRI study. NeuroscienceLetters 382, 211–216.

Levine, B., Dawson, D., Boutet, I., Schwartz, M.L., Stuss, D.T., 2000. Assessment of strate-gic self-regulation in traumatic brain injury: its relationship to injury severity andpsychosocial outcome. Neuropsychology 14, 491–500.

Li, C.S., Sinha, R., 2008. Inhibitory control and emotional stress regulation: neuroimag-ing evidence for frontal-limbic dysfunction in psycho-stimulant addiction. Neuro-science and Biobehavioral Reviews 32, 581–597.

López-Torrecillas, F., Godoy, J.F., Pérez-García, M., Godoy, D., Sánchez-Barrera, M., 2001.Variables modulating stress and coping that discriminate drug consumers fromlow or not drug consumers. Addictive Behaviors 25, 161–165.

Lubman, D.I., Yücel, M., Kettle, J.W., Scaffidi, A., Mackenzie, T., Simmons, J.G., Allen, N.B.,2009. Responsiveness to drug cues and natural rewards in opiate addiction: asso-ciations with later heroin use. Archives of General Psychiatry 66, 205–212.

ot and cold executive functions in polysubstance addiction: Associ-Neuroimaging (2012), doi:10.1016/j.pscychresns.2012.01.006

Original text:
Inserted Text
"-"
Original text:
Inserted Text
"’"
Original text:
Inserted Text
"""
Original text:
Inserted Text
"" "
Original text:
Inserted Text
"""
Original text:
Inserted Text
"" "
Original text:
Inserted Text
"A "

645646647648649650651652653654655656657658659660661662663664665666667668669670671672673674675676677678

679680681682683684685686687688689690691692693694695696697698699700701702703704705706707708709710711

712

713

8 L. Moreno-López et al. / Psychiatry Research: Neuroimaging xxx (2012) xxx–xxx

Lundqvist, T., 2010. Imaging cognitive deficits in drug abuse. Current Topics in Behav-ioral Neurosciences 3, 247–275.

Marinkovic, K., Oscar-Berman, M., Urban, T., O'Reilly, C.E., Howard, J.A., Sawyer, K.,Harris, G.J., 2009. Alcoholism and dampened temporal limbic activation to emo-tional faces. Alcoholism, Clinical and Experimental Research 33, 1880–1892.

Nelson, S.M., Dosenbach, N.U., Cohen, A.L., Wheeler, M.E., Schlaggar, B.L., Petersen, S.E.,2010. Role of the anterior insula in task-level control and focal attention. BrainStructure & Function 214, 669–680.

Passetti, F., Clark, L., Mehta, M.A., Joyce, E., King, M., 2008. Neuropsychological predictorsof clinical outcome in opiate addiction. Drug and Alcohol Dependence 94, 82–91.

Paulus, M.P., Tapert, S.F., Schuckit, M.A., 2005. Neural activation patterns ofmethamphetamine-dependent subjects during decision making predict relapse.Archives of General Psychiatry 62, 761–768.

Posner, M.I., Rothbart, M.K., Sheese, B.E., Tang, Y., 2007. The anterior cingulate gyrusand the mechanism of self-regulation. Cognitive, Affective, & Behavioral Neurosci-ence 7, 391–395.

Romero, M.J., Asensio, S., Palau, C., Sanchez, A., Romero, F.J., 2010. Cocaine addiction:diffusion tensor imaging study of the inferior frontal and anterior cingulatewhite matter. Psychiatry Research 181, 57–63.

Rose, J.E., Behm, F.M., Westman, E.C., Mathew, R.J., London, E.D., Hawk, T.C., Turkington,T.G., Coleman, R.E., 2003. PET studies of the influences of nicotine on neural sys-tems in cigarette smokers. The American Journal of Psychiatry 160, 323–333.

Salloum, J.B., Ramchandani, V.A., Bodurka, J., Rawlings, R., Momenan, R., George, D.,Hommer, D.W., 2007. Blunted rostral anterior cingulate response during a simpli-fied decoding task of negative emotional facial expressions in alcoholic patients.Alcoholism, Clinical and Experimental Research 31, 1490–1504.

Sinha, R., Li, C.S., 2007. Imaging stress- and cue-induced drug and alcohol craving: asso-ciation with relapse and clinical implications. Drug and Alcohol Review 26, 25–31.

Thomas, L.A., Hall, J.M., Skup, M., Jenkins, S.E., Pine, D.S., Leibenluft, E., 2011. A develop-mental neuroimaging investigation of the change paradigm. Developmental Sci-ence 14, 148–161.

Tomasi, D., Goldstein, R.Z., Telang, F., Maloney, T., Alia-Klein, N., Caparelli, E.C., Volkow,N.D., 2007. Widespread disruption in brain activation patterns to a working mem-ory task during cocaine abstinence. Brain Research 1171, 83–92.

UNCO

RRECT

Please cite this article as: Moreno-López, L., et al., Neural correlates of hation between neuropsychological performance..., Psychiatry Research:

PRO

OF

Tucker, K.A., Potenza, M.N., Beauvais, J.E., Browndyke, J.N., Gottschalk, P.C., Kosten, T.R.,2004. Perfusion abnormalities and decision making in cocaine dependence. Biolog-ical Psychiatry 56, 527–530.

Tzourio-Mazoyer, N., Landeau, B., Papathanassiou, D., Crivello, F., Etard, O., Delcroix, N.,Mazoyer, B., Joliot, M., 2002. Automated anatomical labelling of activations in SPMusing a macroscopic anatomical parcellation of the MNI MRI single subject brain.NeuroImage 15, 273–289.

Verdejo-García, A., Bechara, A., 2009. A somatic marker theory of addiction. Neurophar-macology 56, 48–62.

Verdejo-García, A., López-Torrecillas, F., Orozco, C., Pérez-García, M., 2004. Clinical impli-cations and methodological challenges in the study of the neuropsychological corre-lates of cannabis, stimulant and opioid abuse. Neuropsychology Review 14, 1–41.

Verdejo-Garcia, A., Benbrook, A., Funderburk, F., David, P., Cadet, J.L., Bolla, K.I., 2007.The differential relationship between cocaine use and marijuana use on decision-making performance over repeat testing with the Iowa Gambling Task. Drug andAlcohol Dependence 90, 2–11.

Verdejo-García, A., Rivas-Pérez, C., Vilar-López, R., Pérez-García, M., 2007. Strategic self-regulation, decision-making and emotion processing in poly-substance abusers intheir first year of abstinence. Drug and Alcohol Dependence 12, 139–146.

Wechsler, D., 1997. Wechsler adult intelligence scale, 3rd ed. The Psycological Corpora-tion, San Antonio, Texas.

Yaouhi, K., Bertran, F., Clochon, P., Mézenge, F., Denise, P., Foret, J., Eustache, F.,Desgranges, B., 2009. A combined neuropsychological and brain imaging study ofobstructive sleep apnea. Journal of Sleep Research 18, 36–48.

Young, A.W., Perrett, D.I., Calder, A.J., Sprengelmeyer, R., Ekman, P., 2002. Facial expressionsof emotion: stimuli and tests (FEEST). Thames Valley Test Company, Bury St. Edmunds.

Zakzanis, K.K., 2001. Statistics to tell the truth, the whole truth, and nothing but thetruth: formulae, illustrative numerical examples, and heuristic interpretation of ef-fects size analyses for neuropsychological researchers. Archives of Clinical Neuro-psychology 16, 653–677.

Zubieta, J.K., Heitzeg, M.M., Xu, Y., Koeppe, R.A., Ni, L., Guthrie, S., Domino, E.F., 2005.Regional cerebral blood flow responses to smoking in tobacco smokers after over-night abstinence. The American Journal of Psychiatry 162, 567–577.

ED

ot and cold executive functions in polysubstance addiction: Associ-Neuroimaging (2012), doi:10.1016/j.pscychresns.2012.01.006

Original text:
Inserted Text
"’"