24.07.2014 Views

pdf - SRON

pdf - SRON

pdf - SRON

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Universidad de Cantabria<br />

Departamento de Física Moderna<br />

CSIC - Universidad de Cantabria<br />

Instituto de Física de Cantabria<br />

Evolución y agrupamiento<br />

a gran escala de una muestra<br />

de fuentes extragalácticas en rayos-X<br />

Memoria presentada por el Licenciado<br />

Jacobo Ebrero Carrero<br />

para optar al título de Doctor en Ciencias Físicas<br />

2008


Francisco Jesús Carrera Troyano, Doctor en Ciencias Físicas y<br />

Profesor Titular de la Universidad de Cantabria,<br />

CERTIFICA que la presente memoria<br />

Evolución y agrupamiento a gran escala de una muestra<br />

de fuentes extragalácticas en rayos-X<br />

ha sido realizada por Jacobo Ebrero Carrero bajo mi dirección.<br />

Considero que esta memoria contiene aportaciones suficientes para construir la<br />

tesis Doctoral del interesado<br />

En Santander, a 1 de Octubre de 2008<br />

Francisco Jesús Carrera Troyano


A mis padres


Resumen de la tesis en castellano<br />

0.1 Objetivos de la investigación<br />

Las emisiones de rayos-X en el Universo están asociadas a los fenómenos físicos<br />

más energéticos como, por ejemplo, partículas relativistas sometidas a intensos<br />

campos magnéticos, plasmas a temperaturas del orden de millones de grados<br />

o intensos campos gravitatorios. Por lo tanto, el estudio del cielo en rayos-X<br />

significa estudiar el Universo más violento y extremo.<br />

Dentro de nuestra propia Galaxia existen distintos tipos de fuentes que emiten<br />

rayos-X como pueden ser estrellas aisladas o en sistemas binarios (enanas blancas,<br />

estrellas de neutrones) y remanentes de supernova. Fuera de ella, también<br />

existe una variedad de objetos emisores de rayos-X como las galaxias, cúmulos<br />

de galaxias y los Núcleos Galácticos Activos (AGN, en sus siglas en inglés).<br />

El cielo en rayos-X se encuentra dominado por una radiación difusa procedente<br />

de todas direcciones, altamente isótropa, conocida como Fondo Cósmico<br />

de Rayos-X (FCX). Hoy en día sabemos que el origen del FCX es la emisión<br />

integrada de una miríada de fuentes individuales extragalácticas. A energías<br />

superiores a ∼0.2 keV la emisión del FCX se encuentra dominada por AGN, estando<br />

una fracción importante de su emisión absorbida por grandes cantidades<br />

de gas y polvo.<br />

Los AGN son galaxias que emiten cantidades ingentes de energía procedente<br />

de una pequeña región situada en el núcleo de las mismas (de ahí su nombre).<br />

La energía emitida por los AGN es muy superior a la emisión integrada de la<br />

galaxia que los contiene, siendo la región emisora mucho más pequeña (∼1 pc)<br />

que la galaxia anfitriona. Además, los AGN son las fuentes persistentes más<br />

brillantes del Universo, tan sólo superados en brillo por los estallidos de rayos<br />

gamma (aunque éstos últimos tan sólo duran unos segundos). Es imposible<br />

vii


explicar las altas luminosidades de los AGN por la emisión integrada de las<br />

estrellas de la galaxia huésped o por la emisión del gas interestelar, ni tampoco<br />

por brotes de formación estelar. El único fenómeno conocido capaz de explicar<br />

estas altísimas luminosidades y la rápida variabilidad observada en la mayoría<br />

de estos objetos, es la acreción de materia en un agujero negro supermasivo.<br />

De acuerdo con el modelo unificado de los AGN, éstos contienen en su centro<br />

agujeros negros supermasivos con masas entre 10 6 -10 9 M 1 ⊙ . En las regiones en<br />

las que el campo gravitatorio del agujero negro domina sobre el producido por<br />

las estrellas del entorno (∼10 parsecs para un agujero negro de ∼10 8 M ⊙ ), el<br />

agujero negro está acretando materia (gas y estrellas). Como el material acretado<br />

tiene momento angular no cae radialmente, sino que forma un disco plano<br />

alrededor del agujero negro llamado disco de acreción. La materia en el disco<br />

se calienta debido a procesos de fricción y la influencia de fuertes campos magnéticos,<br />

alcanzando temperaturas de 10 5 -10 6 K y liberando energía en forma de<br />

radiación térmica emitida en el rango óptico/ultravioleta (UV). A pesar de estar<br />

generalmente aceptado que la fuente de energía primaria de un AGN es la<br />

acreción gravitatoria de materia, los detalles de cómo se produce la emisión de<br />

rayos-X en estos objetos aún no están totalmente determinados.<br />

Uno de los modelos más aceptados capaces de explicar la emisión de rayos-X<br />

en AGN es el de la Comptonización inversa. Según este modelo, un plasma<br />

caliente de electrones e iones (y, quizá, pares electrón-positrón) rodea la zona<br />

del disco de acreción más próxima al agujero negro. Este plasma reprocesa mediante<br />

efecto Compton inverso los fotones ópticos/UV emitidos térmicamente<br />

en el disco, impulsándolos hasta el rango de los rayos-X. Este proceso produce<br />

un espectro de emisión con forma de ley de potencia, como se observa en espectroscopía<br />

de rayos-X de AGN.<br />

Los AGN muestran una gran variedad de propiedades, dependiendo de las<br />

cuales se clasifican en distintas categorías. Una de las clasificaciones más extendidas,<br />

la cual hemos usado para el trabajo presentado en esta tesis, está basada<br />

en las propiedades de los espectros ópticos de los AGN. Los AGN para los que<br />

se observan tanto líneas de emisión anchas (anchura a media altura ≥ 2000 km<br />

s −1 ) como estrechas (anchura a media altura < 2000 km s −1 ) en el espectro óptico/UV<br />

se clasifican como AGN de tipo 1, mientras que para los que sólo se<br />

observan líneas de emisión estrechas se clasifican como AGN de tipo 2.<br />

1 Las masas se dan en unidades de masa del Sol M ⊙<br />

viii


Los modelos de unificación pretenden explicar la presencia o ausencia de líneas<br />

anchas en los espectros ópticos/UV de los AGN como anisotropías en la geometría<br />

del material absorbente. Según estos modelos, las propiedades observadas<br />

en el óptico/UV dependen de la visual del observador con respecto al núcleo<br />

del AGN. Así, los AGN de tipo 1 serían fuentes en las que se observa directamente<br />

el núcleo, mientras que en los AGN de tipo 2 el toroide intercepta la línea<br />

de visión del observador oscureciendo el núcleo y la región donde se emiten las<br />

líneas anchas.<br />

Se cree que las líneas de emisión anchas se forman en nubes de gas denso (≥<br />

10 8 cm −3 ) que se mueven a altas velocidades (entre 1000 y 10000 km s −1 ) en las<br />

proximidades del agujero negro y el disco de acreción, a distancias del orden de<br />

100 días-luz de los mismos. Esta región se la conoce como la Región de Líneas<br />

Anchas. Por contra, las líneas de emisión estrechas son emitidas en nubes de<br />

gas menos densas (∼ 10 3 − 10 4 cm −3 ) más alejadas de la región central del<br />

AGN (distancias del orden del kiloparsec) y que, por lo tanto, se mueven a<br />

velocidades menores (entre 100 y 1000 km s −1 ). Esta región es conocida como<br />

la Región de Líneas Estrechas.<br />

Otra componente, el gas molecular, se encuentra situado a distancias intermedias<br />

entre la Región de Líneas Anchas y la Región de Líneas Estrechas y, muy<br />

posiblemente, presenta geometría toroidal. La materia presente en el toroide<br />

absorbe la radiación procedente del disco de acreción y de la Región de Líneas<br />

Anchas, oscureciendo así el núcleo del AGN para ciertas líneas de visión. La<br />

radiación absorbida es reemitida en el infrarrojo (IR).<br />

Los AGN son emisores de radiación en todo el rango del espectro, desde IR<br />

y radio hasta los rayos gamma, aunque la emisión en rayos-X (que normalmente<br />

supone el 3-10% de la emisión total del AGN) es una propiedad común<br />

de los AGN. Esta fuerte emisión X supone una prueba directa de la existencia<br />

de procesos muy energéticos y de origen no térmico, ya que son imposibles de<br />

producir en cantidad suficiente mediante procesos puramente estelares.<br />

Los modelos de síntesis del FCX intentan explicar la emisión del fondo cósmico<br />

y la distribución en flujo de las fuentes como una superposición de AGN. Sin<br />

embargo, para que las predicciones teóricas y las observaciones ajusten es necesario<br />

que una fracción importante de la emisión de los AGN se encuentre absorbida<br />

por grandes cantidades de materia. Es decir, los modelos predicen que<br />

la mayor parte de la emisión por acreción del Universo se encuentra oscurecida.<br />

ix


Además, una importante fracción (∼50%) de las galaxias activas cercanas observadas<br />

se encuentran también oscurecidas. Los rayos-X más energéticos son<br />

capaces de atravesar importantes cantidades de material absorbente sin verse<br />

apenas afectados, proporcionando así la única manera (exceptuando el infrarrojo<br />

lejano, FIR) de ver directamente la emisión procedente de las regiones más<br />

internas de los AGN oscurecidos. Por lo tanto, la mejor manera de estudiar la<br />

estructura y los mecanismos de emisión de las regiones más cercanas al agujero<br />

negro central es estudiando la emisión de los AGN en rayos-X.<br />

La actual generación de observatorios de rayos-X, Chandra, XMM-Newton y<br />

Suzaku, está proporcionando datos de calidad sin precedentes que nos están<br />

permitiendo incrementar significativamente nuestro conocimiento acerca de la<br />

naturaleza de los AGN. En concreto, XMM-Newton, gracias a su gran área colectora<br />

nos permite detectar AGN débiles y situados a una gran distancia. No está<br />

clara la dependencia de los mecanismos de emisión de los AGN con distintos<br />

parámetros de las fuentes tales como el flujo, la luminosidad en rayos-X y el<br />

desplazamiento al rojo. Por ello, para entender la naturaleza y evolución de<br />

los AGN es necesario cuantificar estas dependencias. El mejor modo de hacer<br />

esto es mediante análisis estadísticos de muestras amplias y representativas de<br />

fuentes. Este es el camino seguido para la investigación llevada a cabo para esta<br />

tesis.<br />

Algunos de los aspectos que nos hemos planteado abordar en nuestra investigación<br />

son los siguientes:<br />

1. Estudiar la distribución en flujo de fuentes extragalácticas de rayos-X a<br />

flujos medios y débiles. Esta distribución depende de las propiedades<br />

cosmológicas de las fuentes y puede ser descrita por medio de un modelo<br />

empírico que es función del flujo en rayos-X. De esta forma será posible<br />

estimar la contribución de las fuentes emisoras a estos flujos a la intensidad<br />

total del fondo cósmico de rayos-X. ¿A qué flujos contribuyen más las<br />

fuentes que dominan el FCX?<br />

2. Explorar la posible estructura cósmica subyacente del Universo en rayos-<br />

X. Es conocido que las galaxias en el Universo local (con desplazamientos<br />

al rojo de hasta ∼0.2) se agrupan formando estructuras a gran escala en<br />

forma de filamentos alrededor de grandes vacíos. Los AGN se encuentran<br />

a distancias cosmológicas y, por tanto, nos van a permitir descubrir si estas<br />

super-estructuras estaban presentes cuando el Universo era más joven y<br />

x


cuantificar el grado de este agrupamiento.<br />

3. Estudiar como se comporta la fracción de AGN oscurecidos a distintas luminosidades<br />

en rayos-X y desplazamientos al rojo y modelar la absorción<br />

intrínseca de los AGN como función de estos parámetros. De acuerdo<br />

con el modelo unificado descrito con anterioridad, es más probable que<br />

los AGN detectados en rayos-X más energéticos (> 2 keV) sean AGN de<br />

tipo-2, ya que sólo esta radiación puede atravesar el toroide que rodea el<br />

agujero negro central. ¿Depende la fracción de AGN oscurecidos de la luminosidad?<br />

¿Había más AGN oscurecidos en el Universo joven que en el<br />

tardío?<br />

4. Investigar la evolución de los AGN en distintas épocas del Universo. ¿Eran<br />

más o menos luminosos en el pasado? ¿La densidad de AGN ha cambiado<br />

con el tiempo? Podemos integrar un modelo de absorción con modelos<br />

evolutivos para obtener resultados intrínsecos de las fuentes. ¿Ha<br />

cambiado la tasa de acreción de materia en el tiempo? ¿En qué época del<br />

Universo ya se habían formado plenamente los AGN?<br />

El objetivo último es cuantificar qué aportan las muestras seleccionadas en<br />

rayos-X a flujos medios, como las empleadas en esta tesis, a los conocimientos<br />

existentes sobre los temas enumerados anteriormente. Es de esperar que<br />

una selección a estos flujos, en los que la población dominante son AGN, nos<br />

proporcione un mejor muestreo de estos objetos a desplazamientos al rojo bajos/medios<br />

y, en especial, mayor número de AGN absorbidos en las bandas<br />

de rayos-X duros que en otras muestras. Esto nos permitirá estudiar mejor sus<br />

propiedades de absorción, que podremos usar posteriormente para derivar sus<br />

propiedades cosmológicas sin sesgos.<br />

0.2 Planteamiento y metodología<br />

Las misiones XMM-Newton de la ESA y Chandra de la NASA, ambos lanzados<br />

en 1999, son los observatorios de rayos-X más potentes en órbita hasta la<br />

fecha. Chandra es principalmete un observatorio de imagen: se encuentra optimizado<br />

para obtener altas resoluciones angulares (con precisiones del orden<br />

de ∼0.5"). Por su parte, XMM-Newton tiene una resolución espacial moderada<br />

(∼15") pero una mayor superficie colectora, siendo un observatorio ideal para<br />

xi


ealizar estudios de espectroscopía en rayos-X. XMM-Newton lleva a bordo tres<br />

instrumentos operados en paralelo: RGS, que consta de dos espectrómetros de<br />

dispersión de alta resolución (E/∆E ∼ 100 − 700) en el rango de energías 0.3-<br />

2.1 keV, tres cámaras de rayos-X (EPIC) llamadas MOS1, MOS2 y pn, y un monitor<br />

óptico/UV. Las cámaras EPIC cubren un campo de visión de 30’ y tienen<br />

resolución espectral moderada (E/∆E ∼ 20 − 50). Los datos empleados para<br />

construir las muestras empleadas en esta tesis han sido obtenidos con el detector<br />

EPIC-pn.<br />

Para dar respuesta a las preguntas planteadas en la sección anterior, hemos empleado<br />

dos muestras de fuentes extragalácticas de rayos-X. Se tratan de AXIS<br />

(An XMM-Newton International Survey) y XMS (XMM-Newton Medium Sensitivity<br />

Survey), construidas y lideradas por el Grupo de Astronomía de Rayos-X<br />

del Instituto de Física de Cantabria. El autor de esta tesis ha participado directamente<br />

en ambos proyectos, realizando observaciones y reduciendo datos<br />

ópticos y de rayos-X que han sido posteriormente utilizados para obtener los<br />

resultados presentados en esta tesis. La muestra AXIS fue extraída de 36 observaciones<br />

de XMM-Newton a altas latitudes galácticas (|b| ≥ 20 ◦ ) para minimizar<br />

el número de fuentes en nuestra propia Galaxia que pudieran entrar en la muestra.<br />

Las fuentes detectadas cubren un ángulo sólido de ∼4.8 grados cuadrados<br />

y tienen flujos en rayos-X entre ∼ 10 −15 − 10 −12 erg cm −2 s −1 . La muestra XMS<br />

es una submuestra limitada en flujo de 25 campos de AXIS con tiempos de exposición<br />

típicos de ∼20 kilosegundos, y cubre en el cielo un área de en torno a<br />

∼3 grados cuadrados. XMS forma parte de un extensivo programa de identificaciones<br />

mediante espectroscopía óptica por lo que, como resultado del mismo,<br />

una importante fracción (∼90%) de esta muestra se haya identificada y sus desplazamientos<br />

al rojo medidos de forma fiable, siendo la muestra más extensa<br />

con alto grado de completitud construida hasta la fecha.<br />

Los estudios evolutivos de AGN requieren una gran precisión en la determinación<br />

del desplazamiento al rojo de las fuentes. De lo contrario, los resultados<br />

podrían estar sesgados y conducir a conclusiones erróneas. El proceso de<br />

identificación de las fuentes conlleva un procedimiento que se puede resumir<br />

de la siguiente forma. En primer lugar hay que tomar imágenes en el óptico<br />

de las zonas donde se detectan fuentes de rayos-X, que posteriormente se utilizan<br />

para buscar las posibles contrapartidas ópticas de las mismas (las fuentes<br />

detectadas en el óptico cuya posición se encuentra próxima a la posición de<br />

las fuentes de rayos-X). En algunos casos no es posible detectar contrapartidas<br />

xii


ópticas, lo que implica que estas fuentes de rayos-X pueden encontrarse fuertemente<br />

oscurecidas. En estos casos, se usan otras longitudes de onda menos sensibles<br />

al oscurecimiento (infrarrojo cercano, infrarrojo medio) para encontrar las<br />

posibles contrapartidas. Seguidamente se realiza espectroscopía en el óptico de<br />

las contrapartidas seleccionadas. Cuando la densidad de fuentes es elevada es<br />

posible hacer espectroscopía multiobjeto (tomar varios espectros de distintas<br />

fuentes al mismo tiempo). En el caso de XMS, esto se hizo en las primeras fases<br />

de la identificación con espectrógrafos de fibras. Posteriormente, a medida que<br />

las contrapartidas restantes son más débiles y su densidad es menor, hay que<br />

ir una a una mediante espectroscopía de rendija larga. Una vez reducidos los<br />

espectros, se procede a la revisión visual de los resultados. De esta forma se<br />

clasifican en función de su naturaleza (si presentan o no líneas de emisión, y si<br />

éstas son anchas y estrechas - AGN de tipo 1 - o sólo estrechas - AGN de tipo<br />

2) y se obtienen sus desplazamientos al rojo comparando la longitud de onda<br />

observada de las líneas con su longitud de onda en reposo.<br />

Para evaluar la contribución de nuestras fuentes al FCX necesitamos modelar la<br />

distribución en flujo de las mismas. Cuanta mayor precisión obtengamos en los<br />

parámetros que modelan la distribución en flujo, más precisa será nuestra estimación<br />

de la contribución al FCX de nuestras fuentes. Además, para una mayor<br />

precisión y cubrir un mayor rango de flujos, hemos añadido a la muestra AXIS<br />

otras muestras más profundas y brillantes disponibles en la literatura. El procedimiento<br />

que seguimos consta de dos pasos. En primer lugar calculamos la<br />

distribución en flujo diferencial de las fuentes en intervalos de flujo que, esencialmente,<br />

es el sumatorio del inverso del área efectiva de las fuentes en cada<br />

intervalo dividido por la anchura del intervalo. A esta distribución diferencial<br />

le ajustamos mediante una técnica de χ 2 un modelo empírico que, en este<br />

caso, es una doble ley de potencias. Esto nos proporciona un valor inicial de los<br />

parámetros que emplearemos como punto de inicio del siguiente paso. Como<br />

el análisis de la distribución por intervalos puede estar sometido a sesgos (por<br />

ejemplo, fluctuaciones debidas a unas pocas fuentes, sobre todo a flujos altos<br />

en donde el número de fuentes es menor) hemos preferido ajustar nuestro modelo<br />

empírico mediante un método de máxima verosimilitud, que hace uso de<br />

todas y cada una de las fuentes disponibles. La función de verosimilitud a minimizar<br />

tiene en cuenta no sólo la variación del área efectiva como función del<br />

flujo, sino también las incertidumbres en el flujo de la fuente. El algoritmo de<br />

minimización emplea como punto inicial el mejor ajuste que obtuvimos de la<br />

xiii


distribución diferencial que, al estar próximo a la solución real, requerirá menos<br />

iteraciones del algoritmo.<br />

Una vez que se dispone de una medida precisa de la distribución en flujo de<br />

las fuentes, se puede estimar la intensidad con la que éstas contribuyen al FCX,<br />

integrando en flujo el mejor ajuste obtenido anteriormente y comparándolo con<br />

la intensidad total del FCX, valor que se puede encontrar en la literatura.<br />

Las fuentes de la muestra AXIS también se pueden usar para explorar la estructura<br />

a gran escala del Universo en rayos-X. Si existe una estructura cósmica que<br />

se encuentre presente en todos (o, al menos, en una buena parte de) los campos<br />

observados, ésto se debe traducir en un exceso en la distribución de distancias<br />

entre pares de fuentes con respecto de la de una distribución aleatoria. Ésta es<br />

la definición de una función de correlación, que es uno de los medios estadísticos<br />

que hemos empleado para medir el grado de agrupamiento de las fuentes.<br />

Efectivamente, se pueden medir distancias angulares entre pares de fuentes y<br />

buscar excesos en el número de pares de fuentes a una distancia angular dada<br />

respecto al que nos daría una muestra artificial generada aleatoriamente. A<br />

la hora de simular una muestra hay que tener en cuenta las variaciones en la<br />

sensibilidad de la detección de fuentes reales para no obtener sobredensidades<br />

espúreas. La función de correlación angular así medida se ajusta por χ 2 a un<br />

modelo analítico (una ley de potencias en este caso) cuyos parámetros nos dicen<br />

la distancia de correlación típica entre fuentes y la fuerza del agrupamiento.<br />

Alternativamente, también hemos calculado el agrupamiento angular de las<br />

fuentes AXIS mediante un algoritmo basado en estadística Poissoniana. Este<br />

método compara el número esperado de pares de fuentes a distancias angulares<br />

menores o iguales que unas observadas en ausencia de correlación con el<br />

número total de pares de fuentes observados. De esta forma se puede obtener<br />

la probabilidad Poissoniana integrada de obtener menos pares que los observados<br />

para una distribución de media el número de pares aleatorios. Además<br />

de darnos indicios acerca de la presencia o ausencia de agrupamiento angular,<br />

éste método estima también la significancia de la detección para lo que hicimos<br />

10000 simulaciones de muestras sintéticas sobre las que comparar los valores<br />

observados.<br />

No obstante, las separaciones angulares entre fuentes son una proyección en<br />

el cielo de las auténticas separaciones espaciales entre dos fuentes que se encuentren<br />

a distintos desplazamientos al rojo. Aún así, el agrupamiento angular<br />

xiv


es una herramienta poderosa para estimar la estructura a gran escala cuando<br />

las muestras empleadas son grandes. Además, es posible desproyectar el agrupamiento<br />

angular al espacio tridimensional si se asume una distribución de<br />

desplazamientos al rojo. Este método se conoce como inversión de la ecuación<br />

de Limber.<br />

Análogamente a la función de correlación angular existe una generalización en<br />

tres dimensiones llamada función de correlación espacial que emplea el mismo<br />

procedimiento. Como es de esperar, en este caso se emplean distancias físicas<br />

entre pares de fuentes de las que se conoce su desplazamiento al rojo (como es<br />

el caso de las fuentes de la muestra XMS). El proceso de cálculo es análogo al<br />

anterior y proporciona la distancia espacial de correlación típica entre pares de<br />

fuentes y la fuerza del agrupamiento.<br />

Para realizar el estudio evolutivo de los AGN presentes en la muestra XMS en<br />

primer lugar hemos modelado la absorción intrínseca de las fuentes detectadas<br />

en los rayos-X más energéticos (que, como hemos visto, son fuentes absorbidas<br />

en una fracción importante) como función de la luminosidad y el desplazamiento<br />

al rojo. Esto nos ha permitido investigar la evolución de la luminosidad<br />

intrínseca (antes de la absorción) de estas fuentes. Para ello hemos ajustado<br />

las observaciones a distintos modelos de evolución (evolución pura en luminosidad<br />

y evolución tanto en la luminosidad como en la densidad comóvil de<br />

fuentes) de la función de luminosidad de los AGN. La función de luminosidad<br />

mide el número de fuentes por unidad de luminosidad por unidad de volumen<br />

comóvil (o la densidad comóvil de fuentes por unidad de luminosidad).<br />

El método empleado para hacer el ajuste es, de nuevo, un método de máxima<br />

verosimilitud. Como hemos comentado anteriormente, éste método explota la<br />

información disponible para cada fuente y está libre de todos los sesgos y artificios<br />

que suelen acompañar a las estimaciones hechas por intervalos. Como en<br />

el caso del cálculo de la distribución en flujo de las fuentes AXIS, para el estudio<br />

de la evolución de los AGN también hemos añadido a la muestra XMS otras<br />

muestras altamente identificadas más débiles y más brillantes para ampliar la<br />

cobertura del plano luminosidad-desplazamiento al rojo sobre el que vamos a<br />

realizar el estudio.<br />

A partir del mejor modelo que describe la evolución de los AGN es posible estudiar<br />

la historia de la acreción en el Universo. Como hemos comentado, el origen<br />

de la emisión en rayos-X en los AGN es producida por acreción gravitatoria de<br />

xv


materia sobre el agujero negro central. Una vez que se han corregido los posibles<br />

efectos de la absorción, cambios en la luminosidad X de los AGN a lo largo<br />

del tiempo cósmico implican cambios intrínsecos en la acreción de materia. De<br />

esta forma, a partir de la función de luminosidad adoptada es posible calcular<br />

la densidad comóvil de los AGN, la tasa de acreción volúmica y la cantidad de<br />

materia acretada en función del tiempo.<br />

0.3 Aportaciones originales<br />

La investigación llevada a cabo en esta tesis ha contribuído significativamente<br />

al conocimiento de las propiedades evolutivas de los AGN y cómo su emisión<br />

contribuye a la emisión del fondo cósmico de rayos-X. El análisis de la distribución<br />

en flujo de fuentes-X extragalácticas llevado a cabo en esta tesis es<br />

uno de los más precisos hasta la fecha, gracias al tamaño de la muestra empleada,<br />

reduciendo significativamente las incertidumbres que existían en algunos<br />

parámetros del modelo. Además, el método para ajustar las observaciones<br />

al modelo que hemos empleado tiene en cuenta las incertidumbres en<br />

los flujos observados lo que incrementa la fiabilidad de los resultados.<br />

El estudio de estructura a gran escala del Universo en rayos-X nos ha permitido<br />

estimar el grado de agrupamiento de las fuentes. Además hemos podido cuantificar<br />

la densidad mínima de fuentes necesaria para obtener una señal significativa<br />

de agrupamiento. Por debajo de ella, el agrupamiento subyacente sería<br />

indetectable aunque existiese. Adicionalmente, uno de los métodos empleados<br />

para cuantificar la significancia en la detección del agrupamiento, basado en la<br />

estadística Poissoniana, es original.<br />

Finalmente, hemos estudiado la evolución de los AGN a partir de su función de<br />

luminosidad. La muestra combinada empleada es de las mayores compiladas<br />

hasta la fecha, lo que se ha traducido en una gran precisión en la determinación<br />

de los parámetros evolutivos, disminuyendo las incertidumbres en varios de<br />

ellos con respecto a otros trabajos. Además hemos llevado a cabo el estudio<br />

evolutivo sobre una submuestra de AGN detectados en rayos-X muy duros<br />

(energías en el rango 4.5-7.5 keV) por primera vez hasta la fecha.<br />

xvi


0.4 Conclusiones<br />

Los resultados más importantes de los análisis llevados a cabo sobre las fuentes<br />

de las muestras AXIS y XMS detectadas en rayos-X blandos (energías 2 keV) son los siguientes:<br />

1. La distribución en flujo de fuentes detectadas en rayos-X blandos (bandas<br />

0.5-2 keV y 0.5-4.5 keV) siguen un modelo de doble ley de potencias,<br />

presentando un cambio de pendiente a flujos ∼ 10 −14 erg cm −2 s −1 . Las<br />

fuentes detectadas en rayos-X duros (banda 2-10 keV) también siguen este<br />

modelo aunque hay evidencias de que puede existir un cambio de pendiente<br />

adicional a flujos ∼ 3 × 10 −15 erg cm −2 s −1 . Es difícil asignar un<br />

origen a esta variación (que es poco significativa) que puede ser debida a<br />

desviaciones del modelo debido a la simplicidad del mismo (ya que estamos<br />

intentado describir contribuciones de una superposición de distintas<br />

poblaciones a distintos desplazamientos al rojo) o a incertidumbres en la<br />

calibración de los datos. Los resultados obtenidos son compatibles con<br />

otros trabajos anteriores aunque, gracias al tamaño de la muestra y la amplia<br />

cobertura en flujo disponible, hemos reducido significativamente las<br />

incertidumbres en los parámetros del modelo.<br />

2. La distribución en flujo de las fuentes con espectros más duros (aquellas<br />

que cuantitativamente emiten más rayos-X duros que blandos) no presenta<br />

cambio de pendiente. Esto puede implicar que este tipo de fuentes<br />

son detectadas a desplazamientos al rojo más bajos y, por lo tanto, no<br />

muestran evolución, o ésta es muy poco intensa desde esa época hasta<br />

el presente.<br />

3. Hemos resuelto la intensidad de la emisión del FCX hasta las flujos más<br />

debiles a los que tenemos acceso usando nuestra muestra combinada en<br />

un ∼90%. La intensidad producida extrapolando nuestra distribución en<br />

flujo a flujo cero no satura la intensidad total del FCX. Si asumimos que la<br />

distribución en flujo vuelve a cambiar por debajo del flujo mínimo detectado<br />

en nuestra muestra, hemos comprobado que una población de<br />

galaxias dominante a esos flujos sería suficiente para reproducir el FCX<br />

empleando únicamente fuentes discretas.<br />

4. La máxima contribución fraccional al FCX en todas las bandas de energía<br />

xvii


analizadas proviene de fuentes en torno a ∼ 10 −14 erg cm −2 s −1 (aproximadante<br />

donde ocurre el cambio de pendiente en la distribución en flujo).<br />

La emisión a estos flujos supone el ∼50% de la contribución total al FCX<br />

en las bandas 0.5-2 y 2-10 keV, resaltando así la importancia de muestreos<br />

a flujos medios como AXIS para el estudio de la evolución de la radiación<br />

en rayos-X del Universo hasta 10 keV.<br />

5. Hemos encontrado evidencias de agrupamiento angular a gran escala en<br />

el Universo de rayos-X blandos (bandas 0.5-2 y 0.5-4.5 keV) con una significancia<br />

de ∼99%-99.9%. No así para las fuentes duras, que ofrecen resultados<br />

comparables con una distribución aleatoria. Hemos comprobado<br />

que esto puede ser debido a la baja densidad de fuentes por campo en<br />

esta banda, que puede diluir la presencia de agrupamiento hasta hacerlo<br />

indetectable.<br />

6. Dividiendo la muestra entre varias submuestras comprobamos que la señal<br />

de agrupamiento está extendida por todo el cielo, no encontrándose limitada<br />

a sólo unos pocos campos. Esto implica que si esta estructura tiene<br />

un origen cosmológico, debe provenir de desplazamientos al rojo ≤1.5 (el<br />

pico de la distribución de desplazamientos al rojo de muestras a flujos<br />

medios).<br />

7. Hemos usado la información espectral detallada de la que disponemos<br />

para las fuentes duras (>2 keV) para modelar su absorción intrínseca como<br />

función de la luminosidad y el desplazamiento al rojo. Encontramos que la<br />

fracción de AGN absorbidos disminuye al incrementarse la luminosidad<br />

y aumenta con el desplazamiento al rojo para las fuentes detectadas en la<br />

banda 2-10 keV. Las fuentes más duras, detectadas en la banda 4.5-7.5 keV,<br />

presentan también una disminución en la fracción de AGN absorbidos con<br />

la luminosidad pero, sin embargo, no hay dependencia significativa con el<br />

desplazamiento al rojo. Esto puede ser debido al escaso rango de desplazamientos<br />

al rojo abarcado por esta submuestra, con la mayoría de las<br />

fuentes a desplazamientos al rojo bajos (z < 1).<br />

8. La evolución de los AGN no está bien descrita asumiendo que únicamente<br />

su luminosidad evoluciona con el tiempo. Hemos comprobado que el modelo<br />

que mejor se ajusta a las observaciones es aquel en el que existe evolución<br />

tanto en la luminosidad como en la densidad comóvil de objetos. Gracias<br />

al gran número de fuentes involucradas, hemos mejorado la precisión<br />

xviii


en la determinación de los parámetros de este modelo con respecto a trabajos<br />

anteriores. Además, hemos estudiado por primera vez la función de<br />

luminosidad intrínseca (corrigiendo por la absorción) de las fuentes más<br />

duras (4.5-7.5 keV), encontrando que éstas presentan una evolución significativamente<br />

más fuerte a desplazamientos al rojo bajos que el resto de<br />

AGN. El total de materia acretada en el Universo actual (a desplazamiento<br />

al rojo cero) predicha por nuestro modelo coincide con la estimación de la<br />

densidad de agujeros negros supermasivos en el Universo local hecha por<br />

otros autores empleando otros métodos como la función de masa local de<br />

agujeros negros.<br />

9. En todas las bandas de energía estudiadas, los AGN más luminosos (log L X ><br />

44) se forman antes que los menos luminosos (log L X < 44), alcanzando<br />

un máximo en la densidad de objetos a desplazamientos al rojo de ∼1.5,<br />

mientras que los segundos lo hacen a desplazamientos al rojo menores<br />

(∼0.7). Esto indica que los AGN más luminosos acretaban materia más<br />

eficientemente que los menos luminosos en las etapas tempranas del Universo,<br />

encontrándose plenamente formados a desplazamientos al rojo de<br />

∼1.5-2. Este comportamiento confirma que la evolución de la función de<br />

luminosidad observada a lo largo del tiempo cósmico no está causada por<br />

variaciones en el entorno absorbente de rayos-X, sino por variaciones intrínsecas<br />

en la tasa de acreción de materia sobre el agujero negro supermasivo.<br />

0.5 Futuras líneas de investigación<br />

Aunque los resultados obtenidos en esta tesis han aportado sustanciales mejoras<br />

al estudio de las propiedades evolutivas y de agrupamiento de los AGN,<br />

existen aún ciertos aspectos en los que convendría profundizar.<br />

Por un lado, aunque hemos obtenido claras evidencias de que las fuentes extragalácticas<br />

de rayos-X a flujos medios se encuentran fuertemente agrupadas<br />

formando estructuras cosmológicas a gran escala, no hemos podido confirmar<br />

que esto sea cierto para fuentes detectadas en rayos-X duros. En la literatura<br />

es fácil encontrar resultados discrepantes a este respecto, siendo la razón en la<br />

mayor parte de los casos la baja densidad de fuentes por campo.<br />

xix


Haciendo uso del segundo catálogo de fuentes de XMM-Newton, 2XMM 2 , el<br />

mayor catálogo de fuentes de rayos-X hasta la fecha (2XMM es 5 veces mayor<br />

que el anterior catálogo disponible, 1XMM 3 ) que contiene más de 250,000 fuentes<br />

(∼150,000 detecciones únicas) detectadas sobre ∼600 grados cuadrados de cielo,<br />

será posible determinar fuera de toda duda la presencia de agrupamiento angular<br />

de fuentes por encima de 2 keV. Además, gracias a la superior estadística<br />

proporcionada por 2XMM se podrán determinar los parámetros de agrupamiento<br />

con una precisión sin precedentes, y estudiar cómo evolucionan a<br />

distintos flujos límite poniendo de relieve las propiedades de agrupamiento de<br />

las distintas poblaciones de fuentes dominantes a cada flujo.<br />

Por otra parte, las propiedades cosmológicas de las fuentes detectadas en rayos-<br />

X muy duros (en la banda 4.5-7.5 keV) que, como hemos visto, son muy eficientes<br />

para seleccionar muestras de AGN muy poco sesgadas por la absorción,<br />

se encuentran aún sujetas a muchas incertidumbres. En especial, la evolución<br />

de la fracción de AGN absorbidos con el desplazamiento al rojo en esta banda<br />

de energía no se ha podido determinar, y algunos parámetros evolutivos en<br />

la función de luminosidad de estas fuentes requieren un estudio más detallado.<br />

Para resolver estos aspectos es necesario mejorar la estadística de la muestra. Ya<br />

que los AGN absorbidos tienen baja densidad angular, si se quiere disponer de<br />

un número mayor de ellos es necesario muestrear áreas del cielo más grandes.<br />

Correlacionando 2XMM con catálogos ya existentes en otras frecuencias como,<br />

por ejemplo, el Sloan Digital Sky Survey (SDSS) se puede reunir suficiente información<br />

como para ampliar la muestra a flujos altos, con una elevada cobertura<br />

de cielo, que proporcione un número suficiente de AGN oscurecidos que permita<br />

estudiar su evolución a desplazamientos al rojo moderados ( ∼ < 1).<br />

¢¡£¡¥¤§¦©¨¨£¢¡£©£! "#¨£$¥¡%¥&¥ ¢¨'¥(£))#¨<br />

2<br />

¢¡£¡¥¤§¦©¨¨£¢¡£©£! "#¨£$¥¡%¥&¥ ¢¨+*(£))#¨<br />

3<br />

xx


Universidad de Cantabria<br />

Departamento de Física Moderna<br />

CSIC - Universidad de Cantabria<br />

Instituto de Física de Cantabria<br />

Evolution and large scale clustering<br />

of a sample of extragalactic X-ray sources<br />

A dissertation submitted in partial fulfillment of the requirements for<br />

the degree of Doctor of Philosophy in Physics<br />

by<br />

Jacobo Ebrero Carrero<br />

2008


With magic, you can turn a frog into a prince.<br />

With science, you can turn a frog into a PhD and<br />

you still have the frog you started with.<br />

Terry Pratchett<br />

In the beginning, the Universe was created.<br />

This made a lot of people very angry and<br />

has been widely regarded as a bad idea.<br />

Douglas Adams. The Hitchhiker’s Guide to the Galaxy


Acknowledgements<br />

Esta sección es, con toda probabilidad, a la que más tiempo he dedicado de<br />

todas las componen esta tesis. La razón principal es que va a ser la más leída<br />

de toda ella (las estadísticas así lo indican) y, mucho me temo, la única que se<br />

entienda.<br />

Han sido muchos años y muchas personas, por lo que es posible que me deje a<br />

alguien en el tintero. Si es así, espero que el/los afectado/s no se lo tome/n a<br />

mal y acepte/n el agradecimiento implícito en estas líneas.<br />

Me gustaría empezar agradeciendo a mi director de tesis, Francisco Carrera,<br />

todo el esfuerzo que ha puesto en que este trabajo salga adelante, sus consejos<br />

(“más sabe el diablo por viejo que por diablo”) y, sobre todo, por su paciencia.<br />

También quiero hacer pública mi admiración por su capacidad multi-tarea que,<br />

en ocasiones, me ha hecho considerar seriamente la posibilidad de que sea un<br />

replicante.<br />

Igualmente me gustaría agradecer a Xavier Barcons la confianza depositada<br />

y los consejos dados a lo largo de estos años (cuando conseguía parar por el<br />

despacho, que no era todo lo a menudo que él quisiera).<br />

Mi agradecimiento al resto de la gente del grupo (que empezó siendo grupúsculo<br />

y ya se ha convertido en horda) de Astronomía de Rayos-X del IFCA: a<br />

Maite Ceballos (la SUSI) por el apoyo logístico prestado (que no tiene precio<br />

y tampoco habría páginas suficientes para agradecerlo), a mis compañeros de<br />

fatiga Amalia y Ángel (vamos que ya os falta poco), al “infiltrado” Joserra (que<br />

ha visto más de Cantabria en un año que yo en cuatro), a Lánder (perdona por<br />

la puñalada trapera en el Diplomacy), a Rodrigo (por descubrirme las bondades<br />

culinarias de Estrasburgo), y también a los ex-miembros Javier Bussons (nos vemos<br />

en Sigüenza) y Francesca (autora de la mejor tarta de queso que he probado<br />

en mi vida).<br />

xxvii


Un capítulo aparte merecen mis compañeros de despacho a lo largo de estos<br />

años. Desde aquí quiero agradecer a Ibán Cabrillo (por los consejillos informáticos<br />

-“vale, pero antes manda un mail a la USI”- y por pasarme la serie completa<br />

de Gárgolas), a Marcos el Castor (que creo que se lo está pasando pipa en<br />

Cambridge), a Marcos Cruz (sí, sí, tú, el capullo maquiavélico), a Andrés (por<br />

comentar la prensa todas la mañanas y por esas inolvidables noches de poker<br />

-“pero... QK!!”-), al Chals (“que yo mato peña, chorbo”), y a Luis (por asistir impertérrito<br />

a nuestros exabruptos verbales).<br />

Quiero también extender mi agradecimiento al resto de gente del IFCA que he<br />

tenido el placer de conocer: a Diego (por fomentar la ludopatía y el frikismo,<br />

y por sus insondables conocimientos gastronómicos -a lo que ayuda el hecho<br />

de que, al parecer, su estómago también es insondable-), a Chema (por... por...<br />

bueno, por ser Chema), a Patri y a Belén (por el mus y las risas), y a la gente<br />

del piso de abajo Ivan, Marta, Clara y Diego (por las charlas intrascendentes<br />

delante de la máquina del café).<br />

Desde aquí un abrazo a Emmanuel y Almudena, y como a media docena más<br />

de personas (que no voy a enumerar porque seguro que me dejo a alguien y<br />

entonces sí que voy a quedar mal) que me hicieron más llevaderas las estancias<br />

en Munich, con el agravante de que eramos todos compañeros en la Complu<br />

(va a ser verdad que el mundo es un pañuelo). También a mis amigos Sergio,<br />

Pablo, Fede, Ana y Mónica (siempre nos queda Nochevieja).<br />

En un lugar destacable en estos agradecimientos se encuentra toda mi familia<br />

por el cariño y el apoyo incondicional a pesar de la incomprensión que genera<br />

un trabajo como éste, pero muy especialmente a mis padres y a mi abuela<br />

Chencha, y al resto de mis abuelos (que ya no están aquí pero sé que les hubiera<br />

gustado verlo).<br />

Finalmente, quiero agradecer a la gente de la Universidad de Leicester la posibilidad<br />

que me han dado de trabajar unos meses con ellos, que han sido claves<br />

para poder terminar de escribir esta tesis, y a la Spanish Armada del departamento<br />

Silvia, Omaira, Dolores, Jenny (honoríficamente) y, proximamente, Pili<br />

(somos más que ellos, jo, jo, jo).<br />

A todos, muchas gracias.


xxx


Summary<br />

Since its discovery more than 40 years ago, unveiling the origin and nature of<br />

the cosmic X-ray background (CXB) emission has been one of the major questions<br />

in X-ray astronomy. The deepest X-ray surveys conducted by the Chandra<br />

and XMM-Newton imaging observatories have resolved up to a large extent<br />

the extragalactic X-ray emission from 0.1 to 10 keV, confirming the discrete origin<br />

of the CXB. Large campaigns of optical spectroscopic identifications of the<br />

sources detected in these surveys have shown that dominant population of the<br />

extragalactic X-ray sky is a mixture of obscured and unobscured Active Galactic<br />

Nuclei (AGN).<br />

In this work we pursue to confirm the nature and cosmic evolution of the<br />

sources that dominate the CXB emission at medium fluxes. AGN are strong<br />

X-ray emitters, their luminosity being several orders of magnitudes larger than<br />

that of the host galaxy, and are believed to be powered by accretion of matter<br />

onto a central supermassive black hole. Hard X-rays can penetrate through<br />

large amounts of gas and dust without being significantly affected and, hence,<br />

they provide a unique view of the central engine in obscured AGN. Given the<br />

large sensitivity of the current X-ray observatories, it is clear that medium-deep<br />

extragalactic X-ray surveys provide a unique tool to study the cosmic history of<br />

accretion in the Universe.<br />

In the work presented in this thesis we have provided further insight into the<br />

nature and cosmic evolution of the population of AGN that dominate the CXB<br />

emission at intermediate to faint X-ray fluxes at energies from ∼0.1 to ∼10 keV.<br />

In order to do that we have studied the evolutionary and clustering properties<br />

of one of the largest samples of X-ray sources analyzed up to date detected<br />

serendipitously with XMM-Newton, the AXIS survey, and a flux-limited highlyidentified<br />

subsample of AXIS, the XMS survey. In combination with other shallower<br />

and deeper surveys we have constrained with unprecedent accuracy the<br />

xxxi


sky density of extragalactic X-ray sources, resolving up to ∼90% of the CXB<br />

emission into discrete sources and thus confirming that the bulk of the CXB<br />

emission resides at medium X-ray fluxes ∼10 −14 erg cm −2 s −1 .<br />

We have also estudied the large-scale structure of the X-ray Universe at medium<br />

and faint fluxes. For that, we have undertaken two different approaches applied<br />

to the sources of the AXIS survey. One consisted in the calculation of the<br />

angular correlation function which measures the distribution of angular distances<br />

between pairs of sources with respect a random distribution. On the<br />

other hand, we applied a more formal test based on Poissonian statistics. In<br />

both cases, we found strong evidence that AGN detected in soft (2 keV) X-<br />

rays present only marginal clustering or no clustering at all. However, we have<br />

demonstrated that this lack of clustering might be due to the low density of<br />

hard sources per field rather than an intrinsic property of these sources. Additionally,<br />

we have computed the spatial correlation function for the sources in<br />

the XMS subsample, for which we have a very large number of accurate redshift<br />

determinations. We did not find clustering for neither soft nor hard AGN.<br />

This can be explained in terms of the dillution of the clustering signal across the<br />

sky.<br />

Another major aim of these work is to study the cosmic evolution of AGN along<br />

cosmic time. We have carried out an evolution analysis of the sources in the<br />

XMS survey by computing the X-ray luminosity function of these sources. In<br />

order to have intrinsic results, not biased by absorption, we have taken into account<br />

the intrinsic absorption column densities of the sources detected in hard<br />

X-rays., finding that the fraction of absorbed AGN decreases with increasing<br />

luminosity, and increases with redshift. We tested two models for the X-ray luminosity<br />

function. A model in which there is only envolution in luminosity has<br />

been ruled out, being more acceptable a model in which there is evolution in<br />

both the luminosity and the comoving density of AGN. From these results we<br />

have estimated the accretion rate density of the Universe at different epochs,<br />

finding out that the most luminous AGN feed and grow more efficiently than<br />

the less luminous ones, and are fully formed earlier (z ≃ 1.5 − 2).<br />

xxxii


xxxiv


Contents<br />

Resumen de la tesis en castellano<br />

vii<br />

0.1 Objetivos de la investigación . . . . . . . . . . . . . . . . . . . . . vii<br />

0.2 Planteamiento y metodología . . . . . . . . . . . . . . . . . . . . . xi<br />

0.3 Aportaciones originales . . . . . . . . . . . . . . . . . . . . . . . . xvi<br />

0.4 Conclusiones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii<br />

0.5 Futuras líneas de investigación . . . . . . . . . . . . . . . . . . . . xix<br />

Acknowledgements<br />

xxvii<br />

Summary<br />

xxxi<br />

1 Introduction 1<br />

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.2 Brief history of X-ray Astronomy . . . . . . . . . . . . . . . . . . . 3<br />

1.3 The Cosmic X-ray Background . . . . . . . . . . . . . . . . . . . . 6<br />

1.4 The large-scale structure of the Universe in X-rays . . . . . . . . . 8<br />

1.5 Active Galactic Nuclei . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.5.1 Emission mechanisms . . . . . . . . . . . . . . . . . . . . . 10<br />

1.5.2 Observational properties . . . . . . . . . . . . . . . . . . . 11<br />

1.5.3 Origin of the X-ray emission . . . . . . . . . . . . . . . . . 14<br />

1.5.4 Classification of AGN . . . . . . . . . . . . . . . . . . . . . 15<br />

1.5.5 The AGN model and the AGN unification picture . . . . . 17<br />

1.6 Evolution of AGN in X-rays . . . . . . . . . . . . . . . . . . . . . . 21<br />

xxxv


CONTENTS<br />

1.6.1 The K-correction . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

1.6.2 Space density of AGN . . . . . . . . . . . . . . . . . . . . . 22<br />

1.7 Modern instrumentation in X-ray astronomy . . . . . . . . . . . . 23<br />

1.7.1 X-ray telescopes . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

1.7.2 Imaging X-ray detectors . . . . . . . . . . . . . . . . . . . . 24<br />

1.7.3 The XMM-Newton observatory . . . . . . . . . . . . . . . . 27<br />

1.8 Aims of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2 AGN surveys with XMM-Newton 33<br />

2.1 The AXIS survey . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

2.1.1 Introduction and field selection . . . . . . . . . . . . . . . . 34<br />

2.1.2 Data processing and source selection . . . . . . . . . . . . 35<br />

2.1.3 X-ray properties of the sources . . . . . . . . . . . . . . . . 41<br />

2.2 The XMS survey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

2.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

2.2.2 Imaging and selection of optical counterparts . . . . . . . 43<br />

2.2.3 Identification of XMS sources . . . . . . . . . . . . . . . . . 45<br />

2.3 Other surveys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

2.3.1 Chandra Deep Field . . . . . . . . . . . . . . . . . . . . . . . 50<br />

2.3.2 ASCA Medium Sensitivity Survey . . . . . . . . . . . . . . 51<br />

2.3.3 XMM-Newton Bright Serendipitous Survey . . . . . . . . . 51<br />

2.3.4 XMM-Newton Hard Bright Serendipitous Survey . . . . . 51<br />

2.3.5 ROSAT International X-ray/Optical Survey . . . . . . . . . 52<br />

2.3.6 ROSAT Deep Survey - Lockman Hole . . . . . . . . . . . . 52<br />

2.3.7 ROSAT Bright Survey . . . . . . . . . . . . . . . . . . . . . 53<br />

3 Source counts of the AXIS sources 55<br />

3.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

3.2 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

3.2.1 The sky areas . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

xxxvi


CONTENTS<br />

3.2.2 The binned log N − log S . . . . . . . . . . . . . . . . . . . 57<br />

3.2.3 The log N − log S model . . . . . . . . . . . . . . . . . . . . 58<br />

3.2.4 Maximum likelihood fit method . . . . . . . . . . . . . . . 59<br />

3.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

3.4 Contribution to the Cosmic X-ray Background . . . . . . . . . . . 72<br />

3.4.1 Contribution from bright sources and stars . . . . . . . . . 72<br />

3.4.2 Total X-ray background intensity . . . . . . . . . . . . . . . 73<br />

3.4.3 Contribution to the CXB in different flux intervals . . . . . 74<br />

3.4.4 Resolved and unresolved components of the CXB . . . . . 77<br />

3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

4 Clustering of AXIS and XMS sources 83<br />

4.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

4.2 Cosmic variance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

4.3 Angular correlation function of AXIS sources . . . . . . . . . . . . 85<br />

4.3.1 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

4.3.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

4.3.3 Poisson statistics test . . . . . . . . . . . . . . . . . . . . . . 92<br />

4.3.4 Angular correlation of a hardness ratio selected sample . . 99<br />

4.3.5 Inversion of Limber’s equation . . . . . . . . . . . . . . . . 99<br />

4.4 Spatial clustering of XMS sources . . . . . . . . . . . . . . . . . . . 105<br />

4.4.1 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

4.4.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109<br />

4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

5 Luminosity function of XMS sources 115<br />

5.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115<br />

5.2 The N H function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

5.3 Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124<br />

5.3.1 Binned luminosity function . . . . . . . . . . . . . . . . . . 128<br />

xxxvii


CONTENTS<br />

5.3.2 Analytical model . . . . . . . . . . . . . . . . . . . . . . . . 129<br />

5.3.3 Model fitting to soft sources . . . . . . . . . . . . . . . . . . 130<br />

5.3.4 Model fitting to hard and ultrahard sources . . . . . . . . . 131<br />

5.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

5.4.1 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

5.4.2 Accretion history of the Universe . . . . . . . . . . . . . . . 137<br />

5.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140<br />

6 Summary of the results 147<br />

6.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147<br />

6.2 Future prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

A Sensitivity maps 155<br />

References 160<br />

xxxviii


List of Figures<br />

1.1 The electromagnetic window . . . . . . . . . . . . . . . . . . . . . 2<br />

1.2 Spectrum of the extragalactic CXB . . . . . . . . . . . . . . . . . . 7<br />

1.3 Schematic spectrum of a Type-1 AGN . . . . . . . . . . . . . . . . 12<br />

1.4 Schematic diagram of the unified model of AGN . . . . . . . . . . 20<br />

1.5 Light path on a Wolter-1 X-ray telescope . . . . . . . . . . . . . . . 25<br />

1.6 Payload of the XMM-Newton observatory . . . . . . . . . . . . . . 26<br />

1.7 Lightpath of the XMM-Newton telescopes . . . . . . . . . . . . . . 28<br />

2.1 Examples of identified XMS sources . . . . . . . . . . . . . . . . . 49<br />

3.1 Sky area of the AXIS survey . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.2 N(>S) 0.5-2 keV and 2-10 keV best fits . . . . . . . . . . . . . . . . 63<br />

3.3 N(>S) 0.5-4.5 keV and 4.5-7.5 keV best fits . . . . . . . . . . . . . . 64<br />

3.4 dN(S)/dSdΩ and best fit ratio in Soft and hard bands . . . . . . . 67<br />

3.5 dN(S)/dSdΩ and best fit ratio in XID and Ultrahard bands . . . . 68<br />

3.6 N(>S) for two HR selected Soft subsamples . . . . . . . . . . . . . 71<br />

3.7 Relative contribution to the total CXB intensity . . . . . . . . . . . 76<br />

4.1 w(θ) of AXIS in the 0.5-2 keV and 2-10 keV bands . . . . . . . . . 88<br />

4.2 w(θ) of AXIS in the 0.5-4.5 keV and 4.5-7.5 keV bands . . . . . . . 89<br />

4.3 1 − P µ (N θ ) versus θ for the Soft and Hard samples . . . . . . . . . 93<br />

4.4 1 − P µ (N θ ) versus θ for the XID and Ultrahard samples . . . . . . 94<br />

4.5 Correlation tests for a Soft/3 sample . . . . . . . . . . . . . . . . . 96<br />

xxxix


LIST OF FIGURES<br />

4.6 Soft 1 − P µ (N θ ) versus θ in both Galactic hemispheres . . . . . . . 98<br />

4.7 Correlation tests for a HR selected sample . . . . . . . . . . . . . . 100<br />

4.8 Selection function for AXIS soft and hard sources . . . . . . . . . . 103<br />

4.9 Redshift distribution of the real and simulated XMS sources . . . 108<br />

4.10 ξ(r) versus r of XMS sources in Soft and Hard bands . . . . . . . . 110<br />

5.1 Intrinsic N H versus L X4.5−7.5 of XMS sources. . . . . . . . . . . . . . 118<br />

5.2 Observed N H distributions in the hard and ultrahard bands. . . . 120<br />

5.3 Fraction of absorbed Hard AGN as a function of L X and z . . . . . 121<br />

5.4 Fraction of absorbed Ultrahard AGN as a function of L X and z . . 122<br />

5.5 Luminosity-redshift plane of different X-ray surveys . . . . . . . . 126<br />

5.6 Sky area of the survey as a function of flux . . . . . . . . . . . . . 127<br />

5.7 X-ray luminosity function in the soft band . . . . . . . . . . . . . . 134<br />

5.8 X-ray luminosity function in the hard band . . . . . . . . . . . . . 136<br />

5.9 X-ray luminosity function in the ultrahard band . . . . . . . . . . 138<br />

5.10 Comoving density of AGN . . . . . . . . . . . . . . . . . . . . . . . 139<br />

5.11 Accretion rate density . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

5.12 Total accreted mass . . . . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

A.1 ,§-./ vs. crpoisim for all bands . . . . . . . . . . . . . . . . . . . . . 158<br />

xl


List of Tables<br />

2.1 Fields of the AXIS survey . . . . . . . . . . . . . . . . . . . . . . . 39<br />

2.1 Continued . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

2.2 Number of AXIS sources selected in different bands . . . . . . . . 43<br />

2.3 Definition of the XMS samples . . . . . . . . . . . . . . . . . . . . 44<br />

2.4 Spectroscopic facilities used to identify XMS sources . . . . . . . . 46<br />

2.5 Identification completeness of the XMS sample . . . . . . . . . . . 47<br />

2.6 Identification breakdown of the XMS sample . . . . . . . . . . . . 48<br />

3.1 ML fit results to the AXIS source counts . . . . . . . . . . . . . . . 61<br />

3.2 Fit results for HR selected Soft subsamples . . . . . . . . . . . . . 70<br />

3.3 Cosmic X-ray Background intensities . . . . . . . . . . . . . . . . . 75<br />

3.4 Comparison between estimated CXB intensities . . . . . . . . . . 78<br />

4.1 Angular correlation function best χ 2 fits . . . . . . . . . . . . . . . 91<br />

4.2 Spatial correlation lengths for the AXIS soft and hard samples . . 104<br />

5.1 Summary of surveys used for the luminosity function . . . . . . . 125<br />

5.2 Parameters of the X-ray luminosity function. . . . . . . . . . . . . 132<br />

A.1 Summary of the results of the linear fits of crpoisim to ,0-.0/ . . . . 156<br />

xli


Chapter 1<br />

Introduction<br />

In the last few decades, X-ray Astronomy has opened a new window to the<br />

Universe. The capability to observe the sky at such energetic wavelengths has<br />

allowed many astronomers to look at the most violent and obscured side of the<br />

Universe. A X-ray photon, with an energy thousands of times that of an optical<br />

one, can penetrate through large amounts of matter showing phenomena that<br />

were hidden to us before.<br />

X-ray emission can have different origins ranging from ionized plasmas at millions<br />

of degrees to relativistic particles trapped inside strong magnetic fields,<br />

and therefore many astronomical objects are known to be X-ray emitters. Some<br />

of them lay inside our own Galaxy such as stars, binary systems with a compact<br />

companion (neutron stars, black holes) or supernova remnants, but others<br />

are extragalactic: active galactic nuclei (AGN, hereafter) or hot gas in galaxy<br />

clusters.<br />

The origin of the X-ray emission from the intra-cluster medium in clusters of<br />

galaxies is thermal bremsstrahlung from a low-density, optically-thin hot (millions<br />

of degrees) gas, whereas in the case of AGN and binary systems the radiation<br />

comes from matter gravitationally accreting onto a compact object.<br />

In this thesis, we will study the properties of a sample of AGN using the X-ray<br />

data provided by the European X-ray Observatory XMM-Newton.<br />

1


CHAPTER 1: INTRODUCTION<br />

Figure 1.1 The electromagnetic window. The vertical axis shows the transparency of<br />

the atmosphere for a given frequency. The higher the curve, the more opaque is the<br />

atmosphere at that frequency.<br />

1.1 Motivation<br />

The atmosphere of the Earth absorbs X-rays preventing them to reach the surface<br />

of the planet, effectively absorbing all the electromagnetic radiation coming<br />

from space, except a small window at optical and near infrared wavelengths<br />

and radiowaves (see figure 1.1). This has important biological implications but<br />

it is a major problem when trying to do high-energy Astronomy unless X-ray<br />

detectors are placed outside the atmosphere. The first steps in this field were<br />

done through stratospheric balloons followed by rockets in the second half of<br />

the last century.<br />

X-ray Astronomy was born in 1962, in a rocket flight that tried to detect the<br />

X-ray emission from the Sun reflected in the Moon. The result of this mission<br />

was the detection of the first extra-solar X-ray source (named Sco X-1 as it was<br />

discovered in the constellation of Scorpius) and an isotropic emission coming<br />

from all directions in the sky ([78]). We know today that this emission is the<br />

Cosmic X-ray Background (CXB hereafter, see section 1.3).<br />

Observations with the modern X-ray observatories XMM-Newton and Chandra<br />

2


CHAPTER 1: INTRODUCTION<br />

have confirmed that the CXB emission is a composition of a miriad of individual<br />

sources, most of them AGN. These objects are usually extremely bright in<br />

X-rays, with luminosities in the range 10 42 − 10 48 erg s −1 accounting for an important<br />

fraction (∼10%) of their bolometric luminosity. Variability has been observed<br />

in the X-ray flux of many of these sources with temporal scales ranging<br />

from days to minutes (see [179] for a review), which means that the origin of the<br />

X-ray emission must be very close to the objects lying in their centres. Nowadays,<br />

it is widely believed that almost all known galaxies host in their centres<br />

a supermassive black hole (i.e. [126]) with a mass of the order of 10 6 − 10 9 M ⊙ ,<br />

and therefore the X-ray emission is strongly linked with it, generated by gravitational<br />

accretion.<br />

In this process, matter falls onto a compact and massive object. Since the infalling<br />

material possesses considerable specific angular momentum, it forms a<br />

disk (the so-called accretion disk) around the supermassive black hole. X-ray<br />

emission is then generated in this environment when photons gain energy by<br />

inverse Compton scattering in the hot corona that surrounds the disk and thus<br />

are enhanced to X-ray frequencies (see section 1.5.1).<br />

X-rays are very penetrating and can pass through large amounts of matter that<br />

would easily absorb optical or ultraviolet photons. This makes possible to detect<br />

AGN through cosmological distances and to explore the Universe when it<br />

was younger. Hence, X-rays allow us to trace the evolution of the accreting supermassive<br />

black holes across cosmic time. Furthermore, hard X-rays are the<br />

only ones that can escape from obscured AGN, those with intrinsic absorbing<br />

column densities N H 10 22 cm −2 (type-2 AGN, see section 1.5.4). However,<br />

if the absorption column density is 5 × 10 24 cm −2 , X-ray photons with energies<br />

below 10-20 keV are also absorbed (Compton-thick objects). Recent results<br />

seems to confirm that obscured AGN may play a key role to explain the overall<br />

CXB emission.<br />

1.2 Brief history of X-ray Astronomy<br />

After the first rocket missions that probed the X-ray sky in the 1960s, it was<br />

clear that only with instruments onboard satellites orbiting the Earth would be<br />

possible to observe the X-ray sky. The first major advance in X-ray Astronomy<br />

came from the launch of the mission SAS-1 (NASA), unofficially named Uhuru<br />

3


CHAPTER 1: INTRODUCTION<br />

(the Swahili word for Freedom) in 1970. This small satellite observed hundreds<br />

of sources in the range of 2-20 keV, many of them in the galactic plane as they<br />

were mainly binary systems or supernova remnants, but the rest were extragalactic<br />

sources, homogeneously distributed on the celestial sphere. They were<br />

mainly Seyfert galaxies (see section 1.5.4) and galaxy clusters, which meant that<br />

most of the cosmic X-ray emission came from very far objects and therefore the<br />

existence of very energetic physical processes was required. In addition, Uhuru<br />

was capable of observing the sources for relatively large amounts of time (compared<br />

with those of the rocket experiments) which led to the first discovered<br />

variable sources.<br />

After Uhuru the observation of the X-ray sky was continued by a series of small<br />

satellites such as SAS-3 or OSO-8. In 1977 a new leap forward in the history<br />

of X-ray Astronomy was given with the launch of the first of the High Energy<br />

Astronomy Observatories (HEAO-1). Among other discoveries, HEAO-1 measured<br />

the CXB intensity spectrum in the range 10-100 keV, showing a broad<br />

peak at energies 40 keV. This hump was fitted to a thermal bremsstrahlung<br />

model under the assumption of a diffuse origin for the CXB ([146]). However,<br />

when the satellite COBE measured the cosmic microwave background with the<br />

instrument FIRAS in 1992, its emission spectrum perfectly fitted with a black<br />

body spectrum showing no traces of Compton scattered microwave photons<br />

by hot electrons ([152]). This meant that the diffuse origin hypothesis for the<br />

CXB had to be abandoned.<br />

Although the instruments onboard HEAO-1 had an improved sensitivity with<br />

respect the previous missions, it was not until 1978 with HEAO-2 (renamed Einstein<br />

afterwards) that a modern X-ray telescope was finally put into orbit. The<br />

construction and installation of the four nested mirrors in the satellite was a<br />

major technological challenge for the epoch. It could focus X-ray photons with<br />

energies between 0.25 and 4 keV with an angular resolution of several arcseconds.<br />

Einstein detected over 5000 sources and performed the first deep imaging<br />

observations (to a limiting flux of ∼ 2 × 10 −14 erg cm −2 s −1 ) that resolved 25%<br />

of the CXB in the 1-3 keV band into discrete sources ([77], [91], [96], [186]).<br />

The first important European mission was Exosat, which was launched in 1983<br />

and was operative until 1986. Its very eccentric orbit with an orbital period<br />

of 90 hours, allowed to observe the same source for some days and therefore to<br />

study in detail their variability. In fact, the most remarkable discovery achieved<br />

4


CHAPTER 1: INTRODUCTION<br />

by this mission was the existence of quasiperiodic oscillations in the X-ray spectrum<br />

of binary systems.<br />

In 1990 was launched the mission Rosat, a collaboration between Germany, the<br />

United States and the United Kingdom. It operated in the range 0.1-2.4 keV<br />

and, compared with Exosat, increased its sensitivity and field of view (the area<br />

of the sky visible from a telescope). The prime task of Rosat was to carry out an<br />

all-sky survey, something that had not been done since HEAO-1, which yielded<br />

to a catalogue of ∼ 77000 sources. Besides, Rosat performed a deep observation<br />

in the Lockman Hole region resolving up to ∼ 90% of the CXB into discrete<br />

sources down to a flux limit of ∼ 10 −15 erg cm −2 s −1 ([98], [102]) in that energy<br />

band.<br />

The Japanese Space Agency participated actively in the investigation of the X-<br />

ray sky in the 1980s with a couple of two small missions (Tenma, Ginga) but it<br />

was in 1993 with Asca when they fully gave a technological leap forward. This<br />

satellite could observe in a range of energies spanning from 0.5 to 10 keV while<br />

retaining a great effective area. The X-ray detectors placed onboard were based<br />

on the CCD technology (see section 1.7.2) for the first time. Asca carried out<br />

some medium/deep surveys in the 2-10 keV band down to limiting fluxes of<br />

∼ 5 × 10 −14 erg cm −2 s −1 that resolved ∼ 35% of the CXB in this energy band<br />

([69], [74], [226]). These results showed that there was a hidden population<br />

of AGN with harder spectra than the AGN detected at lower energies, many of<br />

them turned out to be X-ray absorbed AGN at low redshifts. This meant that the<br />

unresolved fraction of the CXB at harder energies could be due to a population<br />

of undetected obscured AGN. In addition, the gravitationally redshifted Fe K α<br />

fluorescence emission line was observed with better sensitivity and spectral<br />

resolution (which was first discovered by Ginga) thus proving the existence of<br />

matter moving at relativistic velocities around a supermassive black hole.<br />

In 1996, the Italian Space Agency in collaboration with the Dutch Space Agency<br />

launched the BeppoSAX satellite. Among its achievements was to survey a region<br />

of the sky down to fluxes of ∼ 5 × 10 −14 erg cm −2 s −1 in the 5-10 keV band<br />

resolving ∼ 20 − 30% of the CXB ([40]) and the observation of highly-absorbed<br />

Compton-thick AGN.<br />

The most powerful X-ray observatories up to now, Chandra (NASA) and XMM-<br />

Newton (ESA), were both launched in 1999 and are still operative. Their designs<br />

are very different: Chandra has an unprecedent angular resolution that makes<br />

5


CHAPTER 1: INTRODUCTION<br />

it ideal for imaging, whereas XMM-Newton possesses a larger effective area<br />

(sensitivity) that is better for X-ray spectroscopy. Both of them have reached<br />

limiting fluxes up to ∼ 100 times fainter than the precedent observatories. For<br />

instance, Chandra reached ∼ 2 × 10 −16 erg cm −2 s −1 in the ultra-deep surveys<br />

Chandra Deep Field North (CDF-N, [28]) and South (CDF-S, [79], [194]) resolving<br />

up to a ∼ 90% of the CXB in the 2-8 keV band ([33], [80], [115], [156], [163],<br />

[219]). These surveys have been used in this thesis to complement our X-ray<br />

data (see section 2.3). On the other hand, XMM-Newton reached a 0.5-2 keV<br />

limiting flux of ∼ 5 × 10 −16 erg cm −2 s −1 in the Cosmos survey ([134]) and the<br />

Lockman Hole region ([101]). However, at energies above 10 keV, where the<br />

bulk of the CXB is emitted, only a tiny fraction of it has been resolved into discrete<br />

sources, probably heavily obscured AGN that remain undetected until a<br />

new generation of X-ray telescopes is developed.<br />

Since 1999, when XMM-Newton began to operate, each of its pointings have<br />

provided between ∼ 30 − 150 serendipitous sources, most of them newly detected.<br />

The 1XMM catalogue was released in 2003 and contained all the sources<br />

(∼ 33, 000 in total, ∼ 28, 000 unique sources) detected by XMM-Newton before<br />

May 2002. In 2007, the XMM-Newton Survey Science Center (SSC) 1 released the<br />

2XMM catalogue, the largest X-ray catalogue ever made, comprising ∼ 250, 000<br />

sources, ∼ 150, 000 of them unique 2 .<br />

1.3 The Cosmic X-ray Background<br />

The X-ray sky is dominated by a diffuse radiation of cosmic origin. Since its discovery<br />

(see section 1.2) the hypothetical origin of the CXB has been a challenge<br />

for the X-ray astronomers.<br />

The CXB spans from ∼ 0.1 keV to several keV (see Fig. 1.2) and the contribution<br />

at different energies has different origins. For instance, at very soft energies,<br />

0.1-0.3 keV, the main contribution comes from thermal emission from a local<br />

hot plasma (with T ∼ 10 6 K) known as the Local Hot Bubble. A second contributor<br />

is the thin plasma in the halo of our Galaxy ([153]). From 0.5 to 1 keV<br />

there is contribution to the CXB emission from both the Galactic hot plasma<br />

and extragalactic discrete sources. Above ∼3 keV the CXB is highly isotropic<br />

¢¡£¡¥¤§¦©¨¨£¢¡£©£! "#¨<br />

1<br />

2 A fraction of the sources has been detected more than once because of the overlapping in XMM-<br />

Newton’s observations.<br />

6


CHAPTER 1: INTRODUCTION<br />

Figure 1.2 Spectrum of the extragalactic X-ray background from 0.2 to 400 keV measured<br />

by different missions (taken from [82])<br />

(once the weak Galactic component and the dipolar field caused by the motion<br />

of the Galaxy have been conveniently taken into account) and it is known to<br />

have a extragalactic origin, due to the integrated emission of unresolved discrete<br />

sources.<br />

None of the missions and experiments that have measured the CXB spectrum to<br />

date have found significant deviations in the spectral shape from Γ ∼ 1.4 over<br />

the ∼3-10 keV band. However, the values obtained for its intensity (usually<br />

parameterized as the intensity at 1 keV in units of photons keV −1 cm −2 s −1<br />

sr −1 ) have been significantly different (sometimes even among measurements<br />

taken with the same instrument) with variations that can be as large as ∼40%,<br />

which makes the value of the normalization of the CXB rather uncertain.<br />

It is expected that the discrete nature of the CXB induces a cosmic variance in<br />

the measured value of the CXB intensity. The discrete nature of the CXB produces<br />

confusion noise, caused by unresolved sources in the images ([11], [202]).<br />

All measurements, except the one carried out by HEAO-1, were obtained from<br />

observations over small solid angles (below ∼1 deg 2 ). [14] showed that confusion<br />

noise can account for variations in the measured intensity of the CXB<br />

of ∼10% for solid angles below 1 deg 2 . When cosmic variance is taken into<br />

7


CHAPTER 1: INTRODUCTION<br />

account measurements within the same mission but different instruments are<br />

brought to consistency. However, they also found that cosmic variance alone<br />

cannot acount for the observed discrepancies between missions, and hence significant<br />

instrumental cross-calibration uncertainties between different instruments<br />

must still be present ([52]).<br />

1.4 The large-scale structure of the Universe in X-rays<br />

In the last decades there has been substantial advances in the study and understanding<br />

of the cosmic structure of the Universe and the large-scale matter<br />

distribution (see [214] for a review). New generations of galaxy surveys, such as<br />

the 2-degree Field Galaxy Survey ([39]) or the Sloan Digital Sky Survey (SDSS,<br />

[243]), have quantified the distribution of galaxies in the local Universe. These<br />

wide-angle galaxy surveys have shown the existence of large-scale structure in<br />

the form of filaments around voids with extensions of ∼20-30 h −1 Mpc or forming<br />

superstructures like the ’Great Wall’ ([73]). Such structures have been found<br />

up to distances of z ≃ 0.2. However, the study of these structures at higher redshifts<br />

requires the involvement of AGN which, thanks to their large bolometric<br />

luminosities (see section 1.5), can be detected at cosmological distances. Since<br />

strong X-ray emission is a typical feature of AGN activity and it is less likely<br />

to be affected by absorption, X-ray surveys provide the most efficient mean of<br />

investigating the large-scale structure at high redshifts.<br />

The standard way of describing the clustering of galaxies, clusters of galaxies<br />

and AGN in statistical cosmology have been, for many years, the correlation<br />

functions. They were first suggested in the 1960s by [218] to study the spatial<br />

clustering of galaxies in the near Universe. Nonetheless, the major boost for the<br />

correlation functions as a tool to measure large-scale clustering came up in the<br />

1970s, with the works of P. J. E. Peebles, along with several colleagues. They<br />

carried out a programme to extract estimates of the correlation functions from<br />

the Lick galaxy catalogue and other data sets (see [174] and references therein).<br />

The correlation functions set up a description of the clustering properties of a<br />

set of points distributed in space, although useful results can also be obtained<br />

for two-dimensional distributions of positions on the celestial sphere. A number<br />

of works have used the correlation functions of X-ray selected AGN trying<br />

to trace the underlying cosmic large-scale structure ([24], [18], [38], [84], [159],<br />

8


CHAPTER 1: INTRODUCTION<br />

[161], [242]), although the significancies on the measurements vary and the results<br />

show very large uncertainties in some cases. In addition, other authors<br />

have found similar results for X-ray selected clusters of galaxies (e.g. [206]),<br />

whose clustering properties have been claimed to be linked with that of AGN<br />

([47], [35], [34]) and therefore with the processes of formation of galaxies and<br />

supermassive black holes.<br />

1.5 Active Galactic Nuclei<br />

Active Galactic Nuclei (AGN) were first discovered in the 1940s as galaxies<br />

characterized by the extreme brightness of their centres. The luminosity of<br />

the centre can range from tens to thousands of times that of the underlying<br />

host galaxy, which is completely outshone, appearing in the optical images as<br />

a bright and variable point source when the AGN is distant. The suggested<br />

activity of the central regions of these galaxies led to the term Active Galactic<br />

Nuclei. These objects are the most luminous persistent sources known in the<br />

Universe (gamma ray bursts can be brighter by a few orders of magnitude but<br />

they only last a few seconds). Their bolometric luminosities span a broad range<br />

of values, from 10 42 to 10 48 erg s −1 for the most luminous AGN often known<br />

as quasars. This term was originally applied to radio-loud objects (quasi-stellar<br />

radio source), while QSO was used to radio-quiet objects. More recently, the<br />

term quasar has been frequently used to refer to all luminous AGN, although<br />

we now know that many properties are similar for both low and high luminosity<br />

objects.<br />

With the improvement of sensitivity of X-ray detectors, it has been possible to<br />

detect activity in an increasing number of non-active galaxies, somehow blurring<br />

the division between active and non-active galaxies. There is observational<br />

evidence that supermassive black holes are hosted in the centres of most, if<br />

not all, galaxies (i.e. [126]), including the Milky Way. It is customary in X-ray<br />

astronomy to classify a galaxy as active if its X-ray luminosity is higher than<br />

∼10 42 erg s −1 since nothing but gravitational accretion of matter can produce<br />

such luminosities.<br />

9


CHAPTER 1: INTRODUCTION<br />

1.5.1 Emission mechanisms<br />

The mechanisms that are believed to produce the observed emission in AGN at<br />

different wavelengths can be summarized as follows:<br />

1. Thermal black body emission: A black body is defined as a body in thermal<br />

equilibrium with its surroundings that is both a perfect absorber and a<br />

perfect emitter of radiation. None of the radiation that reaches a black<br />

body is reflected. Real objects are not perfect black bodies, but the black<br />

body approximation works reasonably well in many cases. The emission<br />

spectrum of a black body depends only on its temperature.<br />

Thermal radiation by the gas present in the accretion disk has been traditionally<br />

used as an explanation to the optical/UV continuum emission in<br />

AGN. AGN accretion disks are typically cool (∼10 5 - 10 6 K) producing relatively<br />

soft spectra in the optical/UV energy range. Black body emission<br />

coming from dust heated by the optical/UV radiation field is believed to<br />

be responsible for the mid-/far-IR emission in radio-quiet AGN.<br />

2. Inverse Compton scattering: Photons can interchange energy by interactions<br />

with charged particles, a process known as Compton scattering. The scattering<br />

cross-section of nuclei is much smaller than that of the electrons<br />

and therefore Compton scattering by nuclei can be neglected. If the electrons<br />

are relativistic, photons may gain energy from the electrons and the<br />

process is known as Inverse Compton scattering. The spectrum of the<br />

scattered radiation depends on the electron energy distribution. Inverse<br />

Compton scattering is believed to be responsible for the broad band continuum<br />

observed in the X-ray spectra of radio-quiet AGN.<br />

3. Synchrotron radiation: Electromagnetic radiation produced by relativistic<br />

charged particles accelerated in an external magnetic field is known as<br />

synchrotron radiation.<br />

Assuming a non-thermal particle distribution with a power law energy<br />

spectrum, the emitted spectrum is another power law with an energy index<br />

α (F ν ∝ ν −α ) that is related to the power law of the particle spectrum.<br />

It is believed that synchrotron radiation from a population of relativistic<br />

electrons located in jets is responsible for the radio emission in radio-loud<br />

AGN. However, it is very unlikely that synchrotron radiation is an impor-<br />

10


CHAPTER 1: INTRODUCTION<br />

tant contributor to the high energy spectra of AGN, excepting the small<br />

fraction of radio-loud AGN.<br />

4. Bremsstrahlung (free-free emission): Bremsstrahlung (which is the German<br />

word for ’braking radiation’) is related to emission produced in hot dense<br />

ionized plasma. It is produced when free electrons lose energy and therefore<br />

emit electromagnetic radiation when they are accelerated by the electromagnetic<br />

field of the ions. As free electrons can have a very large range<br />

of energies, the emission has a thermal continuum distribution. If the material<br />

contains metals, there are might also be very intense emission lines<br />

superimposed to this continuum.<br />

The emission mechanism on clusters of galaxies is believed to be bremsstrahlung<br />

from hot (∼10 8 K), low-density (∼10 −3 atoms cm −3 ) gas.<br />

1.5.2 Observational properties<br />

The broadband emission of AGN is very different to that of normal galaxies,<br />

and spans the entire electromagnetic spectrum from radio to gamma rays (unless<br />

stated otherwise, all energies mentioned in this section are rest-frame energies):<br />

1. Radio emission: Only ∼10% of optically and X-ray selected AGN are strong<br />

radio emitters, and the emission in radio rarely contributes more than<br />

∼1% to the AGN bolometric luminosity. The spectrum of AGN at radio<br />

wavelengths is a power law, and the origin of the radio emission is<br />

believed to be synchrotron radiation emitted in outflows of relativistic<br />

charged particles along the poles of a rotating supermassive black hole<br />

(see section 1.5.1).<br />

2. Infrared emission: The IR continuum of AGN rise following a power law<br />

from the optical to the FIR/submm wavelengths. The IR broad band<br />

spectrum can be well reproduced by a model in which a significant fraction<br />

of the nuclear emission is reprocessed by warm molecular gas that<br />

surrounds the active nucleus and the broad line region (see section 1.5.5)<br />

with an axisymmetric torus-like structure ([87], [181], [182]). The gas absorbs<br />

the optical/UV/X-ray radiation and thermally reprocesses it into<br />

the rest-frame mid- and far-IR range. Obscuration in a moderately thick<br />

11


CHAPTER 1: INTRODUCTION<br />

Figure 1.3 Schematic spectrum of a type-1 AGN (black solid line) along with the most<br />

relevant spectral components (taken from [191]).<br />

gas (optical depth from 10 to 300 in the UV) and at maximum distance<br />

to the black hole of tens to hundreds of parsecs, gives a good description<br />

of the broadband IR spectrum of AGN ([88]). This scenario has gained<br />

strong observational support as some observations of the Hubble Space<br />

Telescope to the active galaxy NGC 4261 have resolved a torus structure<br />

([68]).<br />

3. Optical/UV emission: The optical/UV continuum emission of AGN is dominated<br />

by a broad hump known as the Big Blue Bump (BBB), which peaks<br />

at ∼1000 Å. The BBB is usually interpreted as thermal emission from the<br />

accreting matter in the innermost regions of the AGN. The optical/UV<br />

spectrum of Type-1 AGN (see section 1.5.4) also shows strong broad permitted<br />

lines and narrow permitted and forbidden emission lines superimposed<br />

to the continuum.<br />

4. X-ray emission: AGN are very luminous X-ray sources. As a first approximation,<br />

the X-ray spectrum of AGN from ∼1 to ∼100 keV is continuumdominated,<br />

and is well described by a single power law with a typical<br />

slope of ∼1.9. At low-energy gamma rays the spectrum of AGN is also<br />

well reproduced by a power law albeit with a steeper slope. If the signal-<br />

12


CHAPTER 1: INTRODUCTION<br />

to-noise ratio of the X-ray spectrum is good enough, other prominent features<br />

are observed (see Figure 1.3). The most common spectral features<br />

are:<br />

• Photoelectric absorption: The strongest feature in the 0.1-100 keV spectrum<br />

of a significant fraction of AGN is a photoelectric cut-off of the<br />

intrinsic power law at low energies due to absorption by a cold or<br />

partially ionized material in the line of sight to the nucleus. Photoelectric<br />

absorption by our own Galaxy modifies the X-ray spectrum<br />

below 0.6 keV. On the other hand, local absorption lines have been<br />

detected in the X-ray spectrum of bright AGN, which are believed to<br />

be caused by hot plasma (10 5 - 10 7 K) in the so-called Warm-Hot Intergalactic<br />

Medium (WHIM). Apart from this, 50% of Seyfert 1 galaxies<br />

(see section 1.5.4) show X-ray absorption by highly ionized gas by<br />

the intense nuclear UV/X-ray radiation (warm absorber, [76], [189]).<br />

These absorption edges usually appear blueshifted, showing velocities<br />

of the order of a few hundreds km s −1 thus suggesting that the<br />

absorbing material is moving away the central source at these velocities<br />

([120], [121]).<br />

• Soft excess: At low energies (below ∼2 keV) the emission of a significant<br />

fraction (∼30%) of AGN shows a strong soft component superimposed<br />

over the hard X-ray continuum with a peak of emission in<br />

the Extreme UV ([223]), known as the soft excess ([235]). This emission<br />

was first reported by [8] in 1985, but the origin of the spectral<br />

component is not clear. It has been often interpreted as direct thermal<br />

emission from the hot innermost region of the accretion disk ([8], [46],<br />

[132], [185], [195]) and therefore as the high-energy tail extension of<br />

the BBB. However, the typical measured temperatures (>500 eV) are<br />

too high to explain it as thermal emission from the accretion disk. Alternative<br />

models for this emission are reflection of hard X-rays by cold<br />

dense gas ([94]) or, more recently, by ionized absorption of the primary<br />

continuum by a relativistic wind moving at several thousands<br />

km s −1 ([81]).<br />

• Fluorescent Fe K α emission: Emission lines at rest-frame energies ∼6.4 -<br />

6.7 keV are common in high quality X-ray spectra of AGN. Since iron<br />

is the most abundant heavy element, the observed line emission in X-<br />

13


CHAPTER 1: INTRODUCTION<br />

rays has been attributed to iron fluorescence. Fe K α emission has been<br />

detected in both absorbed and unabsorbed AGN ([162], [164], [184]).<br />

The material responsible for the line emission in absorbed objects is<br />

likely to be the molecular torus responsible for the X-ray absorption.<br />

In unobscured AGN the emission originates in optically thick material<br />

out of the line of sight.<br />

• Compton reflection hump: A broad bump above the power law continuum<br />

is observed in many AGN over the ∼7 - 60 keV rest-frame<br />

energy range, peaking at ∼30 keV ([165]). This spectral feature is often<br />

explained as reflection of the hard X-ray photons in the accretion<br />

disk or the molecular torus (i.e. [75], [93], [136]).<br />

• Exponential cut-off : At energies of several tens to hundreds of keV an<br />

exponential cut-off has been obverved in the BeppoSAX spectra of a<br />

few AGN ([92], [166], [176], [177]). This cut-off indicates the presence<br />

of a thermal hot corona around the accretion disk.<br />

1.5.3 Origin of the X-ray emission<br />

In this section we will summarize some of the properties already discussed in<br />

order to put them in the context of the origin of the X-ray emission of AGN. We<br />

know that AGN are cosmological sources that, because of their large bolometric<br />

luminosities, can be detected at very large distances. The large emissivity<br />

of these objects puts very strong constraints on the emission mechanisms that<br />

powers them (see section 1.5.1). It is commonly accepted that the source of energy<br />

of AGN is the conversion into radiation of gravitational energy of accretting<br />

matter in the region surrounding a supermassive central black hole with<br />

masses in the range ∼10 6 - 10 10 M ⊙ . The efficiency of the proccess can be of<br />

the order of ∼10%, much higher than any other source of energy (for instance,<br />

nuclear fusion, the power source in stars, has an efficiency of ∼0.7%).<br />

In spite of this commonly accepted scenario, the details on how X-rays are produced<br />

are still not clear and remain as one of the most challenging problems in<br />

AGN physics. As seen in section 1.5.2, an optically thick accretion disk is not<br />

hot enough to form the soft excess emission often seen in the spectrum of AGN<br />

and therefore it cannot produce the higher X-ray emission observed.<br />

The best model to date for the primary X-ray emission above 1 keV in AGN is<br />

14


CHAPTER 1: INTRODUCTION<br />

Compton up-scattering of optical/UV photons emitted thermally in the accretion<br />

disk by an optically thin hot plasma of thermal electrons that surrounds the<br />

accretion disk, which is modelled as a Comptonizing corona. Comptonization<br />

leads to a power law spectral shape that reproduces well the observed continuum<br />

emission of AGN in X-rays ([46], [94], [95], [133], [196]).<br />

In the simplest geometry of the Comptonization model, half of the X-ray radiation<br />

is emitted towards the observer while the other half is directed to the accretion<br />

disk, where it is absorbed and re-emitted as optical/UV photons that enter<br />

again the corona thus cooling the hot electrons. Hence, the spectral slope of the<br />

X-ray power law depends on the temperature and optical depth of the Comptonizing<br />

plasma. Heating and cooling processes are coupled so the Comptonizing<br />

models predict a narrow range of power law shapes, which is in agreement<br />

with the observations that have found rather small dispersions on the X-ray<br />

continuum of AGN (∆Γ ∼0.2, [149]).<br />

Thermal models for the corona predict a high energy cut-off at energies that depend<br />

on the temperature of the Comptonizing plasma. This has already been<br />

observed in the BeppoSAX spectra of a number of AGN (see section 1.5.2), giving<br />

strong support to the thermal models.<br />

1.5.4 Classification of AGN<br />

The different observational properties of AGN have been described in previous<br />

sections, which have been commonly used to classify AGN into different<br />

groups. The scheme adopted for this work is based on their emission properties<br />

at optical/UV wavelengths. AGN are often classified using a secondary scheme<br />

based on their emission properties at radio wavelengths (radio-loudness).<br />

AGN are classified according to their optical/UV spectral properties, falling<br />

into the following categories:<br />

• Broad Line AGN (Type-1 AGN): The optical spectrum of a Type-1 AGN<br />

shows strong highly ionized broad permitted lines and hydrogen lines,<br />

although HeI, HeII and FeII lines are also frequently detected with typical<br />

widths of 2000-10000 km s −1 . Because of this, these objects are often<br />

referred as Broad Line AGN (BLAGN). They cover a broad range of redshifts<br />

and are the dominant population at bright fluxes ([217]). Most AGN<br />

in this group do not show signs of obscuration at neither X-ray nor optical<br />

15


CHAPTER 1: INTRODUCTION<br />

wavelengths. Some observational evidences ([149]), however, point out<br />

that ∼10% of the sources in this group have X-ray absorption, although<br />

with low absorbing column densities (below 10 22 cm −2 ).<br />

The radio-quiet objects in this class are usually subdivided into Seyfert 1<br />

galaxies, when they have luminosities below 10 44 erg s −1 (and are therefore<br />

seen nearby), and QSO, when they have luminosities above 10 44 erg<br />

s −1 (which appear as point-like sources in both X-ray and optical images).<br />

• Narrow Line AGN (Type-2 AGN): These sources have optical continua with<br />

superimposed permitted and forbidden lines, being all the observed lines<br />

narrow (widths ≤1500 km s −1 ). Their optical spectral continuum is dominated<br />

by starlight and therefore has very weak non-thermal optical/UV<br />

continuum, with little evidence for the optical/UV bump typical of Type-<br />

1 AGN. The X-ray spectrum of a Type-2 AGN usually shows signatures<br />

of X-ray absorption, with a wide range of absorbing column densities but<br />

systematically higher than those of Type-1 AGN. Hence, they are fainter<br />

soft sources, detected at z ≤ 1.5 in current X-ray surveys ([12]).<br />

Like in the Type-1 category, Type-2 radio-quiet AGN are classified as Seyfert<br />

2 galaxies at low luminosities and as Type-2 QSO at high luminosities. A<br />

general name used for all galaxies which only show narrow emission lines<br />

in their optical/UV spectrum is Narrow Emission Line Galaxies (NELG,<br />

[162]), which includes Seyfert 2 galaxies, LINERs, starburst galaxies and<br />

Type-2 QSO.<br />

• Other categories: A small number of AGN are characterized by the absence<br />

of emission lines in their optical/UV spectrum. The radio-loud sources in<br />

this group include BL Lacertae (BL Lac) objects, Optically Violently Variable<br />

Quasars (OVV), Highly Polarized Quasars (HPQ) and Core Dominated<br />

Quasars (CDQ). All of them are characterized by rapid variability<br />

in flux at very short time scales and are generally termed as blazars.<br />

One of the most surprising findings of the Chandra and XMM-Newton<br />

medium-deep surveys was the discovery of X-ray luminous sources (≥<br />

10 42 erg s −1 ) in the nuclei of optically normal galaxies ([15], [41], [79],<br />

[114], [163]). These sources, which present an optical spectrum typical of<br />

an absorption line galaxy (ALG), have been given a variety of names such<br />

as Optically Dull Galaxies ([63]), Passive Galaxies ([90]), X-ray Bright Optically<br />

Normal Galaxies (XBONG, [41]) or Elusive Active Galactic Nuclei<br />

16


CHAPTER 1: INTRODUCTION<br />

([142]). It is uncertain why we do not see AGN signatures in the optical<br />

spectra of these sources, although a possible explanation is the combination<br />

of absorption, faint nuclear emision with respect to the host galaxy,<br />

and an inadequate set-up and atmospheric conditions during the spectroscopic<br />

observations ([208]). Additionally, some of these sources show absorbed<br />

X-ray spectra with column densities higher than 10 24 cm −2 ([142])<br />

which suggests that these sources indeed host an AGN and that the peculiar<br />

optical and X-ray properties are related to heavy obscuration.<br />

A secondary classification of AGN classify them into two broad categories according<br />

to their radio-loudness, defined as:<br />

( )<br />

f5<br />

R L = log<br />

f B<br />

(1.5.1)<br />

where f 5 is the radio flux at 5 GHz and f B is the optical flux on the B band,<br />

centred at 4400 Å. Those objects with R L ≥ 10 are labeled as radio-loud AGN<br />

whereas the rest are radio-quiet AGN. The optical/UV emission line spectra of<br />

both classes of objects are very similar ([200]), although radio-loud AGN appear<br />

to have systematically flatter X-ray spectral slopes than the radio-quiet objects<br />

([210], [235]). Radio-loud AGN constitute a small fraction (5-10%) of the optical<br />

and X-ray selected AGN population.<br />

The radio-loud Type-1 AGN are named Broad Line Radio Galaxies at low luminosities,<br />

and Radio Loud Quasars (RLQ) at high luminosities. Radio-loud Type-<br />

2 AGN are usually termed Narrow Line Radio Galaxies (NLRG), and are subdivided<br />

into two categories according to their morphological type: the low luminosity<br />

Fanaroff-Riley 1 radio galaxies (FRI) and the high luminosity Fanaroff-<br />

Riley 2 radio galaxies (FRII, [67]). Radio emission in FRI objects is found to be<br />

core dominated and show weak optical emission, while the FRII objects show<br />

strong optical emission and their radio emission is lobe dominated.<br />

1.5.5 The AGN model and the AGN unification picture<br />

The large diversity of observational properties of AGN makes the detailed understanding<br />

of the underlying physic very complex. However, the approximate<br />

structure of AGN is believed to be as illustrated in Figure 1.4 and it is widely<br />

accepted that all active galaxies are constituted by the following ingredients:<br />

17


CHAPTER 1: INTRODUCTION<br />

1. A supermassive black hole (10 6 - 10 10 M ⊙ ) accreting matter at the centre.<br />

2. The accreting material posseses significant angular momentum and therefore<br />

the accretion cannot be radial thus forming a planar disk of optically<br />

thick plasma (accretion disk, see section 3.1) that surrounds the black<br />

hole ([209]). The accreting matter is probably heated by magnetic and/or<br />

viscous processes to temperatures of the order of 10 5 - 10 6 K and emits<br />

thermally in the optical/UV and, probably, soft X-rays. The characteristic<br />

radius of an active nucleus is generally parametrized in terms of<br />

the radius of a non-rotating Schwarzschild black hole, R S = 2GM/c 2 ∼<br />

3 × 10 5 M/M ⊙ cm. For a black hole of mass 10 8 M ⊙ , R S = 10 14 cm or 1<br />

light-hour.<br />

3. A quasi-spherical cloud of hot optically thin plasma surrounds (or perhaps<br />

it is mixed with) the inner parts of the accretion disk. The hot electrons<br />

Compton up-scatter the thermal photons emitted in the accretion disk to<br />

hard energy X-rays. This results in the non-thermal power law spectrum<br />

observed at X-rays in AGN.<br />

4. At distances of about 100 light-days from the central engine are located<br />

clouds of high density (10 7 - 10 10 cm −3 ) gas moving at high velocity (thousands<br />

of km s −1 ), which are responsible for the strong highly ionized<br />

Doppler-broadened optical/UV emission lines observed in the spectrum<br />

of many AGN. This is known as the Broad Line Region (BLR).<br />

5. Further out, at distances of parsecs or a few light-years from the nucleus,<br />

there is a region of thick cold material of gas and dust with a toroidal geometry<br />

in a similar plane as the accretion disk ([127], [181], [182]). The<br />

torus absorbs and scatters the nuclear radiation and partially covers the<br />

central engine, blocking the BLR and the accretion disk to certain lines of<br />

sight. It is not clear whether the torus geometry depends on the luminosity,<br />

but the apparent lack of cross-correlation of the amount of absorption<br />

with the host galaxy inclination angle (even for AGN of the same optical<br />

type), suggests that its orientation must be random with respect to the host<br />

galaxy ([143]).<br />

6. At distances ranging from a few tens of parsecs to kiloparsecs are lowvelocity<br />

clouds of gas excited by the ionizing nuclear radiation emitted<br />

along the polar axis of the torus. This is known as the Narrow Line Region<br />

18


CHAPTER 1: INTRODUCTION<br />

(NLR) and is responsible for the narrow optical/UV emission lines and,<br />

as it lies outside the region affected by the high column density, they only<br />

suffer from moderate dust obscuration. The detection of strong forbidden<br />

narrow lines implies that the density of the material in the NLR cannot<br />

exceed the critical density for collisional de-excitation and must be of the<br />

order of ∼10 4 cm −3 or less.<br />

7. Powerful relativistic radio jets and ionization cones flow from the central<br />

black hole and can extend up to tens of kiloparsecs. They are symmetric<br />

with respect to the black hole and propagate in both directions along the<br />

axis perpendicular to the accretion disk. They are best seen in radio wavelengths<br />

thanks to synchrotron emission, but it is unclear how common are<br />

these jets in AGN.<br />

Unification models try to unify all different classes of AGN with only a few parameters<br />

giving rise to the whole observed variety of properties. The basis of<br />

the unification schemes is that all AGN have intrinsically the same mechanism<br />

(they are powered by accretion onto a supermassive black hole) and their different<br />

observational properties are mostly due to the anisotropic distribution of<br />

obscuring matter. Unification models were firstly invoked to explain the results<br />

of spectropolarimetric observations that detected broad line components in the<br />

optical/UV spectrum of Seyfert 2 galaxies ([6]). Further support to a unified<br />

model theory came with the detection of broad permitted emission lines in the<br />

near-IR unpolarized spectra of optically Type-2 AGN and no broad lines features<br />

in the polarized spectra, as the absorption is less effective in the IR than<br />

in the optical band ([86], [192], [199]).<br />

In this context the different observational properties of Type-1 and Type-2 AGN<br />

are mainly due to different orientation angles from which the obscuring matter<br />

is observed. The absorbing matter would be located at intermediate distances<br />

between the BLR and the NLR, but it would not be completely blocking the<br />

nuclear region. The simplest geometry that describes the obscuring matter is a<br />

torus ([5], [228]), the clouds of the BLR, or material associated with the accretion<br />

disk ([111]). The sources in which we can directly see the nuclear emission and<br />

the BLR will therefore have unabsorbed X-ray spectra and would be optically<br />

classified as Type-1 AGN. The sources in which the obscured matter intercepts<br />

the line of sight of the observer, the central engine and the BLR will be hidden.<br />

These sources would therefore have an absorbed X-ray spectrum and will be<br />

19


CHAPTER 1: INTRODUCTION<br />

Figure 1.4 Schematic diagram of the unified model of AGN (image credit: NASA). The<br />

torus is not at scale.<br />

20


CHAPTER 1: INTRODUCTION<br />

optically classified as Type-2 AGN (see Figure 1.4).<br />

Unification schemes provide a simple description of the most general properties<br />

of AGN. X-ray observations show that Type-2 AGN are systematically weaker<br />

in the soft X-ray band than Type-1 AGN, as expected if their nuclear emission<br />

is obscured. However, the number of observational results that cannot be explained<br />

by the standard unified model are increasing rapidly. X-ray surveys are<br />

finding that not all X-ray obscured AGN have Type-2 optical spectra, and that<br />

not all Type-1 AGN are X-ray unobscured (e.g. [149], [150]).<br />

1.6 Evolution of AGN in X-rays<br />

1.6.1 The K-correction<br />

The luminosities of extragalactic objects such as AGN have to be corrected for<br />

their redshift, so that all luminosities are studied over the same emitted band<br />

(we would have different emitted bands for different redshifts otherwise).<br />

Formally, this can be expressed as:<br />

L = L obs × K corr (z) (1.6.1)<br />

where L obs is the luminosity in the observer bandpass and L is the luminosity<br />

over the same bandpass in the frame of reference of the emitting object. For an<br />

object with a power law spectrum (typical of AGN), the K-correction is:<br />

K corr (z) = (1 + z) α−1 (1.6.2)<br />

where α is the energy index defined in section 1.5.1. The K-correction is unity<br />

(and therefore has no effect at all) when α = 1. In any other cases, neglecting<br />

the K-correction would lead to spurious luminosity evolution with redshift.<br />

In this work, the spectral index of most AGN is known and therefore we have<br />

applied the K-correction for individual AGN. However, in some cases in which<br />

we do not know the spectral index for every single source in a sample we have<br />

applied an average K-correction.<br />

21


CHAPTER 1: INTRODUCTION<br />

1.6.2 Space density of AGN<br />

As we have seen before, AGN can be detected at cosmological distances thanks<br />

to their large luminosity and hence they are of great interest to study the Universe<br />

at earlier epochs. However, all observations have a limiting flux for detection,<br />

such that an object with a flux below this limit cannot be distinguished<br />

from the background. The limiting luminosity for detection varies with distance<br />

so that we end up seeing different populations of AGN at different redshifts.<br />

The luminosity function enables us to represent what is observed in<br />

terms of different distances and luminosities, and is thus defined as the number<br />

of objects detected per unit comoving volume per unit luminosity. In general,<br />

the luminosity function will be a function of both luminosity and redshift. The<br />

comoving volume is the volume that a region of space would have if it were<br />

observed at the present epoch. Hence, the comoving volume is independent<br />

of the expansion of the Universe and therefore there is no density evolution of<br />

objects caused by this cosmological expansion.<br />

Evolution in the comoving space density or luminsity of AGN is seen as a<br />

change in the luminosity function. This can be described in different forms:<br />

• Pure Luminosity Evolution (PLE): In PLE, the number of AGN per unit<br />

comoving volume remains constant while the luminosity of each AGN<br />

evolves, at a rate that is the same for AGN of all luminosities. In the<br />

framework of a PLE model, the luminosity function retains its shape at all<br />

redshifts and depends only on the evolution law and the luminosity function<br />

at zero redshift. It has been a fairly succesful model for the evolution<br />

of AGN at radio ([56], [169], [198]), optical ([27], [26], [106], [107], [108],<br />

[205]) and X-ray ([51], [25], [118], [171]) wavelengths, although deviations<br />

from this model have been seen at high redshifts and overproductions of<br />

the X-ray background have been reported ([157], [224]).<br />

• Pure Density Evolution (PDE): In PDE, the number of AGN per unit comoving<br />

volume is assumed to evolve while the distribution of luminosities remains<br />

constant. This model has failed to fit the current observations well<br />

since it clearly overpredicts the measured cosmic X-ray background (e.g.<br />

[99], [157], [224]) and therefore has been ruled out as a possible description<br />

of the evolution of AGN.<br />

• Luminosity-dependent Density Evolution (LDDE): In LDDE, both the num-<br />

22


CHAPTER 1: INTRODUCTION<br />

ber of AGN per unit comoving volume and the luminosity of each AGN<br />

are assumed to evolve. This model, although more complicated than the<br />

previous ones, it is now widely accepted since it reproduces well the observations<br />

in both optical (e.g. [236]) and X-ray ([157], [104], [224], [130],<br />

[211]) wavelengths.<br />

1.7 Modern instrumentation in X-ray astronomy<br />

The sources that dominate the CXB are faint, with typical fluxes of ∼10 −14 erg<br />

cm −2 s −1 (see sections 1.3 and 3.4) and therefore large area detectors with very<br />

high efficiency are required to study them. The rocket that first discovered the<br />

X-ray background in 1962 carried a small detector, a Geiger counter, with very<br />

low spectral resolution and no imaging capabilities (see section 1.2). The first<br />

X-ray orbiting observatories used mechanical collimators to limit the field of<br />

view covered by the detectors. This provided rough angular resolutions (of<br />

the order of several degrees) which significantly limited the sensitivity of the<br />

observations because of the confusion of the faintest sources.<br />

Since then there have been major advances in instrumentation in X-ray astronomy,<br />

the most important one being the introduction fo X-ray imaging optics<br />

(the Einstein observatory in the 1970s).<br />

Proportional counter detectors have been extensively used in several missions<br />

for many decades, until the advent of high resolution solid state detectors (charge<br />

coupled devices, CCDs), which were first used in the Japanese ASCA observatory<br />

and were able to produce X-ray images with photons of up to 10 keV. The<br />

spectral resolution of CCD detectors is one order of magnitude better to that of<br />

proportional counters and are now used in all modern X-ray observatories.<br />

1.7.1 X-ray telescopes<br />

Focusing X-ray photons is an extremely difficult task. Unlike standard reflecting<br />

telescopes, in which photons are directed almost perpendicular to the surface<br />

of the reflector and then collected in the focal plane, X-ray telescopes cannot<br />

make use of normal incidence as it leads to the absorption and scattering of<br />

X-ray photons due to their high energy.<br />

X-ray photons can only be focused in grazing incidence at very small angles.<br />

23


CHAPTER 1: INTRODUCTION<br />

The reflection angle is ∼1 degree for ∼1 keV X-rays and standard reflection<br />

metals such as Au, Pt, Ir or Ni.<br />

The German physicist Hans Wolter showed in 1952 that a system formed by<br />

two reflectors (a paraboloid followed by a co-axial hyperboloid) can focus X-<br />

rays. This system, known as Wolter-1 configuration, is frequently used in X-ray<br />

astronomy (for instance, in the XMM-Newton X-ray telescopes). The light path<br />

in a Wolter-1 configuration is shown in Figure 1.5.<br />

The critical angle for reflection depends inversely on the energy of the incident<br />

photons (reflection is more difficult at higher energies) and increases with the<br />

density of free electrons in the reflecting material. Shorter focal length telescopes<br />

focus mostly soft X-rays, while to focus harder photons longer focal<br />

lengths (which implies longer telescopes) are needed. The effective area of the<br />

telescope is therefore a function of the energy of the incident photons but also<br />

of the off-axis angle (distance to the centre of the optical system or optical axis):<br />

the fraction os incident photons reaching the focal plane decreases with the offaxis<br />

angle. This effect is known as vignetting and it is observed as a decline of<br />

the effective area of the X-ray telescopes at larger off-axis angles.<br />

In grazing incidence the effective area of the mirror seen by an incoming photon<br />

is very small, and therefore a large number of mirrors are usually nested in<br />

order to increase significantly the collecting area of the telescope.<br />

1.7.2 Imaging X-ray detectors<br />

A variety of non-dispersive detectors have been used in X-ray astronomy, including<br />

Position Sensitive Proportional Counters (PSPC), Micro Channel Plates<br />

(MCP), Micro-calorimeters, or Charge Coupled Devices (CCD). PSPCs, for instance,<br />

were the detectors onboard the Einstein and ROSAT observatories.<br />

CCDs were used in the ASCA observatory and are also onboard of the Chandra<br />

and XMM-Newton observatories. CCDs are imaging arrays of pixels (charge<br />

collecting units or capacitors) with low noise and good energy resolution. When<br />

a X-ray photon hits a capacitor, the semiconductor material of the CCD produces<br />

multiple electron-hole pairs via the photoelectric effect. Optical photons<br />

can only release a few electron-hole pairs and hence need to be accumulated<br />

in order to collect enough signal through longer exposures. However, a X-ray<br />

photon is able to produce hundreds of thousands of pairs which, in principle,<br />

24


CHAPTER 1: INTRODUCTION<br />

Figure 1.5 Light path on a Wolter-1 X-ray telescope (image credit: ESA).<br />

make possible to measure the energy of each photon individually. CCD detectors<br />

work as photon counting devices in X-ray astronomy.<br />

The charge carried by the cloud of electrons is proportional to the energy of the<br />

incident photon. To read the signal, the charge is transferred along the surface<br />

of the semiconductor from the capacitor hit by the X-ray photon to neighbouring<br />

capacitors. The last capacitor in the array will transfer then its charge to an<br />

amplifier that converts the charge into voltage, that can be recorded, making<br />

possible to recover the position and energy of the incident photon.<br />

The charge collected in a CCD is recovered in successive read-outs that limit<br />

the time resolution of the observations to a few seconds. CCDs can be operated<br />

in different observing modes, although to improve the time resolution of the<br />

observatories it is necessary to reduce the read-out time of the CCD which is<br />

only achievable by reducing the read-out area and hence the field of view. The<br />

next generation of CCDs for X-ray astronomy will be based on Active Pixel<br />

Sensors in which each pixel is read individually.<br />

25


CHAPTER 1: INTRODUCTION<br />

Figure 1.6 Payload of the XMM-Newton observatory (image credit: ESA).<br />

26


CHAPTER 1: INTRODUCTION<br />

1.7.3 The XMM-Newton observatory<br />

Most of the data used in this thesis have been obtained with the X-ray detectors<br />

onboard the XMM-Newton observatory 3 , which was the second cornerstone of<br />

the Horizon 2000 science programme of the European Space Agency (ESA). It<br />

was launched on the 10th of December of 1999 on an Ariane 504 rocket from<br />

Kourou, French Guyana. The satellite is in a 48-hour elliptical orbit allowing<br />

continuous monitoring of selected targets up to about 40 hours.<br />

XMM-Newton carries six science instruments. Three co-aligned X-ray telescopes<br />

(each with 58 nested Wolter-1 type mirrors) focus X-rays into the CCD cameras<br />

at the focal plane ([117], see Figure 1.6). The telescopes have the largest effective<br />

area achieved so far, 3 to 10 times higher at 1 and 7 keV that that of Chandra, respectively.<br />

The light path of the X-ray telescopes aboard XMM-Newton is shown<br />

in Figure 1.7.<br />

The maximum effective area of the XMM-Newton mirrors is 0.15 m 2 at 1 keV<br />

and 0.05 m 2 at 5 keV. Each telescope has its own Point Spread Function (PSF,<br />

the distribution of light of a point-like source in the focal plane), which varies<br />

little with energy in the 0.1-4 keV range but has a strong dependence on the<br />

off-axis angle.<br />

On the focal plane of each telescope there is an European Photon Imaging Camera<br />

(EPIC) instrument. XMM-Newton carries two different technologies of CCD<br />

detectors: two of the cameras contain 7 individual MOS (Metal-Oxide Semiconductor)<br />

front-illuminated CCD arrays (600 × 600 pixel each, [222]) referred<br />

to as MOS1 and MOS2 cameras, respectively, while the last detector contains<br />

12 pn (p-n junction) back-illuminated CCDs (64 × 200 pixel each, [216]) and is<br />

referred to as pn camera, which was specially developed for XMM-Newton.<br />

The EPIC cameras can perform imaging over a 30 arcmin diameter field of view<br />

(2.5 times that of Chandra), and are sensitive to photons with energies from 0.1<br />

to 15 keV providing moderate energy resolution (E/∆E ∼ 20 − 50). Because<br />

of the small sizes of the pixels in the EPIC cameras (1.1 and 4.1 arcsec in the<br />

MOS and pn, respectively), the angular resolution is determined by the PSF of<br />

their mirror modules. The FWHM of the PSF is 6 arcsec, allowing to obtain<br />

X-ray source positions with accuracies of ∼1-3 arcsec, while Chandra achieves a<br />

positional accuracy of 0.3-1 arcsec. X-ray source confusion in XMM-Newton is<br />

3 ¢¡£¡¥¤§¦©¨¨1¥¢2¥354¡ 27


CHAPTER 1: INTRODUCTION<br />

Figure 1.7 Lightpath of the XMM-Newton telescopes (image credit: ESA). Left panel:<br />

Without grating assembly. Right panel: With grating assembly.<br />

not significant for observations with exposures ≤100-200 ks.<br />

The EPIC cameras can be operated in various observing modes ([122], [123],<br />

[128]). The EPIC CCDs are sensitive to X-rays but also to IR, visible and UV<br />

photons. In order to reduce the contamination of the observations each EPIC<br />

camera is equipped with three aluminized optical blocking systems with different<br />

thicknesses (named thick, medium and thin filters). It is also possible to<br />

perform observations in an open position (with no filter at all) or in a closed<br />

position (only for internal calibration purposes).<br />

The telescopes with MOS cameras in their focal planes have a grating assembly<br />

(RGA, [54]) that diverts part (∼40%) of the incoming photons to a Reflection<br />

Grating Spectrometer (RGS). About 44% of the incoming light reaches the MOS<br />

cameras, while the remaining ∼16% is absorbed by the supporting structures<br />

of the RGA. The RGS instruments perform high resolution dispersive spectroscopy<br />

(resolving power E/∆E ∼ 200 − 800) of bright sources in the soft<br />

energy band.<br />

1.8 Aims of the thesis<br />

In the previous sections we have briefly summarized our current understanding<br />

of the nature of the cosmic X-ray background and AGN emission. We have<br />

tried to highlight the importance of studying the X-ray emission of AGN since<br />

it is originated in the innermost regions of these objects, close the central supermassive<br />

black hole. We know that AGN dominate the overall CXB emission<br />

and that they are detectable at very large distances. Hence, X-ray surveys of<br />

28


CHAPTER 1: INTRODUCTION<br />

AGN are essential to investigate the large-scale structure of the Universe and<br />

to provide an accurate look at the evolution of the central engine that powers<br />

these sources along cosmic time.<br />

The aim of the work presented in this thesis is to give insight to the evolutionary<br />

and clustering properties of a medium-deep sample of AGN, and to evaluate<br />

the importance of the contribution of this population of AGN to the cosmic<br />

X-ray background.<br />

The medium-flux X-ray surveys presented in this work in Chapter 2 are especially<br />

suitable to estimate the intensity of the CXB. Since its origin seems to<br />

be the integrated emission of discrete individual sources we can estimate the<br />

resolved fraction of the CXB integrating the sky density of the sources in our<br />

samples over their flux range. However, the contribution from sources brighter<br />

and fainter than ours, although small, is not negligible. We could obtain a<br />

much more accurate estimation spanning a broader flux range, but to do that<br />

we would need to know the sky density distribution of sources at those fluxes<br />

in detail. The sky density or source counts are dependent on the cosmological<br />

properties of the sources involved and can be described by an empirical model<br />

(which is a function of the X-ray flux). Adding wide-area bright surveys and<br />

deep faint surveys to our medium sample would allow us to constrain the parameters<br />

of such model with unprecedent accuracy and therefore determine the<br />

contribution of these sources to the total intensity of the CXB. Such analysis is<br />

presented in Chapter 3.<br />

Similarly, this kind of surveys (medium-deep moderately wide area) are potentially<br />

useful to explore the underlying cosmic structure of the X-ray Universe<br />

since they achieve both high angular density and width. If there exists a cosmological<br />

structure in the X-ray sky, we would be able to detect it either through<br />

the cosmic variance in the number of sources per field or the distribution of<br />

the angular separations between pairs of sources. This can be estimated using<br />

a variety of mathematical tools that rely on Poissonian statistics and the correlation<br />

functions introduced in section 1.4. This subject will be investigated in<br />

Chapter 4.<br />

As we have seen in the previous sections, the origin of the X-ray emission of<br />

AGN is located in the proximity of the accreting central black hole. Changes in<br />

the accretion rate of matter onto the black hole would therefore lead to changes<br />

in the X-ray luminosity of the object. Since AGN can be detected through enor-<br />

29


CHAPTER 1: INTRODUCTION<br />

mous distances, we can measure their intrinsic luminosity at different epochs of<br />

the Universe and hence estimate how it is related to the accretion rate of matter<br />

at those epochs. It is believed that the evolution observed in the luminosity of<br />

AGN is directly linked to intrinsic changes in the matter inflow, which tell us<br />

how the supermassive black hole grew and fed in the early stages of the Universe.<br />

However, according to the unified model of AGN we have described<br />

before, Type-2 AGN are more likely to be observed at high energies (> 2 keV)<br />

since only hard X-rays are capable of passing through the molecular torus that<br />

enshrouds the central engine of the AGN. Hence, this absorption has to be taken<br />

into account in order to obtain unbiased intrinsic luminosities. In Chapter 5 we<br />

will address this subject as well as the possible evolutionary models for AGN.<br />

Specifically, the issues we have investigated in this thesis are:<br />

1. To study the flux distribution of sources detected in soft and hard X-rays at<br />

medium fluxes. This distribution depends on the cosmological properties<br />

of the sources and can be described by an empirical model. We have used<br />

this model to estimate the contribution of the sources to the total CXB<br />

intensity. At what fluxes do the dominant sources of the CXB contribute<br />

most?<br />

2. To explore the underlying cosmic structure of the X-ray Universe. It is<br />

known that local galaxies (redshifts up to ∼0.2) cluster forming large-scale<br />

structures (filaments and voids). AGN are at cosmological distances and<br />

hence they will allow us to investigate whether these structures were already<br />

present in the early Universe and to quantify the strength of the<br />

clustering.<br />

3. To study the behaviour of the fraction of absorbed AGN at different X-<br />

ray luminosities and redshifts and to model the intrinsic absorption of the<br />

AGN as a function of these parameters. Accordingly to the unified model,<br />

it is more probable to detect Type-2 AGN in hard X-rays since only this<br />

radiation can pass through the molecular torus around the central black<br />

hole. Does the fraction of obscured AGN depend on the luminosity? Were<br />

there more absorbed AGN in the early Universe?<br />

4. To investigate the evolution of AGN at different epochs of the Universe.<br />

Were they more luminous in the past? Has the density of AGN changed<br />

with time? There is evidence of this, so we intend to quantify better these<br />

30


CHAPTER 1: INTRODUCTION<br />

subjects making use of the available absorption information. We can convolve<br />

an absorption model with evolutionary models in order to obtain<br />

intrinsic results for the sources. Has the accretion rate of matter changed<br />

with time? At what epoch of the Universe were AGN fully formed?<br />

We have addressed these points throughout this thesis and the results are discussed<br />

and summarized in Chapter 6.<br />

31


Chapter 2<br />

AGN surveys with XMM-Newton<br />

As discussed in Chapter 1, X-ray source populations studies are essential to understand<br />

their X-ray emission and cosmic evolution, especially those related to<br />

AGN. However, in order to have good enough statistics, these studies need to<br />

collect a moderately large number of objects. Moreover, it is difficult (if not impossible)<br />

for a single survey to gather simultaneously a wide sky coverage and<br />

high sensitivity to detect very faint sources. It is hence imperative to combine<br />

wide field shallow surveys with deep pencil-beam surveys so as to carry out<br />

accurate population and evolutionary studies such as those conducted in this<br />

thesis.<br />

In this chapter we will discuss the X-ray extragalactic surveys that have been<br />

used throughout this thesis. The surveys that compose the core of this work<br />

are the AXIS survey ([13], [38], see section 2.1) and the XMS survey ([12], see<br />

section 2.2). The project involving the construction of both surveys and the<br />

development of the tools needed for that purpose has been led by the X-ray<br />

Group at the Instituto de Física de Cantabria 1 . The author of this thesis has<br />

been actively involved in the development of these surveys by participating in<br />

observation runs and reducing optical and X-ray data.<br />

AXIS and XMS are medium-depth samples, which means that the sources inside<br />

of them have the bulk of their X-ray emission at medium fluxes. The dominant<br />

population at these fluxes are AGN, whose clustering and evolutionary<br />

properties are going to be investigated in this thesis as they are critical issues to<br />

understand the accretion history of Universe across cosmic time.<br />

Important as they are, the AXIS and XMS samples alone cannot fulfill the cov-<br />

1 ¢¡£¡¥¤§¦©¨¨64 +7389¥! 4+34§¢¥¨ ∼ £:¢¨ 33


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

erage and depth requirements described above to conduct this work. We have<br />

combined them with a variety of surveys (some of them deeper, other shallower)<br />

in all energy bands under study. The properties of these samples are described<br />

in section 2.3 and they were chosen so they were complementary to<br />

AXIS and XMS in order to obtain results of unprecedent accuracy in the population<br />

(see Chapter 3) and evolutionary (see Chapter 5) studies carried out in<br />

this thesis.<br />

2.1 The AXIS survey<br />

2.1.1 Introduction and field selection<br />

Thanks to its large collecting area and sensitivity, in each XMM-Newton pointing<br />

∼30-150 serendipitous sources are detected, most of them being entirely<br />

new detections. In 1996, three years before the launch of XMM-Newton, the<br />

XMM-Newton Survey Science Centre (SSC) 2 was appointed by ESA to develop<br />

the software analysis system, the pipelines to process all the XMM-Newton data<br />

and the scientific exploitation of the XMM-Newton’s serendipitous sky surveys.<br />

The SSC is a consortium of ten european institutions which has, among other responsabilities,<br />

the task of following-up and identifying the XMM-Newton serendipitous<br />

sources (XID programme, [234]). The XID programme has two components:<br />

• Core programme: Its aim is to obtain the identifications from a well-defined<br />

sample of X-ray sources. This strategy involves a high galactic latitude and<br />

a low galactic latitud samples. The high latitude sample consists of three<br />

large enough subsamples of sources with fluxes ≥ 10 −13 erg cm −2 s −1<br />

(bright sample), ≥ 10 −14 erg cm −2 s −1 (medium sample), and ≥ 10 −15 erg<br />

cm −2 s −1 (faint sample).<br />

• Imaging programme: Deep optical and infrared imaging of a larger fraction<br />

of objects in order to obtain optical magnitudes and colors necessary for<br />

statistical identifications of the sources.<br />

AXIS (An XMM-Newton International Survey, [13], [38]) is a medium X-ray survey<br />

that constitutes the backbone of the SSC XID programme by providing es-<br />

2 ¢¡£¡¥¤§¦©¨¨£¢¡£©£! " 34


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

sential resources for the exploitation of the XMM-Newton serendipitous survey.<br />

In total, 36 XMM-Newton observations were selected for optical follow-up of<br />

X-ray sources within the AXIS programme. The selected fields (listed in Table<br />

2.1) had high galactic latitude (|b| > 20 ◦ ) to avoid contamination coming<br />

from Galactic sources. They had total exposure times above 15 ks and bright<br />

or extended targets were excluded, except for the A1837 and A399 fields. Some<br />

of the fields ended up with exposure times below 15 ks after discarding observation<br />

intervals with high background rates, and others were added to expand<br />

the solid angle for bright X-ray sources. All of them have been used for the<br />

study of the source counts (see Chapter 3) and for cosmic variance and angular<br />

clustering analysis (two fields were removed in this latter case: HD111812 and<br />

HD117555, see section 4.3 for details).<br />

Of the original 36 fields, 27 were selected for optical follow-up of X-ray sources<br />

at medium fluxes. Two of them (A2690 and MS2137) were excluded later from<br />

the identification programme because of their low Declination, which prevented<br />

them from being observed from the Calar Alto Observatory (Spain) 3 . The remaining<br />

25 fields constitute the XMS sample ([12], see section 2.2).<br />

2.1.2 Data processing and source selection<br />

The Observation Data Files (ODF) of the data used in this thesis were processed<br />

in the SSC Pipeline Processing Facility at the University of Leicester (UK), using<br />

the same Science Analysis System (SAS, [71]) used for the processing of the<br />

1XMM catalogue 4 , except for the field Mkn205 which was processed with the<br />

slightly different SAS version 5.3.3. The field PHL1092 was never reprocessed<br />

originally and a set of data from the XMM-Newton Science Archive reprocessed<br />

with SAS version 5.4.0 was used for this work. The difference between the SAS<br />

versions used in the field Mkn205 and the rest of the fields was the possibility<br />

of including sources with negative count rates while using the source detection<br />

task (;7=7?§;2@1;BAC@<br />

) in the latter version. This was allowed to avoid a bias in the<br />

total count rates of sources that were undetected in one or more energy bands.<br />

However, since negative count rates do not have physical meaning and can<br />

¢¡£¡¥¤§¦©¨¨0 9<br />

3<br />

4 The First XMM-Newton Serendipitous Source Catalogue (1XMM) was released in 2003 and contains<br />

source detections from 585 XMM Newton EPIC observations made between March 1 st 2000 and May<br />

5 th 2002. It comprises over 28000 individual X-ray sources with a median 0.5-12 keV flux of ∼ 3 × 10 −14 erg<br />

cm −2 s −1 .<br />

35


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

cause numerical problems in certain cases, negative count rates for this work<br />

have all been set to zero as well as the detection likelihood of the source.<br />

The SAS products provide X-ray source lists (created ;D


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

(sizes and positions are also listed in Table 2.1) as well as rectangular regions<br />

around Out Of Time (OOT) regions 5 .<br />

Two sets of fields were partially overlapping: G133-69pos_2 and G133-69pos_1<br />

on one hand and SDS-1b, SDS-2 and SDS-3 on the other. The portion of the<br />

second field (and also the third in the case of the latter set) overlapping with the<br />

first field was masked so as to not having in the final lists duplicated sources<br />

and an overestimation of the total sky coverage of the survey.<br />

In the analysis carried out in this thesis we have used the following energy<br />

bands:<br />

• Soft: 0.5-2 keV<br />

• Hard: 2-10 keV<br />

• XID: 0.5-4.5 keV<br />

• Ultrahard: 4.5-7.5 keV<br />

The Soft, XID and Ultrahard bands are identical to the standard SAS (before<br />

version 6.9) bands 2, 9 and 4, respectively 6 . The hard band is a combination of<br />

the standard SAS (before version 6.9) bands 3, 4 and 5, which corresponds to<br />

counts in the range 0.5-12 keV. However, all hard band fluxes in this work have<br />

been calculated in the range 2-10 keV.<br />

The detection likelihood in the hard band (L 345 ) was obtained from the detection<br />

likelihoods in bands 3, 4 and 5 (L 3 , L 4 and L 5 , respectively) using the following<br />

expression:<br />

L 345 = −log(1 − Q( 5 2 , L′ 3 + L′ 4 + L′ 5 )) (2.1.1)<br />

where L<br />

i ′ is obtained from: L i = −log(1 − Q( 3 2 , L′ i )) (2.1.2)<br />

and Q(a, x) is the incomplete gamma function (see ;7=7?§;2@1;BAC@<br />

documentation).<br />

5 Bright band that extends from the target to the pn chip reading edge, due to the photons arriving at<br />

the detector while it was being read out.<br />

6 These bands correspond to SAS version 6.9 or earlier. New energy bands were defined during the<br />

2XMM catalogue processing: former band 2 was splitted into two bands (0.5-1 and 1-2 keV), while former<br />

bands 4 and 5 were merged into a single band (4.5-12 keV).<br />

37


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

detection likeli-<br />

In total, there were 2560 accepted sources with ;D=7?§;@B;BAC@<br />

an<br />

hood ≥10 (the default value) in at least one band.<br />

The final selection of sources in each band was done using their detection likelihood<br />

and the corresponding sensitivity map (see Appendix A for details) at<br />

their sky position. We have chosen a detection confidence limit of 5-σ which<br />

corresponds to L = 15 in the band under consideration. In addition, we have<br />

also imposed that the source has a count rate equal or larger than the value of<br />

the sensitivity map at that sky position in order to ensure that the detection is<br />

reliable. This way, less than 5% of the sources in the soft band, ∼10% of the<br />

sources in the XID and ultrahard bands, and ∼20% of the sources in the hard<br />

band with L ≥ 15 were excluded.<br />

The total of distinct sources that fulfill all the criteria mentioned above in at least<br />

one band is 1433. The number of sources in each band are listed in Table 2.2.<br />

For the analysis done in this thesis we have used only sources detected in the<br />

corresponfing bands, unless otherwise stated. It has to be noted, however, that<br />

once a source has been detected in at least one band it can be considered a real<br />

source and therefore its counts in other bands can be used, even if the source<br />

has a count rate smaller than the detection threshold in those bands.<br />

In order to estimate the number of spurious detections in the survey, it is firstly<br />

needed to calculate the number of independent source detection boxes. The<br />

;D=2?§;2@B;BAD@ task uses an input list ;2KHG7L?§;@B;BAC@ from which is a sliding-box algorithm<br />

using a square 5 × 5 pixels detection cell. Each pixel side has 4 arcsec, so<br />

the individual detection cell has an area of 5 × 5 × (4”) 2 =400 arcsec 2 . Since the<br />

total area of the survey is ∼4.8 deg 2 (see section 3.2.1), or about 155520 independent<br />

detection cells, and that the probability of a false detection at the 5-σ<br />

level is 0.000057, the number of spurious detections is 9 in each detection band.<br />

This is less than 1% in the soft and XID bands, about 2% in the hard band, and<br />

almost 10% in the ultrahard band.<br />

38


39<br />

Table 2.1: Fields of the AXIS survey.<br />

Target name OBS_ID R.A. Dec Texp Filter N H,Gal R.A. Dec. R OOT R.A. Dec. R’<br />

(J2000) (J2000) (s) (10 20 cm −2 ) (J2000) (J2000) (”) (”) (J2000) (J2000) (”)<br />

A2690 0125310101 00:00:30.30 −25:07:30.00 21586.7 Medium 1.84 00:00:21.2 −25:08:12.18 20 0 - - 0<br />

Cl0016+1609 a 0111000101 00:18:33.00 +16:26:18.00 29149.3 Medium 4.07 00:18:33.2 +16:26:07.97 148 0 - - 0<br />

G133-69pos_2 a 0112650501 01:04:00.00 −06:42:00.00 18080.0 Thin 5.19 - - 0 0 - - 0<br />

G133-69Pos_1 a 0112650401 01:04:24.00 −06:24:00.00 20000.0 Thin 5.20 - - 0 0 - - 0<br />

PHL1092 0110890501 01:39:56.00 +06:19:21.00 16180.0 Medium 4.12 01:39:55.8 +06:19:19.67 88 40 01:40:09.0 +06:23:27.67 68<br />

SDS-1b a,d 0112371001 02:18:00.00 −05:00:00.00 43040.0 Thin 2.47 - - 0 0 - - 0<br />

SDS-3 a,d 0112371501 02:18:48.00 −04:39:00.00 14927.9 Thin 2.54 - - 0 0 - - 0<br />

SDS-2 a,d 0112370301 02:19:36.00 −05:00:00.00 40673.0 Thin 2.54 - - 0 0 - - 0<br />

A399 a,c 0112260201 02:58:25.00 +13:18:00.00 14298.7 Thin 11.10 02:57:50.2 +13:03:20.88 272 0 - - 0<br />

Mkn3 a,c 0111220201 06:15:36.30 +71:02:04.90 44506.1 Medium 8.82 06:15:36.6 +71:02:15.95 76 32 - - 0<br />

MS0737.9+7441 a,e 0123100201 07:44:04.50 +74:33:49.50 20209.3 Thin 3.51 07:44:04.3 +74:33:54.56 120 40 - - 0<br />

S5 0836+71 a 0112620101 08:41:24.00 +70:53:40.70 25057.3 Medium 2.98 08:41:24.3 +70:53:41.06 160 52 - - 0<br />

PG0844+349 0103660201 08:47:42.30 +34:45:04.90 9783.5 Medium 3.28 08:47:42.9 +34:45:03.27 200 40 - - 0<br />

Cl0939+4713 a 0106460101 09:43:00.10 +46:59:29.90 43690.0 Thin 1.24 09:43:01.8 +46:59:44.37 160 0 - - 0<br />

B21028+31 a 0102040301 10:30:59.10 +31:02:56.00 23236.0 Thin 1.94 10:30:59.3 +31:02:56.08 140 72 - - 0<br />

B21128+31 a 0102040201 11:31:09.40 +31:14:07.00 13799.8 Thin 2.00 11:31:09.6 +31:14:06.02 140 44 - - 0<br />

Mkn205 a,e 0124110101 12:21:44.00 +75:18:37.00 17199.6 Medium 3.02 12:21:43.8 +75:18:39.08 140 36 - - 0<br />

MS1229.2+6430 a,e 0124900101 12:31:32.32 +64:14:21.00 28700.0 Thin 1.98 12:31:31.2 +64:14:18.06 140 40 - - 0<br />

HD111812 b 0008220201 12:50:42.56 +27:26:07.70 37338.8 Thick 0.90 12:51:42.6 +27:32:23.27 168 40 - - 0<br />

NGC4968 0002940101 13:07:06.10 −23:40:43.00 4898.7 Medium 9.14 13:07:06.3 −23:40:33.23 40 0 - - 0<br />

NGC5044 0037950101 13:15:24.10 −16:23:06.00 20030.0 Medium 5.03 13:15:24.2 −16:23:08.53 340 0 - - 0<br />

IC883 0093640401 13:20:35.51 +34:08:20.50 15849.4 Medium 0.99 13:20:35.4 +34:08:21.37 48 0 13:20:54.4 +33:55:17.26 104<br />

HD117555 a,b 0100240201 13:30:47.10 +24:13:58.00 33225.4 Medium 1.16 13:30:47.8 +24:13:51.07 160 40 - - 0<br />

F278 0061940101 13:31:52.37 +11:16:48.70 4648.3 Thin 1.93 13:31:52.4 +11:16:43.88 48 0 - - 0<br />

A1837 a 0109910101 14:01:34.68 −11:07:37.20 45361.3 Thin 4.38 14:01:36.5 −11:07:43.14 440 0 - - 0<br />

UZLib a 0100240801 15:32:23.00 −08:32:05.00 23391.2 Medium 8.97 15:32:23.4 −08:32:05.32 140 40 - - 0<br />

FieldVI 0067340601 16:07:13.50 +08:04:42.00 9634.0 Medium 4.00 - - 0 0 - - 0<br />

PKS2126-15 a 0103060101 21:29:12.20 −15:38:41.00 16150.0 Medium 5.00 21:29:12.1 −15:38:40.44 120 40 - - 0<br />

PKS2135-14 a, f 0092850201 21:37:45.45 −14:32:55.40 28484.3 Medium 4.70 21:37:45.1 −14:32:55.22 120 44 - - 0<br />

CHAPTER 2: AGN SURVEYS WITH XMM-Newton


R<br />

R<br />

OQP<br />

ONM<br />

40<br />

Table 2.1: Continued.<br />

Target name OBS_ID R.A. Dec Texp Filter N H,Gal R.A. Dec. R OOT R.A. Dec. R’<br />

(J2000) (J2000) (s) (10 20 cm −2 ) (J2000) (J2000) (”) (”) (J2000) (J2000) (”)<br />

MS2137.3-2353 0008830101 21:40:15.00 −23:39:41.00 9880.0 Thin 3.50 21:40:15.1 −23:39:39.32 140 48 - - 0<br />

PB5062 a 0012440301 22:05:09.90 −01:55:18.10 28340.9 Thin 6.17 22:05:10.3 −01:55:20.38 140 40 - - 0<br />

LBQS2212-1759 a 0106660101 22:15:31.67 −17:44:05.00 90892.5 Thin 2.39 - - 0 0 - - 0<br />

PHL5200 a 0100440101 22:28:30.40 −05:18:55.00 43278.5 Thick 5.26 22:28:30.4 −05:18:53.12 16 0 - - 0<br />

IRAS22491-1808 a 0081340901 22:51:49.49 −17:52:23.20 19867.2 Medium 2.71 22:51:49.4 −17:52:25.02 32 0 - - 0<br />

EQPeg a 0112880301 23:31:50.00 +19:56:17.00 12200.0 Thick 4.25 23:31:52.7 +19:56:18.46 160 48 - - 0<br />

HD223460 0100241001 23:49:41.00 +36:25:33.00 6699.0 Thick 8.25 23:49:40.8 +36:25:32.56 172 40 23:50:02.0 +36:25:36.36 116<br />

Columns: (1) Target name, (2) XMM-Newton OBS_ID number, (3)(4) Right Ascension and Declination of the field centre, (5)(6) Clean exposure time and filter used in the pn camera, (7) Galactic<br />

column density in the direction of the field ([55]),(8)(9)(10) Centre and radius of the circular exclusion area for the target, where applicable, (11) Width of the OOT exclusion area, (12)(13)(14)<br />

Centre and radius of an additional exclusion area around bright/extended sources.<br />

a Fields belonging to the XMS<br />

b Fields excluded from the angular correlation studies<br />

c<br />

astrometric correction failed in these fields<br />

d In common with SXDS ([227])<br />

e In common with HELLAS2XMM ([10])<br />

CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

f Same area as one ChaMP field ([124])


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

2.1.3 X-ray properties of the sources<br />

We have fitted the count rates of the sources in several bands to those expected<br />

from a power-law spectrum in order to study their spectral characteristics. The<br />

model has a slope Γ and Galactic Hydrogen column density N H obtained from<br />

21 cm radio measurements ([55], see Table 2.1). Obviously not all the sources<br />

have their spectrum well represented by a power law (for instance, stars and<br />

galaxy clusters show thermal spectra) but for the purpose of calculating fluxes<br />

in the same bands as those in which the fit is performed this is an accurate and<br />

simple model.<br />

We have calculated the expected count rates for different values of Γ ranging<br />

from -10 to 10 in steps of 0.5, linearly interpolating for intermediate values,<br />

using LSIF1;1A<br />

([7]) with the on-axis redistribution matrix files and effective areas<br />

for each field generated by the SAS TJUV§;7W task . The count rates obtained from<br />

are already corrected from vignetting, so the effective areas were<br />

generated disabling the vignetting and PSF corrections in TJ0UV§;7W .<br />

;D=2?§;2@B;BAD@<br />

The fluxes obtained for bands 2 to 5 (in the case of band 5 the flux was calculated<br />

in the range 7.5-10 keV) were also calculated for the same Γ values but<br />

setting N H to zero, i.e. they were calculated ’before the atmosphere’. We have<br />

also checked that the chosen step size in Γ does not significatively affect the<br />

calculations by repeating the spectral fits in a single field with a step of 0.001,<br />

finding that the differences in the slopes were smaller than the uncertainties.<br />

Spectral fits were performed in bands 2 and 3 for the soft and XID fluxes, and<br />

in bands 3, 4 and 5 for the hard and ultrahard fluxes. The average count rates<br />

of our sources in the XID and hard bands were 0.0095 and 0.0032, respectively,<br />

which give (for a typical exposure time of 15 ks) an average of more than 10<br />

counts per bin thus ensuring the validity of Gaussian statistics. The best fit<br />

Γ and flux were calculated by minimising the χ 2 between the observed and<br />

expected count rates. The minimisation was done in Γ, setting the flux from the<br />

normalisation that minimised the χ 2 in the corresponding band. 1-σ error bars<br />

for Γ and flux were obtained from the values which produced ∆χ 2 = 1 from the<br />

minimum. In most cases these error bars were assymetric but we have used a<br />

symetric error bar obtained from the average of the upper and lower error bars<br />

for the weighted Γ and the source counts study (see section 3). In some cases the<br />

fitted slopes are very steep (|Γ| ∼ 10): this corresponds to sources with positive<br />

count rates in only one of the fitted bands, which forces the power law to the<br />

41


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

steepest allowed slope.<br />

In Table 2.2 are shown the weighted average Γ of the detected sources in each of<br />

the fitted bands as well as the number of sources used in the averages excluding<br />

those sources with |Γ| > 9 (typically


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

Table 2.2 Number of sources selected in different bands N, weighted average slopes<br />

〈Γ〉 and errors (taking into account both the error bars in the individual Γ and the<br />

dispersion around the mean), and number of observed sources used in the average<br />

N ave . The soft, hard, XID and ultra-hard bands are as defined in the text. “Soft and<br />

hard” refers to sources selected simultaneously in the soft and hard bands, “Only soft”<br />

refers to sources selected in the soft band but not in the hard band, and “Only hard”<br />

refers to sources selected in the hard band and not in the soft band.<br />

0.5-2 keV 2-10 keV<br />

Selection N N ave 〈Γ〉 N ave 〈Γ〉<br />

Soft 1267 1239 1.811 ± 0.015 1145 1.53 ± 0.02<br />

Hard 397 397 1.76 ± 0.03 394 1.55 ± 0.03<br />

XID 1359 1335 1.773 ± 0.016 1244 1.47 ± 0.02<br />

Ultrahard 91 91 1.80 ± 0.06 91 1.50 ± 0.07<br />

Soft and hard 345 345 1.79 ± 0.03 342 1.67 ± 0.03<br />

Only soft 922 894 1.878 ± 0.017 803 1.04 ± 0.04<br />

Only hard 52 52 −0.29 ± 0.08 52 0.53 ± 0.12<br />

The identification completeness of the sample in the different energy bands is<br />

listed in Table 2.5 and it largely exceeds 90% in the soft and XID bands, and is<br />

around 85% in the hard and ultrahard bands.<br />

2.2.2 Imaging and selection of optical counterparts<br />

Target fields were mainly observed with the Wide Field Camera (WFC) on the<br />

2.5 m Isaac Newton Telescope (INT) since this instrument practically covers all<br />

the field of view of EPIC if it is optimally centered. These observations were<br />

awarded via the AXIS programme and other programmes devoted to image<br />

the XMM-Newton target fields in the optical. All the XMS target fields were<br />

imaged using the Sloan Digital Sky Survey (SDSS) 7 filters g ′ , r ′ and i ′ .<br />

Exposure times were adjusted to be deep enough in order to have an optical<br />

counterpart for most of the X-ray sources in the r ′ and i ′ filters. The chosen<br />

times were 600 seconds for the g ′ and r ′ filters, and 1200 seconds for the i ′ filter,<br />

which produced an image with a limiting magnitude down to r ′ ∼ 23 − 24 for<br />

7 ¢¡£¡¥¤§¦©¨¨0X92©%& 43


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

Table 2.3 Definition of the XMS samples: the first column is the sample energy band,<br />

the second is the flux limit, the third is the number of sources in the sample, the next<br />

four columns give the number of sources belonging to any pair of two samples, and<br />

the last column is the number of sources only present in the corresponding sample.<br />

Overlap<br />

Sample Flux limit N Soft Hard XID UH Only<br />

(10 −14 cgs)<br />

Soft 1.5 210 210 111 209 50 1<br />

Hard 3.3 159 111 159 128 62 20<br />

XID 2.0 284 209 128 284 57 56<br />

Ultrahard 0.7 a 70 50 62 57 70 2<br />

a This band is not artificially flux limited and this value corresponds<br />

to the flux of the faintest detected source.<br />

a seeing of ∼1-1.5 ′′ ([13]). Under bright moon observing conditions, the times<br />

were doubled and only the reddest filters were used. The WFC images were<br />

reduced using the standard reduction pipeline (see the Cambridge Astronomy<br />

Survey Unit -CASU- site 8 ).<br />

The photometric calibration of the WFC images was also done in the standard<br />

way, observing photometric standard stars at different airmasses during the<br />

same nights in which the XMM-Newton target fields were imaged, and fitting an<br />

extinction curve for each optical band afterwards. In a number of target fields<br />

and for some of the bands the scattering in the extinction curve showed that<br />

the observations were carried out under non-photometric conditions. In those<br />

cases, the calibration was done on one band, typically r ′ , and then colour corrections<br />

were applied to propagate the improved photometry to other bands. To<br />

do that, two sets of complementary photometrically calibrated data were used:<br />

the SDSS-Data Release 5, which was used on 6 XMM-Newton fields, and the<br />

Carlsberg Meridian Catalogue (CMC) 9 astrometric survey in the r ′ band, which<br />

was applied to 19 fields. Details on the calibration procedure can be found in<br />

[12].<br />

The astrometric calibration of the WFC was performed using the CASU pro-<br />

8 8¢5 ¢<br />

∼ "#¨ ¢¡£¡¥¤§¦©¨¨0¡¥1¢2Y<br />

9 ¢¡£¡¥¤§¦©¨¨¥£95 +36¢¡¥5B¢2! £"+¨¥5¨ 44


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

cedures. The astrometric reference for the optical images was the APM Catalogue<br />

10 . The astrometry of the XMM-Newton source positions was registered<br />

against the USNO-A2 source catalogue (same as AXIS, see section 2.1). Since<br />

the optical astrometry refers to a different astrometric system, there could be<br />

artificial mismatches. A test that was done measuring the USNO-APM shifts<br />

in each XMM-Newton field by cross-correlating both source catalogues in the<br />

corresponding region, showed that the shifts were systematic but less than the<br />

statistical accuracy in the X-ray source positions (∼0.5 arcsec against 0.6 arcsec<br />

averaged over the entire XMS sample, respectively).<br />

2.2.3 Identification of XMS sources<br />

The search for candidate counterpartes of the X-ray sources was done using the<br />

r ′ WFC images, for which optical source lists were generated using the CASU<br />

standard procedures. The optical counterparts were searched in these lists and<br />

the candidates must be within 5-σ (90% confidence level) of the X-ray source position<br />

or within 5 arcsec from that position (this last criterion was added to accommodate<br />

any residual systematics in the astrometric calibration of the EPIC<br />

X-ray images).<br />

As a result of this, most sources in the XMS ended up having a single counterpart<br />

with only a few exceptions. In the few cases in which the X-ray source<br />

position fell in a gap between CCDs in the WFC images, the counterpart was<br />

found considering other optical imaging data, mainly the USNO-A2 catalogue<br />

or complementary imaging data. Also, for a number of sources (∼80), there<br />

was more than a single candidate counterpart fulfilling the proximity criteria.<br />

In most of these cases (∼90%), the closest candidate to the X-ray source position<br />

was also the brightest and therefore that was the adopted counterpart.<br />

Given the brightness (r ′


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

Table 2.4 Spectroscopic facilities used to identify XMS sources. The first column lists the<br />

ground-based facility; the second, the instrument used, and the last four columns the<br />

spectral range, slit width, spectral resolution and type of the instrument, respectively.<br />

Telescope Instrument Spectral Slit width Spectral Comments<br />

range (Å) (arcsec) resolution a (Å)<br />

WHT/ORM WYFFOS 3900-7100 2.7 b 7 Fibre<br />

WHT/ORM WYFFOS 3900-7100 1.6 b 6 Fibre<br />

WHT/ORM ISIS 3500-8500 1.2-2.0 3.0-3.3 Long slit<br />

TNG/ORM DOLORES 3500-8000 1.0-1.5 14-15 Long slit<br />

NOT/ORM ALFOSC 4000-9000 1.0-1.5 4 Long slit<br />

3.5m/CAHA MOSCA 3300-10000 1.0-1.7 24 Long slit<br />

UT1/ESO FORS2 4400-10000 1.0 6-12 Long slit<br />

a Measured from unsaturated arc lines; b Width of individual fibres<br />

X-ray surveys such as the SXDS ([227]) or the HBS ([50]), which means that<br />

the vast majority of the XMS sources required optical spectroscopy in order to<br />

identify them.<br />

Optical spectroscopy was conducted in a number of ground-based optical facilities<br />

(see Table 2.4). Due to the low surface density of the XMS sources<br />

(∼100 deg −2 ), the use of multiobject slit-mask spectrographs was not very efficient.<br />

In some cases, a fibre spectrometer (WYFFOS) was used to carry out<br />

the identifications, but only for sources brighter than r ′ ∼20.5 mag due to the<br />

problems to substract the sky light that enters the fiber along with the light<br />

coming from the target and, therefore, the strategy that best fitted the optical<br />

magnitude distribution of the sample was long-slit spectroscopy. A number of<br />

these instruments were used in a variety of ground-based telescopes with apertures<br />

ranging from 2.5 m to 8.2 m (see Table 2.4). The spectra thus obtained are<br />

reliable only for identification purposes based on broad/narrow emission lines<br />

and absorption bands, but not for calculating line fluxes or line ratios. The identification<br />

completenesses achieved in each band are listed in Table 2.5. It can be<br />

seen that the high completitudes, in particular in the soft and XID bands, make<br />

the XMS sample specially suitable for cosmic evolution works (see Chapter 5).<br />

The identification breakdown of the XMS sample in several bands is summarised<br />

in Table 2.6, and the different source types are based on those of [13]. Extra-<br />

46


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

Table 2.5 Identification completeness of the XMS sample. The first column is the sample<br />

energy band, the second column N is the total number of sources in that band, the third<br />

column N id is the total number of espectroscopically identified sources, and the fourth<br />

column is the fraction of identified sources.<br />

Sample N N id Fraction (%)<br />

Soft 210 202 96<br />

Hard 159 134 84<br />

XID 284 263 93<br />

Ultrahard 70 60 86<br />

galactic sources showing broad emission lines (velocity widths 1500 km s −1 )<br />

are classified as Broad Line Active Galactic Nuclei (BLAGN, see section 1.5.4).<br />

Those showing only narrow emission lines are termed Narrow Emission Line<br />

Galaxy (NELG). Those with galaxy spectra without obvious emission lines are<br />

plainly classified as Galaxy (Gal) although there are two exceptions: two sources<br />

falling in this category were already catalogued as BL Lac objects and we have<br />

thus mantained this classification, and if a qualitative inspection of the optical<br />

images shows evidences of galaxy concentration, the source is classified as a<br />

Galaxy Cluster (Clus). Finally, all X-ray sources with an stellar espectrum with<br />

or without coronal activity are labeled as Star/AC.<br />

This classification is functional but it has some limitations. For instance, in those<br />

sources labeled as NELG no line diagnosis was carried out to check whether the<br />

source hosts an AGN or not. This is due to the relatively broad range spanned<br />

by these sources and the rather narrow wavelength coverage of the optical spectra<br />

(in particular for the fibres) the number of lines detected is small and the<br />

typical diagnostic lines drop out the spectrum with redshift. Moreover, the<br />

quality of the spectrum is not good enough to detected very weak lines needed<br />

for these diagnostics. In principle the sources classified here as NELG could be<br />

a mixture of type-2 AGN and star-forming galaxies but, given their high X-ray<br />

luminosities, we have concluded that most of them are indeed AGN-powered<br />

sources.<br />

Other limitation arise in the case of galaxy clusters. Only two sources were clas-<br />

47


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

Table 2.6 Identification breakdown of the XMS sample. The first column is the type of<br />

source (see text for the definitions), the next four columns are the number of sources in<br />

each category for the Soft, XID, Hard and Ultrahard bands, respectively. In parenthesis<br />

is the fraction of objects in each class with respect to the total of sources in that band.<br />

Type Soft XID Hard Ultrahard<br />

Total 210 284 159 70<br />

BLAGN 149 (71%) 191 (67%) 83 (52%) 39 (56%)<br />

NELG 27 (13%) 39 (14%) 35 (22%) 16 (23%)<br />

Gal/Clus 9 (4%) 11 (4%) 11 (7%) 3 (4%)<br />

Star/AC 15 (7%) 20 (7%) 3 (2%) 0 (0%)<br />

BL Lac 2 (1%) 2 (1%) 2 (1%) 2 (3%)<br />

Un-ID 8 (4%) 21 (7%) 25 (16%) 10 (14%)<br />

sified as such and this low number is caused by a bias in the detection and classification<br />

method. The detection method was designed for point-like sources<br />

and therefore it could be missing a certain number of extended X-ray sources.<br />

Furthermore, the presence of a cluster of galaxies could not be very obvious on<br />

the optical images and the source might have been misclassified as galaxy with<br />

or without emission lines.<br />

Redshifts have been estimated by matching the most prominent features in<br />

emission or absorption to characteristic line wavelengths displaced at different<br />

redshifts. In some cases, template spectra for AGN and galaxies with a<br />

range of spectroscopic classes were used to assist when no prominent features<br />

were found in the spectra. Some examples of spectra and identifications of XMS<br />

sources are shown in Figure 2.1.<br />

2.3 Other surveys<br />

In this section we will overview all the additional surveys we have made use of<br />

throughout this work in order to complement AXIS (see Chapter 3) and XMS<br />

(see Chapter 5).<br />

48


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

GUIspec z=1.958<br />

GUIspec z=0.238<br />

Flux (cgs/Å)<br />

2×10 −17 4×10 −17 6×10 −17<br />

SiIV CIV<br />

OIV<br />

NIV]<br />

[NIII]<br />

CIII]<br />

MgII<br />

Flux (cgs/Å)<br />

0 5×10 −17 10 −16 1.5×10 −16<br />

[OII]<br />

[OII]<br />

H−edge<br />

He<br />

CaIIK<br />

CaIIH<br />

Hd<br />

CHG<br />

CHG<br />

Hg<br />

CHG [OIII]<br />

Hb<br />

[OIII]<br />

[OIII]<br />

MgI<br />

MgI<br />

MgI<br />

HeI<br />

NaI<br />

NaI<br />

Flux (cgs/Å)<br />

0 2×10 −16 4×10 −16 6×10 −16<br />

4000 5000 6000 7000 8000 9000<br />

Wavelength (Å)<br />

He CHG<br />

CaIIK CHG Hg<br />

CaIIHHd CHG<br />

GUIspec z=0.301<br />

4000 5000 6000 7000 8000<br />

Wavelength (Å)<br />

Hb<br />

MgI<br />

MgI<br />

MgI<br />

NaI<br />

NaI<br />

Flux (cgs/Å)<br />

0 10 −15 2×10 −15 3×10 −15<br />

4500 5000 5500 6000 6500 7000<br />

Wavelength (Å)<br />

He CHG<br />

CaIIK CHG Hg<br />

CaIIHdCHG<br />

Hb<br />

MgI<br />

MgI<br />

MgI<br />

GUIspec z=0.0<br />

NaI<br />

NaI<br />

4000 5000 6000 7000 8000<br />

Wavelength (Å)<br />

Ha<br />

Figure 2.1 Examples of identified XMS sources. Top left: Broad Line AGN (BLAGN);<br />

Top right: Narrow Emission Line Galaxy (NELG); Bottom left: Galaxy; Bottom right:<br />

Star/Active Corona. Overplotted are the most prominent features detected in the spectra.<br />

49


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

2.3.1 Chandra Deep Field<br />

The deepest survey in the soft and hard bands so far is the Chandra Deep Field<br />

North (CDF-N, [19]), obtained from a total exposure of 2 Ms in a Northern<br />

hemisphere location. In a complementary effort in the Southern hemisphere,<br />

the Chandra Deep Field South (CDF-S, [79], [194]) obtained a total exposure<br />

time of 1 Ms. The source counts of both samples are widely discussed in [19].<br />

There are a total of 442 soft and 313 hard sources in the CDF-N, and 282 soft<br />

and 186 hard sources in the CDF-S, covering an area on the sky of ∼0.125 deg 2<br />

and ∼0.108 deg 2 , respectively.<br />

Recent Chandra calibrations have increased the estimate of the ACIS instrument<br />

effective area above 2 keV. We have used the utilities available in the Chandra<br />

calibration database to compare data from Cycle 8 to those from Cycle 5. It<br />

turned out that above 2 keV the relation between the effective areas from both<br />

Cycles is reasonably flat and can be well approximated to a constant factor of<br />

about 12%. This increase in the 2-8 keV fluxes is unlikely to be caused by the<br />

increasing contamination in the optical blocking filter since the latter only manifests<br />

itself below 1 keV.<br />

All the CDF-N and CDF-S hard fluxes and their corresponding errors, which<br />

were taken from [19] and belonged to Cycle 5, were therefore decreased by a<br />

12% factor. Moreover, when comparing the sky areas calculated individually<br />

for each source with those calculated from interpolating the sky area of the full<br />

survey, the faintest sources in both CDF samples showed large discrepancies<br />

up to several orders of magnitude. In the study of the source counts of X-ray<br />

sources (see Chapter 3) we make use of a model of the sky area for the full<br />

survey (see 3.2.4), so we have thus decided not to use soft sources fainter than<br />

3 and 7 × 10 −17 erg cm −2 s −1 in the CDF-N and CDF-S, respectively.<br />

For the X-ray luminosity function analysis in Chapter 5 we have only used the<br />

CDF-S sample, since this is the survey with the most secure identifications (up<br />

to ∼60%). The optical imaging strategy and optical counterparts catalogue can<br />

be found in [80], while the spectroscopic identifications have been taken from<br />

[217]. In both cases the ground-based facilities used were the FORS1 and FORS2<br />

instruments at the VLT. The redshifts of the unconclusive identifications and<br />

unidentified sources have been taken from photometric redshift estimations of<br />

[246]. These estimations make use of 12-band data in near ultraviolet, optical,<br />

infrared and X-rays along with sets of power law models for type-1 AGN<br />

50


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

and various templates of galaxies and have been calculated using two parallel<br />

models: HyperZ ([21]) and the Bayesian Model BPZ ([20]). The photometric<br />

redshift estimations were checked against the secure spectroscopic identifications<br />

in [217] and the COMBO-17 survey ([237], [239], [238]) to ensure their<br />

reliability, matching well that of both surveys.<br />

2.3.2 ASCA Medium Sensitivity Survey<br />

The ASCA Medium Sensitivity Survey (AMSS, [225]) is one of the largest high<br />

galactic latitude, broad band, X-ray surveys to date. It includes 606 sources in<br />

the 2-10 keV band over an sky area of 278 deg 2 with hard band fluxes spanning<br />

between 10 −13 and 10 −11 erg cm −2 s −1 , which have been used in Chapter 3.<br />

For the X-ray luminosity function (Chapter 5), we have used a flux-limited subsample<br />

of the AMSS in the northern sky (AMSSn). The flux limit of this subsample<br />

is 3 × 10 −13 erg cm −2 s −1 in the 2-10 keV band, and all but one of the 87<br />

hard X-ray selected sources have been optically identified, covering an area of<br />

∼68 deg 2 . Details on the imaging and spectroscopic identifications procedures<br />

can be found in [1].<br />

2.3.3 XMM-Newton Bright Serendipitous Survey<br />

The XMM-Newton Bright Serendipitous Survey (BSS, [50]) is a bright sample<br />

of XMM-Newton high galactic latitude sources down to a flux of 7 × 10 −14 erg<br />

cm −2 s −1 in the 0.5-4.5 keV band, with an uniform coverage of ∼28 deg 2 above<br />

that flux. This ideally complements the AXIS survey since it is selected exactly<br />

in the same energy band (XID) with the same X-ray observatory (but with a<br />

different detector: EPIC-MOS2 instead of the EPIC-pn used for AXIS), and it<br />

is both wider and shallower. For our source counts calculation purposes (see<br />

Chapter 3) we have used a total of 389 sources from this survey, including those<br />

identified as stars that have been excluded from the analysis performed in [50].<br />

2.3.4 XMM-Newton Hard Bright Serendipitous Survey<br />

Similarly to the BSS, the XMM-Newton Hard Bright Serendipitous Survey (HBSS,<br />

[50]) is a sample of sources at high galactic latitudes detected in the 4.5-7.5 keV<br />

band down to the same flux limit (7 × 10 −14 erg cm −2 s −1 ), with a flat sky cov-<br />

51


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

erage of ∼25 deg 2 . The sample contains 67 sources that have been used to<br />

complement the AXIS and XMS samples in the ultrahard band.<br />

All of these sources but two have been spectroscopically identified, as reported<br />

in [48], with 62 of them being AGN. Along with the redshifts, information on<br />

the intrinsic absorption column densities are also provided in [48] which allow<br />

us to model the absorption of these hard sources (most of them are likely to<br />

be type-2 AGN) and convolve it with the X-ray luminosity function in order to<br />

obtain a purely observation-based evolution model for these objects (see section<br />

5.2).<br />

2.3.5 ROSAT International X-ray/Optical Survey<br />

The ROSAT International X-ray/Optical Survey (RIXOS, [147]) is a medium<br />

sensitivity survey of high galactic latitude sources. The sample contains 401<br />

sources divided in two subsamples: 64 ROSAT fields with 296 sources down to<br />

a flux limit of 3 × 10 −14 erg cm −2 s −1 (known as RIXOS3), plus 18 further fields<br />

containing 105 sources down to a flux limit of 8 × 10 −14 erg cm −2 s −1 in the<br />

0.5-2 keV (RIXOS8).<br />

RIXOS3 covers 15.77 deg 2 in the sky while RIXOS8 covers 4.44 deg 2 , to a total<br />

survey area of ∼20.2 deg 2 . The RIXOS3 subsample has been spectroscopically<br />

identified in a 94%, and the RIXOS8 subsample has been completely identified<br />

which makes the full sample ideal for evolution studies in the soft band.<br />

2.3.6 ROSAT Deep Survey - Lockman Hole<br />

The ROSAT Deep Survey (RDS, [102]) collects all the observations performed<br />

by ROSAT in the period 1990-1997 in the direction of the Lockman Hole, which<br />

is one of the areas of the sky with a minimum of the Galactic Hydrogen column<br />

density ([138]). This sample comprises 50 sources down to a flux limit of<br />

5.5 × 10 −15 erg cm −2 s −1 in the 0.5-2 keV band, with an sky coverage of about<br />

∼0.3 deg 2 . All but four of the 50 sources that compose the RDS have been spectroscopically<br />

identified ([204]), the vast majority of them being AGN. This deep<br />

pencil-beam sample ideally complements the XMS soft sample and hence it has<br />

been used in the calculations done in Chapter 5.<br />

52


CHAPTER 2: AGN SURVEYS WITH XMM-Newton<br />

2.3.7 ROSAT Bright Survey<br />

The ROSAT Bright Survey (RBS, [70], [207]) aimed to identify the brightest<br />

∼2000 sources detected in the ROSAT All-Sky Survey (RASS, [232]) at high<br />

galactic latitudes, excluding the Magellanic Clouds and the Virgo cluster, with<br />

PSPC count rates above 0.2 s −1 . This sample cointains bright sources (flux limit<br />

∼ 2.5 × 10 −12 erg cm −2 s −1 in the 0.5-2 keV band) with redshifts mainly below<br />

1 and a total sky coverage of ∼20300 deg 2 . We have selected a subsample of<br />

∼600 objects that have been completely identified via spectroscopic observations,<br />

most of them being AGN, plus a strong population of galaxy clusters.<br />

53


Chapter 3<br />

Source counts of the AXIS sources<br />

3.1 Motivation<br />

The X-ray Background ([78], see section 1.3) is a record of the accretion power<br />

history of the Universe [66], and seeking its origin and composition has been a<br />

cornerstone in the modern X-ray Astronomy [64]. As a consequence of the diverse<br />

studies on the the Cosmic X-ray Background (CXB hereafter) carried out<br />

in the last decades, it is now widely accepted that the CXB emission arises from<br />

the integrated emission of discrete extragalactic sources. At soft X-rays (0.5-<br />

2 keV) most of the CXB ( 90%) has been resolved into individual sources using<br />

deep pencil-beam surveys ([137],[19],[80]), with smaller resolved fractions<br />

at higher energies. Most of these resolved sources are obscured and unobscured<br />

AGN, while the normal galaxies population only contributes significantly at the<br />

faintest fluxes [113]. Models of the CXB combining obscured and unobscured<br />

AGN populations have reproduced successfully the CXB spectrum. However,<br />

although it has been widely confirmed that the evolution of the X-ray luminosity<br />

is well described by a luminosity-dependent density evolution (see Chapter<br />

5 for details) it is not clear whether the fraction of obscured AGN depends on<br />

the luminosity ([224],[104]), redshift ([220]), both ([130]) or even none of them<br />

([57]).<br />

The fainter sources from deep pencil-beam surveys do not contribute much<br />

to the CXB intensity and it is often difficult to study them individually both<br />

optically and in X-rays. Wide shallow surveys over a large portion of the sky<br />

detect minority populations and allow studies on individual sources but again<br />

scarcely contributing to the CXB.<br />

55


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Medium depth wide band surveys which combine many sources with a relatively<br />

wide sky coverage ([103],[38],[12]) are potentially useful to solve the CXB<br />

problem, since almost 50% of the CXB below 10 keV comes from sources at intermediate<br />

fluxes ([64]) and the X-ray spectrum of many of them can be studied<br />

individually ([149]).<br />

In this chapter we will calculate the sky density or source counts of the AXIS<br />

sources (see 2.1), since these distributions are dependent on the cosmological<br />

properties of the source populations that configure the CXB emission. Accurately<br />

constraining the shape of the source counts is essential to determine the<br />

contribution from discrete sources to the CXB. This chapter is structured as follows:<br />

in section 3.2 we will explain the method used in the source counts calculation,<br />

the results will be discussed in section 3.3 and, finally, in section 3.4 we<br />

will calculate and discuss the contribution of the sources to the total CXB intensity<br />

using the obtained source counts results. For simplicity, in this chapter we<br />

will use cgs as a shorthand for the cgs system units for the flux: erg cm −2 s −1 .<br />

3.2 Method<br />

In this section we will explain the method used to calculate the sky density<br />

of the detected sources in the AXIS survey as a function of flux (log N − log S<br />

relations) in four energy bands: Soft (0.5-2 keV), Hard (2-10 keV), XID (0.5-<br />

4.5 keV) and Ultrahard (4.5-7.5 keV).<br />

3.2.1 The sky areas<br />

In order to compute the log N − log S we need to determine the sky area over<br />

which we are sensitive for a given flux. This can be obtained by summing the<br />

sky area corresponding to the values of the sensitivity maps that are below that<br />

flux. As the conversion between count rate and flux depends on the spectral<br />

shape, for each band i we have calculated the sky area values at each flux for<br />

each spectral slope Ω i (S, Γ). To do this, we have converted S to count rate using<br />

Γ and the response matrices in each field of the survey and then summed the<br />

area for which the corresponding sensitivity map is below that rate. The overall<br />

sky area thus obtained is just the sum of the areas found in each field.<br />

Therefore, the sky area in each band i for each source j would be Ω ji = Ω i (S ji , Γ ji ),<br />

56


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Figure 3.1 Sky area of the AXIS survey as a function of flux for different energy bands<br />

where the fluxes S ji and the slopes Γ ji are those explained in section 2.1.<br />

But we are interested in obtaining sky areas independent of the spectral shape<br />

Ω i (S). This can be calculated by weighting Ω i (S, Γ) with the number of detected<br />

sources in each (S, Γ) bin. This spectrally averaged areas are the sky<br />

coverage of our survey (taking into account the excluded areas), summing a<br />

maximum value of 4.8 deg 2 (see Fig. 3.1).<br />

3.2.2 The binned log N − log S<br />

The number of sources per unit flux and unit sky area at a given flux S is called<br />

differential log N − log S. A binned estimate for it can be built by summing the<br />

inverse of the Ω ji for the sources in each flux bin and then dividing that number<br />

by the width of the bin:<br />

dN<br />

dSdΩ = ∑ j 1/Ω ji<br />

∆S<br />

(3.2.1)<br />

The errors are calculated by dividing the above value by the square root of the<br />

number of sources in each bin. We have chosen to have a minimum number of<br />

sources per bin for reasons that will be explained below. This means that the<br />

bin width is determined by the flux of the first source in the bin and the flux<br />

57


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

of the first source in the next bin, which can make all of them to have different<br />

widths in principle. The width of the last bin would be undefined then, but we<br />

have dropped the brightest source of each sample and used its flux as the upper<br />

bound for this bin when we compute the binned log N − log S.<br />

The number of sources per unit sky area with fluxes higher than a given flux<br />

N(> S) is the integral log N − log S. In this work, each bin contains 15 sources<br />

(except the last one which can contain up to 29 sources) in order to avoid some<br />

features that arise due to fluctuations caused by a few sources, especially at<br />

bright fluxes where the number of sources is low. The N(> S) is built by just<br />

summing the inverse of Ω ji for all sources with fluxes above the lower limit of<br />

each bin:<br />

N(> S) = ∑<br />

j<br />

1<br />

Ω ji<br />

(3.2.2)<br />

The errors are calculated by dividing the N(> S) by the square root of the<br />

number of sources with fluxes equal or greater than the lower limit of the bin.<br />

It may happen that sources below the survey’s flux limit are promoted into<br />

the survey due to statistical fluctuations in their fluxes. Analogously, the faint<br />

sources placed just above the flux limit could drop out of the survey but, as<br />

there are many more faint sources than bright ones, the former have more probabilities<br />

of being promoted than the latter of being demoted. This leads to a net<br />

increase in the source counts at very faint fluxes, thus re-steepening the curve.<br />

This effect is known as the Eddington bias ([58]) and can be observed in the<br />

AXIS Ultrahard source counts (see Fig. 3.3 bottom panel).<br />

3.2.3 The log N − log S model<br />

In previous works it has been shown that a log N − log S is well represented<br />

by a double power law with a steep slope at bright fluxes and a flatter slope<br />

at faint fluxes ([225],[19],[160],[10],[102],[32]). Therefore, we have adopted the<br />

following model for the differential log N − log S:<br />

⎧ ( ) ⎫<br />

⎪⎨<br />

−Γd<br />

K SSb ⎪⎬<br />

dN(S)<br />

dSdΩ = Sb<br />

; S ≤ S b<br />

( )<br />

⎪ −Γu (3.2.3)<br />

⎩ K SSb<br />

S<br />

; S > S<br />

⎪<br />

b<br />

b<br />

⎭<br />

58


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

This model has four independent parameters: the normalisation K, the break<br />

flux at which the change in the slope occurs S b , the slope at bright fluxes Γ u ,<br />

and the slope at faint fluxes Γ d . When there are no significant changes in the<br />

slope because the flux limit of our survey is too close to (or above) the break<br />

flux and there are insufficent sources to perform the fit, we have used a single<br />

power law model in which we have fixed S b ≡ 10 −14 cgs and used a single<br />

slope leaving only two independent variables, K and Γ:<br />

dN(S)<br />

dSdΩ =<br />

K ( ) S −Γ<br />

10 −14 10 −14 (3.2.4)<br />

The integral log N − log S (number of sources above a flux S per unit sky area)<br />

form is thus:<br />

N(> S) =<br />

∫ ∞<br />

S<br />

dS dN(S)<br />

dSdΩ<br />

(3.2.5)<br />

and, therefore, the expected number of sources with fluxes between S min and<br />

S max assuming a given log N − log S is:<br />

N(S min ≤ S ≤ S max ) =<br />

∫ Smax<br />

S min<br />

dS dN(S) Ω(S) (3.2.6)<br />

dSdΩ<br />

3.2.4 Maximum likelihood fit method<br />

We have fitted our data to the models described in section 3.2.3 using a Maximum<br />

Likelihood method that takes into account the uncertainties in the fluxes<br />

of the sources, and the dependence of the sky areas with the flux.<br />

More formally, the expression to be minimised is:<br />

L = −2<br />

(<br />

N sample N∑<br />

)<br />

∑ log(P i (S j )) − 2 log(P λi (N))<br />

i=1 j=1<br />

(3.2.7)<br />

where P i (S j ) is the probability of finding a source with flux S j in the sample i:<br />

P i (S j ) =<br />

∫ S ′ max<br />

S min<br />

′<br />

exp<br />

dS dN(S)<br />

dSdΩ Ω i(S)<br />

59<br />

λ<br />

( )<br />

−(S j −S) 2<br />

2σ<br />

j<br />

2 √<br />

2πσj<br />

(3.2.8)


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

and P λi (N) is the Poissonian probability of finding N sources in a sample i when<br />

the expected number of sources is λ i (equivalent to equation 3.2.6):<br />

P λi (N) = e−λ iλ N i<br />

N!<br />

(3.2.9)<br />

λ i =<br />

∫ Smax<br />

S min<br />

dS dN(S)<br />

dSdΩ Ω i(S) (3.2.10)<br />

This likelihood expression takes into account not only the variation of sky area<br />

in a sample i with the flux, Ω i (S) but also the uncertainty in the flux of the<br />

source σ j ([173]) via the exponential term in the numerator of equation 3.2.8,<br />

assuming that the error around the measured flux value follows a Gaussian<br />

distribution. Since the tails of a Gaussian decrease very quickly, we have used<br />

S ′ min = max(S j − 4σ j , S min ) and S ′ max = min(S j + 4σ j , S max ) as the integration<br />

limits. Additionally, with this method we also consider the normalisation K as<br />

a free parameter; since it appears both in the numerator and the denominator<br />

of equation 3.2.8, we have included the second term in equation 3.2.7 to have<br />

it taken into account. This approach is fully equivalent to that described by<br />

[145] to determine the luminosity function parameters of a complete sample of<br />

AGN (see section 5.3 for details) and, essentially, it takes into account the full<br />

distribution of sources in the Ω(S) plane.<br />

The 1-σ uncertainties in the log N − log S parameters are estimated from the<br />

range of values of each parameter around the minimum L min that make ∆L =<br />

1. The procedure is as follows: the parameter of interest is fixed to a value<br />

close to the best-fit value and we let the rest of the parameters vary until a new<br />

minimum is found. This is repeated for several values of the parameter until<br />

the new minimum L new<br />

min<br />

satisfies Lnew<br />

min = L min + 1.<br />

60


61<br />

Table 3.1 Maximum likelihood fit results to the log N − log S in different bands and using different samples: the first column is the band<br />

used, the second is the power-law slope above the flux break, the third is the slope below that break, the fourth is the flux break, the fifth is<br />

the normalisation, the last six columns indicate the number of sources from each sample used in the fit.<br />

N used /N tot<br />

Band Γ u Γ d S b K AXIS BSS HBS CDF-N CD F-S AMSS<br />

(10 −14 cg s) (deg −2 )<br />

Soft 2.39 +0.06<br />

−0.06<br />

1.69 +0.07<br />

−0.06<br />

1.15 +0.16<br />

−0.13<br />

123.4 +18.8<br />

−17.1<br />

1267/1267<br />

Soft f 2.38 +0.14<br />

−0.09<br />

1.56 +0.01<br />

−0.01<br />

1.02 +0.14<br />

−0.17<br />

141.4 +42.1<br />

−7.3<br />

1267/1267 429/442 a 269/282 b<br />

Hard 2.72 +0.07<br />

−0.08<br />

- 1.00 684.4 +74.1<br />

−84.0<br />

348/397 c<br />

Hard 2.03 +0.12<br />

−0.11<br />

1.00 +0.11<br />

−0.12<br />

0.30 +0.07<br />

−0.05<br />

1086.6 +134.3<br />

−146.5<br />

313/313<br />

Hard 2.51 +0.50<br />

−0.28<br />

0.89 +0.12<br />

−0.12<br />

0.72 +0.11<br />

−0.10<br />

743.7 +109.6<br />

−105.6<br />

186/186<br />

Hard 2.12 +0.13<br />

−0.01<br />

1.10 +0.01<br />

−0.01<br />

0.44 +0.04<br />

−0.01<br />

799.1 +226.8<br />

−8.4<br />

313/313 186/186<br />

Hard 2.66 +0.08<br />

−0.05<br />

1.20 +0.01<br />

−0.01<br />

1.00 +0.08<br />

−0.01<br />

611.5 +49.1<br />

−34.3<br />

397/397 313/313 186/186<br />

Hard 2.58 +0.02<br />

−0.02<br />

- 1.00 606.5 +46.8<br />

−46.3<br />

348/397 c 606/606<br />

Hard 2.53 +0.25<br />

−0.18<br />

1.18 +0.14<br />

−0.08<br />

0.92 +0.66<br />

−0.19<br />

607.8 +366.4<br />

−208.1<br />

313/313 186/186 606/606<br />

Hard f 2.58 +0.17<br />

−0.02<br />

1.30 +0.01<br />

−0.01<br />

1.17 +0.01<br />

−0.05<br />

485.3 +10.1<br />

−24.3<br />

397/397 313/313 186/186 606/606<br />

XID 2.46 +0.11<br />

−0.07<br />

1.29 +0.09<br />

−0.18<br />

1.45 +0.16<br />

−0.26<br />

212.2 +47.4<br />

−22.8<br />

1359/1359<br />

XID f 2.54 +0.03<br />

−0.04<br />

1.35 +0.06<br />

−0.25<br />

1.64 +0.13<br />

−0.28<br />

193.0 +33.1<br />

−15.8<br />

1359/1359 389/389<br />

Ultrahard 2.59 +0.09<br />

−0.05<br />

- 1.00 95.0 +11.1<br />

−12.8<br />

84/89 f<br />

Ultrahard f 2.62 +0.10<br />

−0.10<br />

- 1.00 102.2 +20.0<br />

−21.1<br />

84/89 d 58/65 e<br />

a S min,CDF−N = 3 × 10 −17 cgs; b S min,CDF−S = 6 × 10 −17 cgs; c S min,AXIS = 1.5 × 10 −14 cgs; d S min,AXIS = 10 −14 cgs; e S max,HBS = 2.5 × 10 −13 cgs;<br />

CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

f<br />

Best fit used for contribution to CXB


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

3.3 Results<br />

In order to span a wider range of fluxes and hence to achieve greater accuracy<br />

in the source counts parameters determination, we have combined the AXIS<br />

sample with other shallower and deeper surveys in all the energy bands under<br />

study (see 2.3 for details on them). In the Soft and Hard band we have added<br />

both Chandra Deep Field surveys ([19], [80], [79]) which give us coverage at faint<br />

fluxes. In addition, in the Hard band we have also added the AMSS ([225])<br />

survey which helps us to constrain better the slope at the bright end of the<br />

log N − log S. In both XID and Ultrahard bands we have combined the AXIS<br />

sources with those of the BSS and HBSS ([50]) in their respective energy bands.<br />

Adding these samples will help us to better constrain the shape of the log N −<br />

log S at both bright and faint ends and hence diminishing the uncertainties in<br />

the source counts parameters.<br />

The results of the Maximum Likelihood fits to various single and combined<br />

surveys in different energy bands are summarized in table 3.1. The flux interval<br />

used in the fit is S min = 10 −17 and S max = 10 −12 unless stated otherwise. This<br />

numbers have been chosen to span the observed fluxes of the sources, but the<br />

results are not very dependent on them since the sky area drops abruptly at low<br />

fluxes and the sky density of bright sources is very small. The best fit curves<br />

are plotted along with the observed integral log N − log S in figures 3.2 and 3.3.<br />

For each energy band in Table 3.1 we have performed several fits. In the first<br />

two rows we have fitted the AXIS sources alone, using a fixed spectral index to<br />

calculate fluxes and sky areas (first row) or the best fit spectral index for each<br />

source (second row). In the former case, the spectral slopes were Γ = 1.8 in<br />

the Soft and XID bands and Γ = 1.7 in the Hard and Ultrahard bands. It can<br />

be seen that the results of the fit are mutually compatible within the error bars,<br />

although the uncertainties in the parameters are larger when we fix the spectral<br />

index (especially in the XID and Ultrahard bands). These differences in the bestfit<br />

log N − log S parameters between both methods can be (partly) explained in<br />

terms of the uncertainties in the fitted fluxes of the sources. When the photon<br />

index for the power law slope of the spectrum of the source is obtained via their<br />

spectral fitting, the uncertainties in the fitted fluxes thus obtained are smaller<br />

than those calculated using a fixed spectral slope. In light of this, we have thus<br />

used the spectral fitted slopes to compute the fluxes and count rates for the<br />

AXIS sources.<br />

62


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Figure 3.2 Integral log N − log S in different bands (Top panel: Soft, Bottom panel: Hard),<br />

along with our best fits and some previus results<br />

63


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Figure 3.3 Integral log N − log S in different bands (Top panel: XID, Bottom panel: Ultrahard),<br />

along with our best fits<br />

64


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

The absolute value of the likelihood, unlike the χ 2 , is not an indicator of the<br />

goodness of the fit. Although there appears to be a good agreement between<br />

the data and the best fit model in Figs. 3.2 and 3.3 this is only a qualitative impression.<br />

We have therefore plotted the ratio between the binned differential<br />

log N − log S and the best fit model so that any significative deviations from<br />

unit would arise (see Figs. 3.4 and 3.5). Overplotted is the 1-σ uncertainty interval<br />

calculated as follows: for each flux it has been calculated the differential<br />

log N − log S in the corners of the hypercube defined by the 1-σ errors of the<br />

best fit parameters and then taken the minimum and maximum values.<br />

The agreement with the CDF samples is very good in general. In the Soft band,<br />

the combined AXIS-CDF fit is in excellent agreement with the slope provided<br />

by [19] (which is what we expected since at faint fluxes the only contribution to<br />

the log N − log S comes from the CDF sources) whereas the AXIS-only fit has<br />

a steeper slope below the flux break. However, it is compatible with the AXIS-<br />

CDF Γ d slope at


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

them to rise and drop above and below this flux respectively thus explaining<br />

the bump seen in Fig. 3.4.<br />

When we fit only the CDF and AMSS samples we can reproduce the overall<br />

behavior of the log N − log S since the former constrain the Γ d value at faint<br />

fluxes and the AMSS constrains the Γ u value at bright fluxes. Hence, the best fit<br />

parameters between CDF+AMSS and AXIS+CDF+AMSS are consistent within<br />


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Figure 3.4 Ratio between the differential log N − log S and the best fit model in different<br />

bands (Top panel: Soft, Bottom panel: Hard). The dashed bowtie is the 1-σ uncertainty in<br />

the log N − log S<br />

67


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Figure 3.5 Ratio between the differential log N − log S and the best fit model in different<br />

bands (Top panel: XID, Bottom panel: Ultrahard). The dashed bowtie is the 1-σ<br />

uncertainty in the log N − log S<br />

68


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

when we consider all sources (the Maximum Likelihood method is very insensitive<br />

to a few discrepant points). The AXIS-only and AXIS-HBS fits are<br />

very similar and compatible within −0.2 and HR < −0.2, respectively).<br />

To construct the log N − log S it would formally be necessary to recalculate<br />

the effective area of these subsamples but when we used our method (see<br />

69


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Table 3.2 Maximum likelihood fit results to the log N − log S in the Soft band for two<br />

HR selected subsamples. The first column is the band used, the second is the powerlaw<br />

slope above the flux break, the third is the slope below that break, the fourth is the<br />

flux break, the fifth is the normalisation, and the sixth column indicate the number of<br />

sources from each sample used in the fit.<br />

N used /N tot<br />

Band Γ u Γ d S b K AXIS<br />

(10 −14 cg s) (deg −2 )<br />

Soft HR < −0.2 2.33 +0.04<br />

−0.04<br />

1.69 +0.03<br />

−0.04<br />

1.40 +0.20<br />

−0.10<br />

88.7 +5.9<br />

−7.9<br />

1058/1058<br />

Soft HR > −0.2 2.77 +0.07<br />

−0.06<br />

- 1.00 16.9 +0.05<br />

−0.06<br />

209/209<br />

section 3.2.1), it produced gaps in those flux bins which have been depleted of<br />

sources after the splitting of the sample. Since it is extremely difficult to deal<br />

with this problem when the area varies rapidly with the flux (see Fig. 3.1), we<br />

decided to use as a first approximation the Soft Ω(S) for both subsamples. A<br />

similar approach has been used in [125] where they use the corresponding area<br />

curve for each energy band under study.<br />

It can be seen in Fig. 3.6 that the source counts of Soft sources with HR < −0.2<br />

clearly still follows a broken power law model, while for the HR > −0.2 sources<br />

it is a simple power law. The best fit parameters are listed in table 3.2. Those<br />

for the softest sources (HR < −0.2) are fully compatible, well within the 1-σ<br />

errors, with the best fit parameters obtained for the Soft AXIS-only sample (see<br />

table 3.1). This is reasonable, since the AXIS Soft sources with HR < −0.2 are<br />

over 83% of the total Soft sample. However, the Soft HR > −0.2 subsample<br />

best fit parameters are more similar to those obtained for the hard AXIS-only<br />

sample.<br />

In [125] they found similar results using a discriminating value of HR = 0 for<br />

all energy bands in the Chandra Multiwavelength Project (ChaMP). They demostrated<br />

that the missing break is not due to small number statistics, since<br />

the HR < 0 sources largely outnumbered the HR > 0 sources, by performing<br />

simulations in randomly selected subsets of the HR < 0 sample with the same<br />

number of sources as the HR > 0 subsample. They calculated the log N − log S<br />

for these subsets finding that, in spite of the reduced statistics, they all showed<br />

a detectable break. A possible explanation for this behavior is that spectrally<br />

70


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

N(>S) (sources/deg 2 )<br />

10 100<br />

10 −15 10 −14 10 −13<br />

S 0.5−2 keV<br />

(cgs)<br />

N(>S) (sources/deg 2 )<br />

10 100<br />

10 −15 10 −14 10 −13<br />

S 0.5−2 keV<br />

(cgs)<br />

Figure 3.6 Integral log N − log S for two HR selected 0.5-2 keV AXIS subsamples (Top<br />

panel: HR < −0.2, Bottom panel: HR > −0.2). The solid line is the best fit curve.<br />

71


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

hard sources (HR > −0.2), which are likely to be obscured, are detected predominantly<br />

at relatively close redshifts and hence do not show the break flux in<br />

the source counts that is an observable effect of cosmological evolution ([97]).<br />

3.4 Contribution to the Cosmic X-ray Background<br />

Now that we have accurately determined the parameter of the source counts<br />

we can estimate the intensity contributed by sources in different flux intervals<br />

using:<br />

I(S min ≤ S ≤ S max ) =<br />

∫ Smax<br />

S min<br />

dSS dN(S)<br />

dSdΩ<br />

(3.4.1)<br />

We are interested here in comparing the total intensity contributed by resolved<br />

sources with the total CXB intensity. Before that, we need to estimate both the<br />

contribution from bright sources and the total CXB intensity.<br />

3.4.1 Contribution from bright sources and stars<br />

Since the surveys we are using in this section to calculate the log N − log S<br />

clearly lack bright sources we must first estimate which would be their contribution<br />

to the CXB, which is small but not negligible.<br />

In the Soft band we have followed the same method as [160] summing the soft<br />

fluxes of the sources in the RBS ([207], see section 2.3) higher than 10 −11 cgs,<br />

and dividing by the area covered by the survey. In the Hard band we have<br />

used the sources in the HEAO-1 A2 survey [180], which is complete down to<br />

3.1 × 10 −11 cgs. For the Ultrahard band we have taken the values from the<br />

Hard band converting both the flux limit and the intensities to 4.5-7.5 keV using<br />

a power law with slope Γ = 1.7. The XID band is a bit more problematic since<br />

it is partly in the Soft band and partly in the Hard band. Here we have just<br />

estimated the 2-4.5 keV flux limit and intensities from the Hard band the same<br />

way as before and then added them to the Soft band values.<br />

Moreover, we need to estimate the contribution from galactic stars down to our<br />

fainter flux limit so that we can substract it from our source intensities, since<br />

we are pursuing to determine the resolved fraction of the extragalactic CXB.<br />

We have taken the estimations from [19] of the fraction of stars contributing<br />

72


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

to the CXB and their total CXB intensities, obtaining the following values for<br />

stars: 0.19 +0.05<br />

−0.04 × 10−12 cgs deg −2 and 0.36 +0.34<br />

−0.19 × 10−12 cgs deg −2 in the Soft and<br />

Hard bands, respectively. For the XID band, assuming that most of the contribution<br />

from stars comes from high fluxes ([19]) and that the stars at such high<br />

fluxes have a low temperature thermal spectrum ([50]), we have tried to determine<br />

the XID band contribution from the Soft band assuming a T = 0.5 MK<br />


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

of these values to the Soft band using spectral shape described above provides<br />

a result of (7.46 ± 0.09) × 10 −12 cgs deg −2 and (7.78 ± 0.06) × 10 −12 cgs deg −2 ,<br />

respectively, which are again fully compatible with our adopted value of (7.5 ±<br />

0.4) × 10 −12 cgs deg −2 .<br />

More recently, there has been a re-estimation of the total CXB intensity in the<br />

1-2 keV and 2-8 keV bands using Chandra CDF data and contributions from<br />

bright sources ([109]). They have measured the unresolved CXB by extracting<br />

spectra of the sky and then removing all point-like and extended sources detected<br />

in the CDF. Their total CXB intensity estimation comes from the sum of<br />

the unresolved components and the contributions from the CDF sources and<br />

from brighter sources for which they have used the source counts derived in<br />

[230] and [231], with different spectral slopes at different fluxes. Their 1-2 keV<br />

CXB intensity is thus (4.6 ± 0.3) × 10 −12 cgs deg −2 , quite similar to our value.<br />

3.4.3 Contribution to the CXB in different flux intervals<br />

Our calculated CXB intensities for the observed flux intervals along with the<br />

XRB intensities from [160] are shown in Table 3.3.<br />

In Fig. 3.7 is plotted the relative contribution of two flux intervals per decade<br />

to the total CXB intensity for the four energy bands under study assuming our<br />

best fit log N − log S. It can be seen that the maximum contribution comes from<br />

fluxes around ∼ 10 −14 cgs, where the break in the source counts is. The contribution<br />

from the bins one decade around this value are close to the 50% of the<br />

total CXB intensity in the Soft and Hard bands. In the Ultrahard band, we are<br />

not deep enough to detect the break, although using the 5-10 keV source counts<br />

from [101] and [194] it could be possible that a break just below ∼ 10 −14 cgs<br />

exists.<br />

The extrapolation of the Soft and Hard bands log N − log S to zero flux using<br />

our best fit model does not saturate the CXB intensity (Table 3.3), although in<br />

the Soft band the total CXB intensity is within the intensities spanned by the<br />

uncertainties in the best fit parameters. However, the possibility of a new dominant<br />

population at fainter fluxes remains open, especially in the Hard band.<br />

Assuming that the fraction of truly diffuse CXB is negligible (in [156] is shown<br />

that the source counts in the Soft band grows down to at least 7 × 10 −18 cgs) it is<br />

possible to estimate the slope needed to saturate the CXB at zero flux supposing<br />

74


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Table 3.3 Intensity in different bands from different origins and flux intervals. The first<br />

column indicates the band, the second gives the origin of the intensity, the third and<br />

fourth columns list the flux interval, the fifth column total intensity from that interval,<br />

the sixth column shows the fraction of the XRB intensity contributed by sources in that<br />

interval.<br />

Band Origin S min S max I(S min ≤ S ≤ S min ) f XRB<br />

(cgs) (cgs) (10 −12 cgs deg −2 )<br />

Soft Best fit AXIS+CDF log N − log S 0 3 × 10 −17 0.25 0.03<br />

Soft Best fit AXIS+CDF log N − log S 3 × 10 −17 2 × 10 −13 5.60 0.75<br />

Soft Best fit AXIS+CDF log N − log S 3 × 10 −17 10 −11 6.54 0.87<br />

Soft RBS sources a 10 −11 0.20 0.03<br />

Soft Total resolved 3 × 10 −17 6.55 e 0.87<br />

Soft XRB b 7.5 ± 0.4<br />

Hard Best fit AXIS+CDF+AMSS log N − log S 0 3 × 10 −16 0.62 0.03<br />

Hard Best fit AXIS+CDF+AMSS log N − log S 3 × 10 −16 1 × 10 −11 17.08 0.85<br />

Hard Best fit AXIS+CDF+AMSS log N − log S 3 × 10 −16 3.1 × 10 −11 17.18 0.85<br />

Hard HEAO-1 A4 sources c 3.1 × 10 −11 0.43 0.02<br />

Hard Total resolved 3 × 10 −16 17.25 e 0.85<br />

Hard XRB b 20.2 ± 1.1<br />

XID Best fit AXIS+BSS log N − log S 3 × 10 −15 10 −12 8.48 0.56<br />

XID Best fit AXIS+BSS log N − log S 3 × 10 −15 2.38 × 10 −11 9.00 0.59<br />

XID Bright sources d 2.38 × 10 −11 0.39 0.02<br />

XID Total resolved 3 × 10 −15 9.20 e 0.60<br />

XID XRB d 15.3 ± 0.6<br />

Ultrahard Best fit AXIS+HBS log N − log S 9 × 10 −15 2 × 10 −13 1.50 0.21<br />

Ultrahard Best fit AXIS+HBS log N − log S 9 × 10 −15 1.05 × 10 −11 1.74 0.24<br />

Ultrahard Bright sources d 1.05 × 10 −11 0.15 0.02<br />

Ultrahard Total resolved 9 × 10 −11 1.79 e 0.25<br />

Ultrahard XRB d 7.2 ± 0.4<br />

a Schwope et al. ([207])<br />

b Moretti et al. ([160])<br />

c Piccinotti et al. ([180])<br />

d<br />

See text<br />

e After subtracting the stellar contribution (see Section 3.4.1)<br />

75


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Figure 3.7 Relative contribution to the total X-ray Background intensity from equally<br />

spaced logarithmic flux intervals using the best fit models in each band: Soft (solid),<br />

Hard (dashed), XID (dot-dashed) and Ultrahard (dotted).<br />

the log N − log S steepens below the minimum flux studied here. We find out<br />

that these slopes are 1.85 and 1.84 in the Soft and Hard bands, respectively. If<br />

we compare these values to the observed slopes of the separate source counts of<br />

AGN and galaxies in the CDF field ([19]) only the absorbed AGN with an slope<br />

of ∼1.62 are close to saturate the soft CXB (the rest of AGN estimates have values<br />

in the range 1.2-1.5), whereas the galaxies could do so with any estimate<br />

(slopes ranging from 2.1 to 2.7) if we assume they grow at the same rate below<br />

the resolved fluxes. In the Hard band the situation is quite similar, being the<br />

absorbed AGN the only AGN population capable of saturating the hard CXB<br />

(slope ∼1.95, against slopes ∼1.5 for the rest of AGN) while the galaxies, having<br />

slopes ranging 3-3.5, could satisfy this condition at any estimate. Even if a<br />

higher intensity of the CXB is assumed ([109]) this conclusion is still valid since<br />

that would need a steeper source counts slope to saturate the CXB.<br />

In any case, it has been made clear that the maximum contribution to the CXB<br />

comes from sources at fluxes around ∼ 10 −14 cgs at all energy bands (including,<br />

probably, the Ultrahard band) and hence medium flux surveys with limiting<br />

fluxes close to this value, like AXIS, are essential to understand the evolution of<br />

76


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

the X-ray emission in the Universe at these energies. Sources at fainter fluxes or<br />

being heavily absorbed are very important for the overall energy content of the<br />

CXB ([85]), but they reside at harder X-ray energies where the resolved fraction<br />

is much smaller ([240]).<br />

3.4.4 Resolved and unresolved components of the CXB<br />

A direct comparation between the results in [109] and our results conveys a<br />

certain uncertainty since the energy bands under consideration are different<br />

(0.5-2 keV versus 1-2 keV in the Soft band, 2-10 keV versus 2-8 keV in the Hard<br />

band). For the conversions between these bands we have assumed a power law<br />

model with photon indices of 1.5 for the unresolved component (fitted by [109]<br />

to the unresolved CXB spectrum), 1.43 for the resolved faint sources (again, fitted<br />

by [109] to the summed spectrum of resolved sources) and 2 for the resolved<br />

sources brighter than ∼ 10 −14 cgs ([149]).<br />

Also, in [109] they have excluded the areas around the sources catalogued in<br />

[4] from their study of the unresolved CXB, with flux limits of 2.5 × 10 −17 cgs<br />

in the 0.5-2 keV and 1.4 × 10 −16 cgs in the 2-8 keV and therefore they define as<br />

unresolved any intensity coming from fluxes below those limits or from a truly<br />

diffuse component. Additionally, in the 2-8 keV band, the sources in [4] were<br />

also the basis of the work in [19], and we have seen in sections 3.3 and 2.3 that<br />

the fluxes need to be decreased by ∼12% due to changes in Chandra calibrations.<br />

Assuming all these corrections and a spectral slope of 1.43, the flux limit in the<br />

2-10 keV band becomes 1.6 × 10 −16 cgs.<br />

In table 3.4, there is a comparison between the intensities from [109] and the<br />

results of this and previous works in the 0.5-2 keV, 1-2 keV and 2-10 keV bands.<br />

The error bars in our estimates of the intensities have been derived from the<br />

errors in the log N − log S best fit parameters. For the extrapolation to zero flux,<br />

the errors have been calculated using the values spanned by the uncertainties<br />

in the best fit parameters as explained in section 3.3.<br />

The most remarkable difference is that the total soft CXB intensities are barely<br />

compatible at the ∼1-σ level, whereas at 1-2 keV they are compatible at 0.16-σ<br />

(see section 3.4.2). This discrepance is caused by the way the 0.5-2 keV intensities<br />

have been estimated. Our 0.5-2 keV value is an extrapolation from the<br />

1-2 keV total CXB intensity in [160] assuming Γ = 1.4, while in [109] they use<br />

77


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

Table 3.4 Comparison between the estimated X-ray intensities from [109] and this and<br />

previous works, from different origins and flux intervals. The first two columns indicate<br />

the flux limits between which has been estimated the intensity. If the first one is<br />

missing the intensity is calculated from zero flux, while if the second one is missing,<br />

the intensity is calculated to infinity. The third column is the intensity from [109], and<br />

the fourth column the intensity from this work (or previous ones as indicated).<br />

0.5-2 keV<br />

Flux limits<br />

Intensity<br />

(cgs) (10 −12 cgs deg −2 )<br />

[109] Here<br />

2.5 × 10 −17 1.8 ± 0.3 0.23 ± 0.09 a<br />

2.5 × 10 −17 5.0 × 10 −15 2.4 ± 0.4 2.2 ± 0.4 a<br />

5.0 × 10 −15 1.0 × 10 −11 4.2 ± 0.3 4.4 ± 1.4<br />

1.0 × 10 −11 - 0.2 b<br />

Total 8.4 ± 0.6 7.5 ± 0.4 c<br />

Total resolved 6.6 ± 0.5 6.6 ± 1.7 e<br />

Fraction resolved 0.79 ± 0.07 0.88 ± 0.23<br />

1-2 keV<br />

Flux limits<br />

Intensity<br />

(cgs) (10 −12 cgs deg −2 )<br />

[109] Here<br />

1.5 × 10 −17 1.0 ± 0.1 0.13 ± 0.05 a<br />

1.5 × 10 −17 3.0 × 10 −15 1.5 ± 0.3 1.3 ± 0.2 a<br />

3.0 × 10 −15 0.5 × 10 −11 2.1 ± 0.1 2.2 ± 0.7<br />

0.5 × 10 −11 - 0.1 b<br />

Total 4.6 ± 0.3 4.5 ± 0.2 c<br />

Total resolved 3.6 ± 0.2 3.7 ± 0.7 e<br />

Fraction resolved 0.78 ± 0.07 0.81 ± 0.16<br />

2-10 keV<br />

Flux limits<br />

Intensity<br />

(cgs) (10 −12 cgs deg −2 )<br />

[109] Here<br />

1.6 × 10 −16 4.2 ± 2.1 0.40 ± 0.04 a<br />

1.6 × 10 −16 1.6 × 10 −14 9.5 ± 1.3 9.3 ± 0.4 a<br />

1.6 × 10 −14 1.0 × 10 −11 7.0 ± 0.4 7.9 ± 1.5<br />

1.0 × 10 −11 3.1 × 10 −11 - 0.09 ± 0.06<br />

3.1 × 10 −11 - 0.4 d<br />

Total 20.7 ± 2.5 20.2 ± 1.1 c<br />

Total resolved 16.5 ± 1.3 17.3 ± 1.7 e<br />

Fraction resolved 0.80 ± 0.11 0.86 ± 0.10<br />

a Extrapolating our best fit log N − log S; b Schwope et al. ([207]); c Moretti et al. ([160]); d Piccinotti et al. ([180]); e After<br />

subtracting the stellar contribution<br />

78


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

different contributions with different photon indices. Since the contribution<br />

from the brightest sources is almost half of the total and their slopes are steeper,<br />

the effective spectral slope in the conversion of the CXB intensity in [109] from<br />

1-2 keV to 0.5-2 keV is Γ ∼ 1.72, steeper tan our value of 1.4, and hence providing<br />

a much larger contribution in the 0.5-1 keV interval. The resolved contribution<br />

in [109] also increases in the Soft band with respect to the 1-2 keV band, but<br />

the resolved fraction only increases slightly because of the very similar effective<br />

spectral slopes of the resolved component and the CXB intensity. On the contrary,<br />

our resolved component is very similar to that of [109], but our assumed<br />

CXB intensity in the Soft band has a flatter spectral shape, resulting in a smaller<br />

resolved fraction but still compatible within the errors.<br />

Furthermore, we have converted our resolved contributions from the Soft band<br />

to the 1-2 keV band using different spectral slopes for different flux intervals<br />

(as explained at the beginning of this section), and hence the differences in the<br />

resolved components and the resolved fraction with respect to those of [109]<br />

are very small (see table 3.4) and well within the errors.<br />

So, to summarize, the differences in the Soft band in the total CXB intensity are<br />

around ∼1-σ and mostly due to the different effective spectral shape assumed<br />

for the total CXB intensity between [109] and most of the previous works. On<br />

the contrary, the resolved intensities are very similar both in [109] and in this<br />

work.<br />

Assuming the Γ = 1.4 CXB spectral shape below 1 keV, where its real shape<br />

is unknown, is an important source of uncertainty. Supposing that the CXB<br />

spectrum below 1 keV steepens with a power law shape, a slope Γ ∼ 2.2 would<br />

be needed to produce the estimate of the 0.5-2 keV CXB intensity from [109]<br />

from their 1-2 keV intensity. Given that the typical slope of the faintest sources<br />

detected and of the unresolved component in the Soft band are around Γ ∼<br />

1.5, the uncertainty in the spectral shape below 1 keV cannot fully explain the<br />

difference in the CXB intensities discussed above.<br />

In the Hard band there is full agreement in the total CXB intensity as well as in<br />

the resolved component, where our estimate is slightly higher but well within<br />

the 1-σ errors. Our estimate of the bright source contribution in the Hard band<br />

is quite robust since the AMSS sources cover the same region of the log N −<br />

log S parameter space.<br />

Extrapolating the log N − log S to zero flux we cannot saturate the unresolved<br />

79


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

component so either there is a diffuse component, or the source counts has to<br />

steepen again somewhere below the current flux limits.<br />

3.5 Summary<br />

In this chapter we have performed a maximum likelihood fit to a broken power<br />

law model for the log N − log S in four different energy bands: Soft (0.5-2 keV),<br />

Hard (2-10 keV), XID (0.5-4.5 keV) and Ultrahard (4.5-7.5 keV). The results of<br />

the fit are summarized in table 3.3. We have found that fits to the source counts<br />

using fixed spectral slopes produce similar results to those using the spectral<br />

slopes of the individual sources, but with larger error bars.<br />

The source counts in the Soft, Hard and XID bands present breaks at fluxes<br />

around ∼ 10 −14 cgs. A detailed examination of the ratios between the data and<br />

the best fit to this model do not show significant differences in the Soft and XID<br />

bands, whereas in the Hard band there seems to be some evidence of another<br />

break at ∼ 3 × 10 −15 cgs and several changes of slope. This only happens in the<br />

Hard band so it is difficult to tell whether is due to the simplicity of the model<br />

(that has to take into account several contributions from different populations at<br />

different redshifts) or to calibration uncertainties. The best fit model parameter<br />

values obtained are compatible with previous smaller and similar surveys, but<br />

the combination of large number of sources and wide flux coverage in this work<br />

produce smaller uncertainties in the fitted parameters. The source counts of<br />

two HR selected subsamples in the Soft band reveals that the spectrally hard<br />

sources (those with HR > −0.2) show no flux break in their log N − log S. This<br />

may imply that such hard sources are detected at relatively low redshifts and<br />

therefore they show no evolution.<br />

We have also used our best fit log N − log S to calculate both the total resolved<br />

fraction of the Cosmic X-ray Background (including the contribution from bright<br />

sources, table 3.3) and the relative contribution in different flux bins to the total<br />

CXB (see Fig.3.7). We have used the estimates of the average CXB intensity in<br />

the Soft and Hard bands from [160] and converted them to the XID and Ultrahard<br />

bands using a power law spectrum with Γ = 1.4.<br />

The total resolved fraction down to the lowest fluxes of our combined sample<br />

reaches 88% in the Soft band and 86% in the Hard band (where we have been<br />

able to reach deep fluxes usin pencil beam surveys [19]), but it is only 60% in<br />

80


CHAPTER 3: SOURCE COUNTS OF THE AXIS SOURCES<br />

the XID band and 25% in the Ultrahard band. The total intensity produced<br />

extrapolating our log N − log S to zero flux does not saturate the CXB intensity.<br />

Assuming that another flux break occurs just below our detected minimum<br />

fluxes, we have estimated the minimum slope needed to reproduce the CXB<br />

with only discrete sources obtaining values of 1.85 in the Soft band and 1.84 in<br />

the Hard band. If we compare these values with those of the galaxy and AGN<br />

source counts from the CDF ([19]), we find that galaxies could easily provide<br />

this steepening and, among the AGN population, only the absorbed ones could<br />

do so as well.<br />

The maximum fractional contribution to the CXB in the Soft, Hard and XID<br />

bands comes from sources within a decade around ∼ 10 −14 cgs (where the<br />

break in the source counts power law approximately occurs). This is almost<br />

50% of the total contribution in the Soft and Hard bands, which means that<br />

medium deep surveys such as AXIS are essential to understand the evolution<br />

of the X-ray emission in the Universe up to 10 keV.<br />

A recent re-estimation of the Soft and Hard CXB intensities using CDF data<br />

[109] finds lower resolved fractions of the CXB than in this work. We have<br />

shown that the difference in our results in the Soft band, where it is highest, is<br />

only at the ∼1-σ level and that it is due to the different ways in which the CXB<br />

intensity is calculated. In [109] they add different contributions at different<br />

fluxes, which produces an effective slope Γ ∼ 1.8 which is much steeper than<br />

the CXB spectral slope Γ = 1.4 that we have assumed.<br />

81


Chapter 4<br />

Clustering of AXIS and XMS sources<br />

4.1 Motivation<br />

The dominant population at medium and low X-ray fluxes are AGN, which are<br />

known to cluster strongly ([242], [84], [161], [37]).<br />

To detect source clustering it is needed a survey that is both deep (to achieve<br />

high angular density) and wide (to prevent a single structure to bias the overall<br />

average). To achieve both requirements simultaneously is almost impossible,<br />

so some sort of compromise between them must be reached. The AXIS sample<br />

comprises 36 fields outside the Galactic plane (see section 2.1) and has a relatively<br />

high source density (about 35 sources per field in the Soft and XID bands).<br />

This will allow us to compute the angular correlation function of AXIS sources<br />

and to translate the results into spatial clustering inverting Limber’s equation<br />

([175], see section 4.3.5) knowing the luminosity function of the populations<br />

involved.<br />

Several studies have investigated the angular clustering in the Soft and Hard<br />

bands on different scales ranging from tens or hundreds of arcsecs to degrees,<br />

with varying success. For instance, angular clustering have been detected in the<br />

Soft band ([72], [18], [3], [229]) with different values at the clustering strength.<br />

Angular clustering in the hard band has evaded detection in some cases ([72],<br />

[187]), but not in others ([241], [17]) where clustering strengths are found to<br />

be marginally compatible with the observed spatial clustering of optically and<br />

X-ray selected AGN.<br />

The spatial correlation function provides useful information on how extragalac-<br />

83


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

tic X-ray sources cluster forming physical structures of cosmological origin although<br />

it is more expensive in terms of resources than the angular correlation<br />

function because it needs highly complete surveys (which require hundreds of<br />

secure redshift identications obtained via ground-based espectroscopy). Many<br />

works have reported spatial clustering using this method with different significancies<br />

in the strength if the signal ([242], [84], [161], [45]) at all energy bands.<br />

Since the XMS sample has identification completenesses over 95% in the Soft<br />

band and about ∼85% in the Hard band, it would allow us to test these results<br />

using its sources.<br />

This chapter is structured as follows: in sections 4.2 and 4.3 we will study the<br />

cosmic variance and the angular clustering of the AXIS sources, while in section<br />

4.4, we will investigate the spatial clustering of XMS sources. Throughout<br />

this chapter we will assume H 0 = 70 km s −1 Mpc −1 and a flat Universe framework<br />

with Ω M = 0.3 and Ω Λ = 0.7 ([213]).<br />

4.2 Cosmic variance<br />

The simplest test for source clustering is to compare the actual number of sources<br />

detected in each field N k , with the number of expected sources λ k from the best<br />

fit log N − log S (see Chapter 3) and the sky area of each individual field Ω k .<br />

This procedure is similar to the traditional counts-in-cells method. In principle,<br />

if one field (or a few fields) were looking through strong cosmic structures<br />

the number of sources detected should be significantly different from the expected<br />

number of sources for a random, uniform distribution as measured by<br />

the overall log N − log S.<br />

The statistical tools we have used to measure the deviation from such a random<br />

uniform distribution are the cumulative Poisson distributions:<br />

P λk (≥ N k ) =<br />

∞<br />

∑<br />

l=N k<br />

P λk (l); N k > λ k (4.2.1)<br />

P λk (≤ N k ) =<br />

N k<br />

∑ P λk (l); N k < λ k (4.2.2)<br />

l=0<br />

where P λ (l) is the Poisson probability of detecting l sources when the expected<br />

number of sources is λ. The cumulative distributions above give the probability<br />

84


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

of finding ≥ N k (eq. 4.2.1) or ≤ N k (eq. 4.2.2) sources when the expected number<br />

from the source counts is λ k . This method is similar to the one used in [37],<br />

which uses P λ (l) instead of the cumulative probabilities, being our approach a<br />

more conservative estimate to determine the likelihood of finding a number of<br />

sources in a field which is different from the expected value.<br />

Then, the maximum likelihood for the whole sample is:<br />

L ′ = −2 ∑ ln P λk (≥ N k ) − 2 ∑ ln P (≤ N λk ′ k ′) (4.2.3)<br />

k<br />

k ′<br />

where k runs over the fields for which N k > λ k and k ′ runs over the fields for<br />

which N ′ k < λ′ k .<br />

We have compared the observed L’ values in each band with 10000 simulated<br />

values using the values of λ k and Poisson statistics. The number of simulations<br />

with likelihood values above the observed ones were 1388, 4580, 778 and 4958<br />

for the Soft, Hard, XID and ultrahard bands, respectively. As it can be seen, we<br />

found nothing significant in both the hard and ultrahard bands. However, in<br />

the Soft and XID bands we have found some deviations below or at about the<br />

90% significance level. These results are not conclusive at all but indicate that,<br />

at least in the Soft and XID bands, a more detailed study of clustering may be<br />

worthwhile.<br />

4.3 Angular correlation function of AXIS sources<br />

If cosmic structure is present in all (or most) fields, a test that measures significant<br />

deviations from the mean number of sources in each field from some<br />

overall average will not provide any significant results, as seen in section 4.2.<br />

Instead, we should look for evidence of sources tending to appear together in<br />

the sky with respect to an unclustered source distribution.<br />

4.3.1 Method<br />

According to [174], the angular two-point correlation function of objects distribution<br />

w(θ) determines the joint probability of finding two objects in two small<br />

angular regions δΩ 1 and δΩ 2 separated at an angular distance θ with respect to<br />

85


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

that expected from a random distribution:<br />

δP = n 2 δΩ 1 δΩ 2 [1 + w(θ)] (4.3.1)<br />

where n is the mean sky density of objects in the sky. If w(θ) = 0 the distribution<br />

is homogeneous.<br />

The angular separation is a projection in the sky of the real spatial separation<br />

between two sources at different redshifts thus hiding the underlying spatial<br />

clustering (which needs accurate redshifts for a very high fraction of the sources<br />

to be properly measured). The spatial clustering, however, can be estimated<br />

by deprojecting the computed angular correlation function assuming a given<br />

redshift distribution for the sources (which can be either empirically measured<br />

or derived from a luminosity function model) via the inversion of Limber’s<br />

equation (see section 4.3.5). The angular correlation function is, nevertheless,<br />

a powerful approach given the large sizes of the two-dimensional extragalactic<br />

surveys.<br />

There are several methods to measure the angular correlation function, most<br />

of them based on the idea of looking for an excess number of source pairs at a<br />

given angular separation θ with respect to a simulated random homogeneous<br />

sample ([131], [59]). Among all the proposed estimators for w(θ), we have chosen<br />

the one proposed by [59]:<br />

w(θ) = f DD<br />

DR − 1 (4.3.2)<br />

where DD is the number of actual pairs of sources with an angular separation<br />

θ, and DR is the number of pairs having one real and one simulated source at<br />

the same angular separation, while f is a normalisation constant that takes into<br />

account the different numbers of real and simulated sources. The error bars are<br />

given by ([174]):<br />

√<br />

1 + w(θ)<br />

∆w(θ) =<br />

(4.3.3)<br />

DR<br />

To produce the random source sample we tried to follow as closely as possible<br />

the real distribution of the source detection sensitivity of the survey so that we<br />

can provide an accurate random sky against which to judge the presence or absence<br />

of significative overdensities at different angular separations. To do this,<br />

86


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

we have used bootstrap simulations in the following manner. First, we have<br />

formed a pool of real sources with detection likelihoods higher than 15 (see section<br />

2.1) in the band under study, regardless of having their count rates above or<br />

below the sensitivity map of the corresponding field at the source position (see<br />

Appendix A). Then, keeping the number of simulated sources equal to the real<br />

number of sources in each field N k , we have randomly extracted sources from<br />

this pool keeping their count rates and distances to the optical axis of the X-ray<br />

telescope but randomising the azimuthal angle around it. If the source had a<br />

count rate above the sensitivity map of the field under consideration at its new<br />

position (X, Y), the source was kept in the simulated sample. Otherwise, a new<br />

source is randomly extracted from the pool until N k valid simulated sources are<br />

found. This way, we are able to mimic the decline of the source detection sensitivity<br />

with the off-axis angle in the simulated sample. The number of random<br />

samples used in this calculations N sim have been chosen so as to give a total of<br />

about one million simulated sources in each band.<br />

With this in mind, the normalisation constant f in equation 4.3.2 is:<br />

f = ∑ k N k (N k − 1)<br />

2N sim ∑ k N 2 k<br />

(4.3.4)<br />

where N k is the number of sources in field k and N sim is the number of simulations.<br />

4.3.2 Results<br />

We have applied the method described in section 4.3.1 to the AXIS sample for<br />

each energy band discussed in this thesis: Soft (0.5-2 keV), Hard (2-10 keV), XID<br />

(0.5-4.5 keV) and Ultrahard (4.5-7.5 keV).<br />

The values of w(θ) calculated using equation 4.3.2 in each band are shown in<br />

Figs. 4.1 and 4.2. These values are corrected from the so-called integral constraint,<br />

which is a bias in the correlation function that occurs when a positive<br />

correlation is present at angular scales comparable to the individual field size.<br />

When this happens, the estimate of the mean surface density of objects from<br />

the survey is too high, thus generating a negative bias in the angular correla-<br />

87


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Figure 4.1 w(θ) versus θ (Top panel: Soft, Bottom panel: Hard). Solid dots are the integral<br />

constraint-corrected observed values. Solid triangles are the estimated values of the<br />

integral constraint, displaced to the left for clarity. The solid line is the best χ 2 fit to our<br />

data.<br />

88


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Figure 4.2 w(θ) versus θ (Top panel: XID, Bottom panel: Ultrahard). Solid dots are the<br />

integral constraint-corrected observed values. Solid triangles are the estimated values<br />

of the integral constraint, displaced to the left for clarity. The solid line is the best χ 2 fit<br />

to our data. The panels have different scales on their Y-axis.<br />

89


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

tion function [17]. Formally, the integral constraint can be estimated by:<br />

W =<br />

∫ ∫<br />

dΩ1 dΩ 1 w(θ)<br />

∫ ∫<br />

dΩ1 dΩ 2<br />

(4.3.5)<br />

where the integral is carried out over the whole area of the survey. However,<br />

given the complicated dependence of the sensitivity with the area of our survey,<br />

we decided to evaluate the integral constraint empirically. To do that, we<br />

calculated the angular correlation function that would have been detected in<br />

the absence of correlation w IC (θ) via the average of N sim simulated realisations<br />

of w(θ), where the real data were replaced by random samples simulated independently<br />

using the method described in section 4.3.1. The w IC (θ) points at<br />

each angular scale are overplotted as triangles in Figs. 4.1 and 4.2, and have<br />

been used to increase the corresponding w(θ), convolving the error bars using<br />

Gaussian statistics.<br />

Our first results to the full sample resulted in a ∼3-σ significant bump at scales<br />

about 200 arcsec, which was also present in the simulations but at lower significance.<br />

The origin of this bump turned out to be in only two fields: HD 111812<br />

and HD 117555. Both fields have stellar clusters in their field of view ([60])<br />

and, since we are only interested in the extragalactic large-scale structure, we<br />

have opted to exclude these two fields from all subsequent angular clustering<br />

analysis.<br />

The observed w(θ) values can be fit to a power law model in the form:<br />

w(θ) =<br />

( θ<br />

θ 0<br />

) −γ<br />

(4.3.6)<br />

The best χ 2 fits to this model are overplotted in Figs. 4.1 and 4.2. The best<br />

fit parameters are listed in table 4.1 along with the significance of the detection<br />

from an F-test, comparing the χ 2 value of a power law model a w(θ) = 0,<br />

no-clustering model. The F-test suggests significant correlation at the ∼99%<br />

confidence level in the Soft and XID bands, whereas no significant clustering<br />

detection is seen the the Hard and Ultrahard bands.<br />

In [72], they find significant angular correlation in the XMM-LSS sample ([183])<br />

in the Soft band with a similar slope γ = 1.2 ± 0.2, and a lower correlation<br />

length θ 0 = 7 ± 3 arcsec, but compatible with ours within ∼1-σ. In [229] find<br />

γ = 0.7 ± 0.3 and θ 0 = 4 ± 3 arcsec, again a lower correlation length than in our<br />

90


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Table 4.1 Information on the simulations for the angular correlation function together<br />

with fit results: Band is the band where the sources were selected, N pool is the number<br />

of sources in the pool used for the bootstrap simulation (see section 4.3.1), N is number<br />

of sources selected in the real sample, N sim is the number of simulations, χ 2 0<br />

is the value<br />

of the χ 2 fit when using a no clustering model, χ 2 is the best fit value for a power-law<br />

model (see Eq. 4.3.6) with the best fit parameters and their 1-σ uncertainty intervals<br />

given by the last two columns, N bin is the number of bins used in the fit and P(F) is the<br />

F-test probability that the improvement in the fit is significant (the smaller the P(F) the<br />

higher the significance). The second row in each band corresponds to the best fit fixing<br />

γ to the “canonical” value 0.8.<br />

Band N pool N N sim χ 2 0<br />

χ 2 N bin P(F) θ 0 (“) γ<br />

Soft 1177 1131 1000 17.10 5.10 10 0.0079 19 +7<br />

−8<br />

1.2 +0.3<br />

−0.2<br />

7.80 10 0.0096 6 +2<br />

−2<br />

≡ 0.8<br />

Hard 415 351 2500 8.47 7.33 10 0.5622 12 +20<br />

−12<br />

1.1 +2.8<br />

−2.3<br />

7.47 10 0.3017 4 +5<br />

−4<br />

≡ 0.8<br />

XID 1301 1218 1000 16.00 5.30 10 0.0120 19 +7<br />

−8<br />

1.3 +0.4<br />

−0.3<br />

8.80 10 0.0238 4 +2<br />

−2<br />

≡ 0.8<br />

Ultrahard 88 77 10000 1.97 1.87 10 0.8119 0.7 a 0.22 a<br />

1.90 10 0.5759 8 +26<br />

−8<br />

≡ 0.8<br />

HR 250 225 10000 3.60 1.10 10 0.0087 42 +8<br />

−12<br />

3.9 +2.5<br />

−1.3<br />

3.22 10 0.3165 5 +9<br />

−5<br />

≡ 0.8<br />

Soft/3 392 380 2500 14.28 10.42 10 0.2835 35 +11<br />

−18<br />

1.7 +1.0<br />

−0.6<br />

12.53 10 0.2912 7 +7<br />

−6<br />

≡ 0.8<br />

a<br />

Parameter unconstrained by the fit<br />

91


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

results but compatible within less than 1-σ with our γ ≡ 0 result. On the other<br />

hand, in [18] is also found significant angular correlation in the Soft band but<br />

with a higher correlation length θ 0 = 10.4 ± 1.9 arcsec with the canonical slope<br />

γ = 0.8, compatible with our results at about the 2-σ level. Using the RASS-<br />

BSC survey ([232]), [3] have also found significant correlation in this band, but<br />

at much higher angular distances (∼ 8 ◦ ).<br />

In the Hard band there are no evidences of correlation in [72] (413 sources) and<br />

[187] (205 sources). However, there is very strong (> 4σ) clustering detected by<br />

[17] (171 sources) with γ = 1.2 ± 0.3 and θ 0 = 49 +16<br />

−25<br />

arcsec. There is also clustering<br />

detected at similar significance by [241] with γ ≡ 0.8 and θ 0 = 40 ± 11 arcsec<br />

(278 sources). These results are somewhat surprising, since the sample in [17]<br />

is less than a half of the sample in [72] and similar to the one in [187], while<br />

the sample in [241] is also smaller than the [72] sample. Our angular correlation<br />

length is compatible with [17] and [187] results within 1-σ, but not with the<br />

correlation length obtained by [241]. Nevertheless, the F-test on our correlation<br />

function and a Poissonian statistics test (see section 4.3.3) indicates that no significant<br />

clustering have been detected in this band. Our best fit results are again<br />

intermediate between the results obtained by [17] and [241] on one hand, and<br />

[72] and [187] on the other.<br />

4.3.3 Poisson statistics test<br />

Each bin in w(θ) contains information from many pairs, probably involving the<br />

same real and simulated sources many times. The underlying independent bins<br />

hypothesis in the χ 2 test may not be applicable in this case. To solve this problem<br />

and check the significance of the correlation we have applied a test using<br />

Poisson statistics ([37]) to compare the expected number of pairs at angular distances<br />

smaller or equal than each individual observed distance in the absence<br />

of correlation:<br />

DR(θ<br />

µ = ∑<br />

′ )<br />

f<br />

θ ′


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Figure 4.3 1 − P µ (N θ ) versus θ (Top panel: Soft, Bottom panel: Hard). The median (50%),<br />

99% and 99.9% levels from random simulations are also shown. The Y-axis scaling is<br />

different in each panel for clarity.<br />

93


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Figure 4.4 1 − P µ (N θ ) versus θ (Top panel: XID, Bottom panel: Ultrahard). The median<br />

(50%), 99% and 99.9% levels from random simulations are also shown. The Y-axis<br />

scaling is different in each panel for clarity.<br />

94


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

This way we can calculate the integrated Poisson probability of obtaining < N θ<br />

pairs for a distribution of mean µ:<br />

P µ (< N θ ) ≡<br />

N θ −1<br />

N θ −1<br />

µ<br />

∑ P µ (N) = ∑<br />

N e −µ<br />

N!<br />

N=0<br />

N=0<br />

(4.3.9)<br />

This way we can use the quantity 1 − P µ (N θ ) as an estimator of the significance<br />

of the clustering detection. A plot of 1 − P µ (N θ ) versus the angular distance θ<br />

is shown in Figures 4.3 and 4.4 for each band. We have estimated the signal<br />

in the absence of correlation in a similar way in order to quantify the integral<br />

constraint (see section 4.3.2) but replacing the real data by randomly placed<br />

sources. For each sample we have performed 10000 simulations in this way.<br />

We have thus determined the distribution of the expected value of 1 − P µ (N θ )<br />

for each angular scale if the sources were not correlated. We also show in Figures<br />

4.3 and 4.4 the mean and 99% and 99.9% percentiles of those distributions<br />

for each angular distance.<br />

The simulations show clearly that the detection significance is actually lower<br />

than the value given by purely Poisson statistics when there are more than 1000<br />

sources, as it can be seen in the Soft and XID bands where the 99% percentile<br />

is about the 3-σ level, whereas in the Hard and Ultrahard samples the 99%<br />

percentile is closer to the standard Poisson position, although showing large<br />

excursions.<br />

As we can see from the percentile values obtained from the simulations, we<br />

have found a >99% (and mostly


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Figure 4.5 Top panel: w(θ) versus θ. Symbols are the same as in Fig. 4.1. Bottom panel:<br />

1 − P µ (N θ ) versus θ for a random selection of one third of the Soft sources. The median<br />

(50%), 99% and 99.9% levels from random simulations are also shown.<br />

96


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

band sample. When we repeated both correlation tests to this sample, we found<br />

that the clear correlation signal observed in the full Soft sample dissapeared<br />

(see Table 4.1, sixth row, and Figure 4.5). We also repeated this test keeping at<br />

random a half and two thirds of the Soft sources, finding correlation at the 99%<br />

significance level only in the last case. Hence we conclude that clustering in the<br />

Hard and Ultrahard bands may be at least as strong as in the Soft band, but our<br />

relatively low number of sources prevent us from detecting it, which would<br />

be consistent with the results of [84] who did not find significant differences<br />

between the clustering properties of soft and hard X-ray selected sources in the<br />

CDF.<br />

The significant clustering found for the P µ (N θ ) test in the Soft band corresponds<br />

to a broad bump between ∼150 and ∼700 arcsecs, with secondary peaks at<br />

∼300, ∼400 and ∼650 arcsecs, the first two reaching 99.9% significancies (see<br />

Figure 4.3, top panel). These peaks also appear in the XID band (see Figure 4.4,<br />

top panel). While the increase in the number of real pairs that produce the<br />

broad bump is very gradual, the narrow peaks must correspond to significant<br />

increases in relatively narrow ranges of angular distances. To investigate<br />

whether these peaks come from a small number of fields in particular, we divided<br />

the Soft sample between the fields in the northern and southern Galactic<br />

hemispheres respectively. We performed the 1 − P µ (N θ ) test on these regions<br />

(see Figure 4.6) finding that, despite the lower significance of the results due to<br />

the lower number of sources, the broad bump and the peaks are still present<br />

in both hemispheres as well as the ∼300 arcsec peak. The peak at ∼400 arcsecs<br />

appears predominantly in the northern hemisphere while the peak at ∼650 arcsecs<br />

is much more significant in the southern hemisphere.<br />

We have repeated analysis excluding the two fields with obvious cluster emission<br />

(A1837 and A399) from the northern and southern Galactic hemispheres,<br />

respectively, but there were no significant changes in the strength or the position<br />

of the peaks. Additionally, we have performed this test one more time on a<br />

sample composed of the 7 deeper fields (exposure time over 40 ks, ∼400 sources)<br />

and in the remaining 27 fields (∼800 sources) to see whether the signal is introduced<br />

by a few deep fields. We found strong correlation only in the latter case<br />

and hence the signal must be relatively widespread over the survey area. This<br />

means that if it has cosmic origin, it must be generated by sources at z ≤ 1.5,<br />

which is the peak of the redshift distribution of medium-depth X-ray surveys<br />

([12]).<br />

97


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Figure 4.6 1 − P µ (N θ ) versus θ for the soft fields in the northern Galactic hemisphere<br />

(top panel) and southern Galactic hemisphere (bottom panel). The median (50%), 99% and<br />

99.9% levels from random simulations are also shown. The Y-axis scaling is different in<br />

each panel for clarity.<br />

98


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

4.3.4 Angular correlation of a hardness ratio selected sample<br />

The low significance or even the absence of the clustering signal detected in the<br />

hard band is a surprising result since it is known that most of the sources at the<br />

fluxes of our sources are AGN ([12]), which are known to be strongly clustered<br />

when selected both in X-rays ([161], [242]) and optically ([44]). Additionally,<br />

obscured AGN are more common in hard X-ray selected samples ([30], [50],<br />

[12]). However, we have just seen in section 4.3.3 that the significance of the<br />

clustering signal depends strongly on the density of sources in the fields under<br />

study. If the source density is too low, the signal might be dilluted and therefore<br />

we obtain results compatible with a no-clustering scenario, even if angular<br />

clustering is present.<br />

Furthermore, in [72] is shown that, in spite of the absence of significant angular<br />

clustering in their hard sample, when they selected the sources with the hardest<br />

spectrum (using their hardness ratio as a discriminator) the significance of<br />

the detection increased to 2-3σ. To select this sample, they had to reduce the<br />

significance of the detection threshold so as to increase the number of sources<br />

available to perform the analysis.<br />

Since this could be a potentially interesting result, we have performed a similar<br />

analysis defining the same hardness ratio HR = (H − S)/(H + S) where H and<br />

S are the 2-12 keV and 0.5-2 keV count rates, respectively. We have selected the<br />

sources in either the Soft and Hard band which have HR ≥ −0.2 (same as in<br />

section 3.3).<br />

The F-test indicate correlation at >99% confidence level (see Table 4.1, fifth<br />

row), but when we perform the Poisson 1 − P µ (N θ ) method no clustering is<br />

detected at any angular scale (see Figure 4.7), with the data consistent with<br />

the median of the random simulations. We cannot thus confirm the results obtained<br />

in [72], in spite of having a slightly larger sample of sources (250 versus<br />

209 in [72]). Moreover, this would be in agreement with the results of [242],<br />

who could not find any dependence on the shape of the source spectra in their<br />

spatial clustering analysis.<br />

4.3.5 Inversion of Limber’s equation<br />

The two-dimensional angular correlation function is a projection in the sky of<br />

the real three-dimensional spatial correlation function ξ(r) along the line of<br />

99


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Figure 4.7 Top panel: w(θ) versus θ. Symbols are the same as in Fig. 4.1. Bottom panel:<br />

1 − P µ (N θ ) versus θ for a HR selected sample (HR ≥ −0.2). The median (50%), 99%<br />

and 99.9% levels from random simulations are also shown.<br />

100


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

sight, where r is the physical separation between sources.<br />

We can model the spatial correlation function as ([53]):<br />

ξ(r, z) =<br />

( r<br />

r 0<br />

) −γ<br />

(1 + z) −(3+ɛ) (4.3.10)<br />

where ɛ parametrizes the type of clustering evolution. For instance, if ɛ =<br />

γ − 3, the clustering is constant in comoving coordinates, which means that the<br />

amplitude of the correlation function remains fixed with redshift in comoving<br />

coordinates as the pair of sources expands together with the Universe. On the<br />

other hand, if ɛ = −3 the clustering is constant in physical coordinates ([53]).<br />

The angular amplitude θ 0 can be related to the spatial amplitude r 0 by inverting<br />

Limber’s integral equation ([175]). In the case of a spatially flat Universe,<br />

Limber’s equation can be expressed as ([18]):<br />

w(θ) = 2<br />

∫ ∞<br />

0<br />

∫ ∞<br />

0 D4 C φ2 (D c )ξ(r, z)dD c du<br />

(∫ ∞<br />

0 D2 c φ(D c)dD c<br />

) 2<br />

(4.3.11)<br />

where φ(D c ) is the selection function (the probability that a source at a comoving<br />

distance D c is detected in the survey). The comoving distance D c is related<br />

to the redshift through ([110]):<br />

D c (z) =<br />

c H 0<br />

∫ z<br />

0<br />

dz ′<br />

E(z ′ )<br />

(4.3.12)<br />

with:<br />

E(z) =<br />

[<br />

Ω m (1 + z) 3 + Ω Λ<br />

] 1/2<br />

(4.3.13)<br />

in a spatially flat Universe (Ω k = 0), and H 0 being the Hubble constant ([175],<br />

[110]).<br />

The number of objects in a survey that subtend a solid angle Ω S in the sky<br />

within a redshift shell dz is:<br />

( )<br />

dN<br />

c<br />

dz = Ω SDc 2 φ(D c ) E −1 (z). (4.3.14)<br />

H 0<br />

101


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Combining these equations, Limber’s equation becomes:<br />

w(θ) = 2 H 0<br />

c<br />

∫ ∞<br />

0<br />

( 1<br />

N<br />

)<br />

dN 2 ∫ ∞<br />

E(z)dz ξ(r, z)du. (4.3.15)<br />

dz<br />

0<br />

Considering that the physical separation between two sources that project an<br />

angle θ in the sky can be written as:<br />

r ≃ 1<br />

1 + z (u2 + x 2 θ 2 ) 1/2 (4.3.16)<br />

under the small angle approximation assumption, if we now combine equations<br />

4.3.10 and 4.3.15 we obtain:<br />

where<br />

θ γ−1<br />

0<br />

= H γ<br />

(<br />

r<br />

γ<br />

0 H 0<br />

c<br />

) ∫ ∞<br />

( 1<br />

N<br />

0<br />

)<br />

dN 2<br />

E(z)(1 + z) −3−ɛ+γ<br />

dz Dc<br />

γ−1 (z)<br />

dz (4.3.17)<br />

with Γ being the gamma function.<br />

H γ = Γ( 2 1 γ−1<br />

)Γ(<br />

2 )<br />

Γ( γ 2 ) (4.3.18)<br />

We can now invert this equation and obtain r 0 given an angular amplitude θ 0 ,<br />

but we also need to determine the source redshift distribution dN/dz. Since<br />

we have no redshift information of the AXIS sources we can approximately<br />

measure dN/dz using an estimate of the AGN luminosity function. We can<br />

do that via the equation 4.3.14 writing the selection function (degradation of<br />

sampling as a function of the distance in flux-limited surveys) as:<br />

φ(D c ) =<br />

∫ ∞<br />

L min (z)<br />

Φ(L x , z)dL x (4.3.19)<br />

which depends on the evolution of the source luminosity function Φ(L x , z) but<br />

is independent of the cosmological model.<br />

We have determined the selection function (see Figure 4.8) of the AXIS sources<br />

using a luminosity function following a Luminosity-dependent Density evolution<br />

model (see Chapter 5). The best-fit model used for both the Soft and Hard<br />

bands is that obtained for the luminosity function of the XMS sources along<br />

with other deeper and shallower surveys (see Chapter 5 for details). The computed<br />

redshift distribution plotted in Figure 4.8 has been calculated using the<br />

102


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Figure 4.8 The redshift selection function for the AXIS soft (solid line) and hard (dashed<br />

line) sources.<br />

103


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Table 4.2 Spatial correlation lengths r 0 for different clustering models (ɛ) obtained from<br />

Limber’s equation. The slope γ is equivalent to that of Table 4.1<br />

Band (Ω m , Ω Λ ) γ ɛ r 0 (Mpc)<br />

Soft (0.3,0.7) 1.8 -1.2 17.6 +2.4<br />

−2.9<br />

Soft (0.3,0.7) 1.8 -3 8.9 +1.2<br />

−1.5<br />

Soft (0.3,0.7) 2.2 -0.8 11.1 +2.0<br />

−2.9<br />

Soft (0.3,0.7) 2.2 -3 6.8 +1.3<br />

−1.7<br />

Hard (0.3,0.7) 1.8 -1.2 5.5 +2.4<br />

−5.5<br />

Hard (0.3,0.7) 1.8 -3 4.8 +2.1<br />

−4.8<br />

Hard (0.3,0.7) 2.1 -0.9 2.6 +1.7<br />

−2.6<br />

Hard (0.3,0.7) 2.1 -3 2.5 +1.6<br />

−2.5<br />

effective area Ω of AXIS so that it is representative of this sample. Our results<br />

on the inversion of Limber’s equation for different clustering evolution parameters<br />

are listed in Table 4.2.<br />

In the Soft band, our r 0 results using the canonical slope γ ≡ 1.8 and a comoving<br />

clustering model ɛ = −1.2 are fully compatible with those of [18],<br />

but significantly larger than those obtained for optically selected AGN surveys<br />

r 0 ≃ 5.4 − 8.6h −1 Mpc ([3], [43], [89]) and for X-ray selected surveys ([161],<br />

r 0 ≃ 7.4h −1 Mpc). Nonetheless, if we assume that the clustering remains constant<br />

for physical coordinates ɛ = −3, our results and those of [18] are in excellent<br />

agreement with the results of [161].<br />

In the Hard band, our deprojected spatial amplitude is significantly lower than<br />

the value from [17] who obtained r 0 ≃ 12 − 19h −1 Mpc using hard sources<br />

from the XMM-Newton 2dF survey. However, they find that their correlation<br />

length are much larger (over a factor 2) than the ones provided in the literature<br />

for AGN ([45]) or 2dF ([105]) and SDSS ([29]) galaxy distributions. In fact, the<br />

r 0 values obtained by [17] can be compared instead to that of Extremely Red<br />

Objects (EROs) and luminous radio sources ([193], [168], [197]) which are in the<br />

range r 0 ≃ 12 − 15h −1 Mpc. One of the latests works in deprojection of the<br />

angular correlation function is [159], who use the sources in the XMM-Newton<br />

COSMOS field. If we compare our results in the hard band to theirs it can be<br />

seen that we are fully consistent (within the 1-σ level) with their results for the<br />

104


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

2-4.5 keV band, although we are sistematically below their values (at ∼2σ) in<br />

the 4.5-10 keV band. It must be taken into account that the values of θ 0 which<br />

we have used to invert Limber’s equation in the hard band correspond to a<br />

marginal detection (or no detection at all) of angular clustering in that band,<br />

and hence we might be underestimating that value somehow. We would like to<br />

stress that the absence of angular clustering measured in the hard band is likely<br />

to be due to the low number of pairs of sources per field compared that of the<br />

soft band (see section 4.3.3).<br />

4.4 Spatial clustering of XMS sources<br />

The angular correlation function discussed in section 4.3 is an useful tool as<br />

a first approach to the clustering analysis of a sample which lacks reliable redshift<br />

determinations or has low identification completeness. However, it is only<br />

a rough measure of the real clustering problem since it deals with projected angular<br />

distances in the sky between sources that might be at different redshifts<br />

and therefore do not form part of the same physical structure and belong to different<br />

populations. A three-dimensional deprojection of the angular clustering<br />

can be done by inverting Limber’s equation as explained in section 4.3.5 assuming<br />

a redshift distribution. Since the XMS sample (see section 2.2) is highly<br />

identified in all energy bands, we will make use of it to check whether our AGN<br />

present spatial clustering.<br />

4.4.1 Method<br />

As we explained in section 4.3.1, the most common method employed to quantify<br />

clustering is the two-point correlation function ([174]), which measures the<br />

excess probability of finding a pair of objects as a function of the physical separation<br />

between them. The spatial equivalent to equation 4.3.1 is:<br />

δP = n 2 0 δV 1δV 2 [1 + ξ(r)] (4.4.1)<br />

where n 0 is the mean density of objects and r is the comoving distance between<br />

two sources. ξ(r) is the two-point spatial correlation function. If the distribution<br />

of objects is random and homogeneous the probabilities of finding objects<br />

105


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

in the comoving volumes δV 1 and δV 2 are independent such that ξ(r) = 0. If<br />

objects are more clustered than this homogeneous distribution, then ξ(r) > 0.<br />

Analogously to the angular correlation function, the spatial one is well described<br />

by a power law model:<br />

ξ(r) =<br />

( r<br />

r 0<br />

) −γ<br />

(4.4.2)<br />

To compute the correlation function we have chosen the estimator proposed by<br />

[131], which has been demonstrated to minimise the variance:<br />

ξ(r) = N r(N r − 1) DD(r)<br />

N d (N d − 1) RR(r) − N r − 1 DR(r)<br />

N d RR(r) + 1 (4.4.3)<br />

where DD(r) is the number of actual data pairs with a separation r in redshift<br />

space. Similarly, RR(r) is the number of random pairs and DR(r) is the number<br />

of pairs having one real and one randomly simulated source at the same separation<br />

r. N d and N r are the total number of objects in the sample and the number<br />

of objects in the random sample, respectively. The uncertainties are calculated<br />

assuming that the error of the DR and RR pairs are zero, and the uncertainty of<br />

the DD pairs is Poissonian ([242]):<br />

∆ξ(r) = 1 + ξ(r) √<br />

DD(r)<br />

. (4.4.4)<br />

To calculate the comoving distances r between pairs of sources we use the standard<br />

relations described in [175] or [110]. The transverse comoving distance<br />

between the observer and an object at redshift z is:<br />

⎧<br />

⎪⎨<br />

D M =<br />

⎪⎩<br />

1<br />

D H √ sinh[ √ D<br />

Ω C<br />

Ωk<br />

k D H<br />

] ; i f Ω k > 0<br />

D C ; i f Ω k = 0<br />

1<br />

D H √ sin[√ |Ω |Ωk | k | D C<br />

D H<br />

] ; i f Ω k < 0<br />

⎫<br />

⎪⎬<br />

⎪⎭<br />

(4.4.5)<br />

where D C is:<br />

D C = D H<br />

∫ z<br />

0<br />

dz ′<br />

√<br />

ΩM (1 + z ′ ) 3 + Ω k (1 + z ′ ) 2 + Ω Λ<br />

(4.4.6)<br />

106


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

and D H is the Hubble distance:<br />

D H = c H 0<br />

. (4.4.7)<br />

If we consider two sources separated in the sky an angle θ and with transverse<br />

comoving distances to the observer D M1 and D M2 , respectively, the comoving<br />

separation of the second object measured from the first object is ([161]):<br />

r =<br />

√<br />

d 2 D 2 M1 + D2 M2 − 2dD M1D M2 cosθ (4.4.8)<br />

where ([167], [148])<br />

d =<br />

√<br />

1 + Ω k<br />

(<br />

H0 D M1<br />

c<br />

⎡ √<br />

) 2<br />

+ D M1cosθ<br />

⎣1 −<br />

D M2<br />

1 + Ω k<br />

(<br />

H0 D M2<br />

c<br />

) 2<br />

⎤<br />

⎦ . (4.4.9)<br />

In a spatially flat Universe (Ω k = 0), d = 1 and equation 4.4.8 reduces to the<br />

cosine rule for Euclidean space:<br />

r =<br />

√<br />

D 2 M1 + D2 M2 − 2D M1D M2 cosθ. (4.4.10)<br />

The procedure to generate the random sample is very similar to that we followed<br />

for the angular clustering. We formed a pool with real sources with detection<br />

likelihood higher than 15 in the band under study and a secure redshift<br />

identification. For each field, we drew a source from the pool and generated<br />

its new position at random but keeping its distance to the optical axis of the<br />

telescope (see section 4.3.1) and checked whether its count rate was above the<br />

field sensitivity map. In this case, the source is kept, along with its redshift,<br />

and a new source is drawn from the pool until the number of extracted random<br />

sources is equal to the number of real sources in the field. If the source count<br />

rate is below the sensitivity map in the new position, the source is discarded<br />

and the procedure is repeated. Following this method, the average redshift distribution<br />

of the simulated samples resembles the real distributions with great<br />

accuracy (see Fig. 4.9).<br />

107


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

Number of sources<br />

Number of sources<br />

0 10 20 30<br />

0 10 20 30<br />

0 1 2 3 4<br />

Redshift<br />

0 1 2 3 4<br />

Redshift<br />

Figure 4.9 Redshift distributions of the real XMS (solid) and the average simulated<br />

sources (dashed) in the Soft (top panel) and Hard band (bottom panel).<br />

108


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

4.4.2 Results<br />

Unlike the angular correlation function of AXIS sources (see section 4.3.2) and<br />

the predictions obtained via the inversion of Limber’s equation (see section 4.3.5)<br />

based in that results, we do not find evidence of spatial clustering in the XMS<br />

sample.<br />

Our results for the spatial correlation function are shown in Fig. 4.10. Red triangles<br />

are the estimate of the spatial correlation in the absence of clustering,<br />

calculated via simulations of a random homogeneous distribution similarly as<br />

we did when we evaluated the integral constraint in section 4.3.2. It can be<br />

seen that in both Soft and Hard bands that such a bias is almost negligible in<br />

this case and that our method obtains the expected value ξ(r) = 0 for a truly<br />

random distribution of sources. Black solid dots in Fig. 4.10 represent the ξ(r)<br />

of XMS sources at different spatial scales. It is clear that our results are fully<br />

compatible with no clustering in both bands.<br />

These results are in contradiction with a group of recent spatial clustering works<br />

([45], [161], [84], [242]). They all find strong clustering evidence fitting a power<br />

law in the form of equation 4.4.2. [45] uses more than 10000 sources from the<br />

2dF QSO Redshift (2QZ) survey, obtaining spatial correlation lengths about<br />

r 0 ≃ 5.7 ± 0.5h −1 Mpc and slopes γ ≃ 1.6 ± 0.1, quite similar to the known<br />

clustering values of local normal galaxies ([221], [188]). In [161] they use over<br />

200 AGN from the ROSAT NEP survey ([233]) to calculate the spatial correlation<br />

function. They claim to have found significative clustering (>4-σ) in the<br />

0.5-2 keV band with r 0 = 7.5 +2.7<br />

−4.2 h−1 Mpc and γ = 1.85 +1.90<br />

−0.8<br />

, also consistent<br />

the values found for normal galaxies and optically selected AGN. [84] have<br />

made use of the CDF North and South surveys (see section 2.3) gathering 240<br />

and 124 sources, respectively, finding r 0 = 8.6 ± 1.2h −1 Mpc in the CDF-S and<br />

r 0 = 4.2 ± 0.4h −1 Mpc in the CDF-N, respectively, with slopes mutally compatible:<br />

γ = 1.33 ± 0.11 and γ = 1.42 ± 0.07. [242] have also used Chandra sources<br />

from the CLASXS survey in the 2-8 keV band, which sum up to 233 sources.<br />

They found strong clustering (>6.7σ) for pairs at scales


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

ξ(r)<br />

ξ(r)<br />

−2 −1 0 1 2<br />

−2 −1 0 1 2<br />

100 1000<br />

r (Mpc)<br />

100 1000<br />

r (Mpc)<br />

Figure 4.10 ξ(r) versus r of XMS sources in the Soft (top panel) and Hard band (bottom<br />

panel). Solid dots are the observed values. Red triangles are the values obtained for a<br />

random homogenous simulated sample.<br />

110


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

131 in the Hard band) have found such strong evidences of clustering while we<br />

have not. This can be explained in terms of the dillution of the signal across<br />

the sky. The XMS survey consists of 25 fields widely spread over the sky (see<br />

section 2.2), which means that, given the amount of sources stated above, the<br />

source density is scarce (about 7 sources per field in the Soft band, only about 5<br />

in the Hard band). If all our sources were concentrated in a single deep pencilbeam<br />

field or in a few contiguous fields, we would probably have reported an<br />

spatial clustering signal similar to that found in the literature.<br />

4.5 Summary<br />

In this chapter we have studied the large scale structure of the extragalactic<br />

sources (mainly AGN) comprised in the AXIS and XMS surveys. After excluding<br />

two fields in which the presence of stellar clusters would have distorted our<br />

results, we used the AXIS sources to study the presence of cosmological structure<br />

in the X-ray sky through the cosmic variance in the number of sources per<br />

field and the distribution of the angular separations between pairs of sources.<br />

The former method is in principle more sentitive to the presence of significant<br />

overdensities in a few fields, while the latter detects any overall angular clustering<br />

of sources. No cosmic variance is detected at all in the Hard (2-10 keV) and<br />

Ultrahard (4.5-7.5 keV) bands while some signal at the ∼90% level is present in<br />

the Soft (0.5-2 keV) and XID (0.5-4.5 keV) bands, probably due to their larger<br />

number of sources.<br />

Angular clustering studies the excess probability of finding a pair of sources at<br />

a given angular distance projected in the sky. We have studied it in two ways.<br />

The first is the standard two-point angular correlation function w(θ). Using<br />

this method we have detected evidence of clustering at about the 99% level in<br />

the Soft and XID bands but not in the Hard and Ultrahard bands (see Table 4.1<br />

and Figures 4.1 and 4.2). The strength of the clustering signal we have found is<br />

intermediate between those of previous results ([72], [18], [17], [229]). A more<br />

appropiate method using Poisson statistics detects clustering at the 99-99.9%<br />

level in the Soft and XID bands, but again not in the Hard and Ultrahard bands<br />

(see Figures 4.3 and 4.4). We have checked that if we remove at random two out<br />

every three sources in the soft sample, the clustering signal disappears which<br />

implies that angular clustering may also exist in the hard band but we have<br />

111


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

failed to detect it due to the smaller number of sources in that band. Dividing<br />

the soft sample into several subsamples reveals that the signal is widespread<br />

over the sky and it is not limited to a few deep fields. This means that if this<br />

structure has a cosmic origin it must come from z ≤ 1.5, the peak of the redshift<br />

distribution of medium flux surveys ([12]). Additionally, we cannot confirm the<br />

detection of clustering signal among hard-spectrum sources reported by [72].<br />

We have used the angular amplitudes θ 0 obtained in the angular correlation<br />

function and some assumptions in the luminosity function of these sources to<br />

deproject these values by inverting Limber’s equation in order to find an estimate<br />

of their spatial clustering. The deprojected spatial amplitudes r 0 that we<br />

have found for the AXIS sources are in agreement with those calculated by [18]<br />

and [161] in the Soft band for different clustering models. In spite of the lack of<br />

signal in the hard band, the values of r 0 obtained in this band are also in excellent<br />

agreement with those of [159] using sources from the COSMOS survey, but<br />

not with the ones of [17] which are sistematically larger than any other spatial<br />

correlation lengths reported for AGN.<br />

The XMS sample has very high completenesses of secure espectroscopic redshifts<br />

in all bands. Since we can compute the real physical distances between<br />

these sources, we have studied the possible spatial clustering among them. For<br />

that, we have calculated the two-point spatial correlation function in the Soft<br />

and Hard bands, without finding evidences of large scale structure in the redshift<br />

space (see Figure 4.10). Previous works with number of sources comparable<br />

to ours ([161], [84] or [242]) have reported strong spatial clustering for<br />

AGN. This discrepancy can be explained in terms of the dillution of the clustering<br />

signal since the XMS fields are widespread over the sky leading to a low<br />

source density (about 7 and 5 sources per field in the Soft and Hard bands,<br />

respectively).<br />

112


CHAPTER 4: CLUSTERING OF AXIS AND XMS SOURCES<br />

114


Chapter 5<br />

Luminosity function of XMS sources<br />

5.1 Motivation<br />

One of the main goals of X-ray surveys is to study the comological properties<br />

of Active Galactic Nuclei (AGN) as they are strongly linked to the accretion history<br />

of the Universe and the formation and growth of the supermassive black<br />

holes that are believed to reside in the centre of all galaxies, active or not ([126],<br />

[141], [190]). The first studies in this field were constrained to the soft X-ray<br />

band (≤2 keV) which could be biased against absorbed AGN ([140], [23], [172],<br />

[157]). Therefore, hard X-ray surveys (>2 keV) are essential to describe the<br />

luminosity function of the entire AGN population, including obscured AGN<br />

which are the main contributors to the cosmic X-ray background (see [64] for<br />

a review). Moreover, the large majority of energy density generated by accretion<br />

power seems to take place in obscured AGN ([65]), as demonstrated by<br />

the integrated energy density of the cosmic X-ray background (see e.g. [42],<br />

[83]). Not taking into account the obscured AGN population could therefore<br />

bias our understanding of the cosmic evolution of the structures in the X-ray<br />

Universe. Furthermore, it has been confirmed that the fraction of absorbed<br />

AGN decreases with increasing X-ray luminosity in X-ray selected samples of<br />

AGN ([224], [215]) as well as in optically selected AGN (i.e. [212]). On the contrary,<br />

some authors claim to have found a positive evolution of the fraction of<br />

absorbed AGN with redshift ([130], [220] and, more recently, [100]) while others<br />

([224], [57], [83]) have not.<br />

Hard X-ray photons with energies


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

sources with intrinsic column densities up to N H < 10 24 cm −2 , although they<br />

are not energetic enough to escape from Compton-thick objects (N H > 1.5 ×<br />

10 24 cm −2 , for which the Compton-scattering optical depth σ T equals unity).<br />

Previous hard X-ray surveys in the 2-10 keV band ([32], [69], [79], [4]) have provided<br />

unbiased and highly complete samples of sources, many of them being<br />

obscured AGN. Notwithstanding this, an important fraction of the AGN detected<br />

in deep fields fail to provide good quality X-ray information, and their<br />

optical counterparts are far too faint to make reliable spectroscopic identifications<br />

of the totality of the samples.<br />

Previous works have studied the X-ray luminosity function (XLF, hereafter) of<br />

AGN, mainly in the 0.5-2 keV and 2-10 keV bands. For instance, [157] and [104]<br />

studied the the 0.5-2 keV XLF of AGN testing a variety of models. Their results<br />

seem to rule out both Pure Luminosity and Pure Density Evolution models in<br />

favour of a Luminosity Dependent Density Evolution model. This paper aims<br />

to confirm this results and to improve the accuracy of the determination of the<br />

model parameters by using a sample that contains ∼30% more AGN than the<br />

[157] sample and is comparable to that of [104].<br />

[211] studied the XLF of high redshift AGN in the 2-8 keV band but they did<br />

not consider the intrinsic absorption of the sources due to the limited count<br />

statistics. Other works in the 2-10 keV band such as [224] and [130] have taken<br />

into account the N H distribution of the sources when calculating the XLF, but<br />

an important fraction of the N H values were derived from the hardness ratios<br />

of the sources rather than from a spectral analysis. This may introduce a certain<br />

degree of uncertainty since a template spectrum has to be assumed to compute<br />

the N H . The XLF of very hard sources (>5 keV) is almost unexplored so far with<br />

the exception of the work by [48], who studied the de-evolved (z = 0) luminosity<br />

function of absorbed and unabsorbed AGN in the 4.5-7.5 keV band but was<br />

unable to perform detailed evolutionary studies due to the low statistics.<br />

In this chapter we use the XMM-Newton Medium Survey (XMS, [12], see section<br />

2.2), along with other highly complete deeper and shallower surveys, to<br />

compute the X-ray luminosity function (XLF, hereafter) in several energy bands.<br />

Furthermore, given the availability of high-quality X-ray spectral information<br />

in the XMS, we are able to model the intrinsic absorption of the hard sources<br />

(those detected in the 2-10 and 4.5-7.5 keV bands) as a function of the X-ray<br />

luminosity up to column densities of ∼ 10 24 cm −2 . These issues are key tools<br />

116


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

to probe the accretion history of the Universe across cosmic time. Thanks to<br />

the extremely high identification completeness of the XMS sample and the accompanying<br />

surveys we have assembled an overall sample of ∼1000 identified<br />

AGN in the 0.5-2 keV, ∼450 identified AGN in the 2-10 keV band, and ∼120<br />

identified AGN in the 4.5-7.5 keV bands, leading to one of the largest and most<br />

complete sample up to date in all three energy bands.<br />

This chapter is organized as follows: in section 5.2 we study the absorbing column<br />

density distribution of the spectroscopically identified AGN detected in<br />

hard X-rays (>2 keV) and we model the intrinsic fraction of absorbed AGN as<br />

a function of both luminosity and redshift in the 2-10 and 4.5-7.5 keV bands.<br />

In section 5.3 we discuss the method used to compute the X-ray luminosity<br />

function of AGN in several energy bands using two methods: a modified version<br />

of the classic 1/V a method to construct binned luminosity functions, and a<br />

Maximum Likelihood fit technique to an analytical model using all the sources<br />

available without binning. In section 5.4 we discuss the results obtained and we<br />

compare them with those obtained in previous works. Finally, the conclusions<br />

extracted from this work are reported in section 5.5.<br />

Throughout this chapter we have assumed a cosmological framework with<br />

H 0 = 70 km s −1 Mpc −1 , Ω M = 0.3 and Ω Λ = 0.7 ([213]).<br />

5.2 The N H function<br />

In this chapter we are pursuing to calculate the cosmological evolution of the<br />

X-ray luminosity function of all AGN (both Type-1 and Type-2) within our sample<br />

in three energy bands. This should be done by calculating the intrinsic (before<br />

absorption) luminosity and absorption (N H ) function in order to obtain<br />

purely observation-based results. Since hard X-ray photons are less affected<br />

by absorption, absorbed AGN are more likely to be detected at energies above<br />

2 keV and hence we have applied such method to the hard (2-10 keV) and ultrahard<br />

(4.5-7.5 keV) sources, for which we have detailed spectral information<br />

(photon index Γ and instrinsic absorbing column densities N H ). The ultrahard<br />

sources are particularly interesting since we would expect to detect absorbed<br />

sources more easily in this band than in the softer ones. Around 25% of the<br />

XMS sources in the Hard and Ultrahard bands have been classified as Type-2<br />

AGN, having most of them an intrinsic log N H > 22, which is the N H value<br />

117


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

log N h<br />

20 22 24<br />

AGN1<br />

AGN2<br />

42 42.5 43 43.5 44 44.5<br />

log L x (4.5−7.5 keV)<br />

Figure 5.1 Intrinsic N H versus X-ray luminosity of XMS sources detected in the Ultrahard<br />

band. Dots represent those sources identified as Type-1 AGN while the Triangles<br />

are those classified as Type-2 AGN. The dashed line at log N H = 22 marks the standard<br />

separation between absorbed and unabsorbed AGN.<br />

commonly used to separate absorbed AGN from the unabsorbed ones (see Figure<br />

5.1). An appropiate calculation of the XLF of these sources would therefore<br />

require a previous modelling of the intrinsic absorption in order to avoid possible<br />

selection effects. On the contrary, sources detected in the soft band (0.5-<br />

2 keV) are so strongly affected by absorption that we can assume that we are<br />

mainly sampling unabsorbed AGN in this band.<br />

The N H function f (L X , z; N H ) is a probability distribution function for the absorption<br />

column density as a function of the X-ray luminosity and redshift. It is<br />

measured in units of (log N H ) −1 and is normalized to unity over a defined N H<br />

region:<br />

∫ log NH max<br />

f (L X , z; N H )d log N H = 1 (5.2.1)<br />

log N Hmin<br />

We have chosen the N H limits to be log N Hmin = 20 and log N Hmax = 24 given<br />

118


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

the intrinsic N H range spanned by our joint samples (see Figure 5.2).<br />

The observed fraction of obscured AGN (those with log N H > 22) is measured<br />

by the parameter ψ, which is generally function of the both luminosity and<br />

redshift. If we compare the value of ψ at different luminosity ranges we observe<br />

that the fraction of absorbed AGN is not constant, decreasing as the luminosity<br />

increases in both the hard and ultrahard bands (see Figures 5.3 and 5.4). On the<br />

other hand, there is not significant variation in the value of ψ with redshift at<br />

a given luminosity in the ultrahard band (see Figure 5.4), which can be partly<br />

explained by the poor redshift coverage of this sample. The fraction, however,<br />

clearly increases with redshift in the case of the hard band (see Figures 5.3).<br />

We will therefore assume that the formal expression of ψ is dependent on both<br />

the X-ray luminosity and redshift and hence we have formalized it as a linear<br />

function of both log L X and z, similarly as in [130]:<br />

ψ(L X , z) = ψ 44 [(log L X − 44)β L + 1] [(z − 0.5)β z + 1] (5.2.2)<br />

where ψ 44 is the fraction of absorbed AGN at log L X = 44 and z = 0.5, and<br />

β L and β z are the slopes of the linear dependencies on luminosity and redshift,<br />

respectively.<br />

Given equation 5.2.1, the normalized N H function can be written as:<br />

f (L X , z; N H ) =<br />

{ 1−ψ(LX ,z)<br />

2<br />

; 20 ≤ log N H < 22<br />

ψ(L X ,z)<br />

2<br />

; 22 ≤ log N H ≤ 24<br />

}<br />

(5.2.3)<br />

In order to obtain the best-fit values of the free parameters ψ 44 , β L and β z we<br />

have performed a χ 2 fit on the sources of the hard and ultrahard samples, although<br />

in the latter case we have fixed β z = 0 to account for the absence of<br />

dependence on redshift observed in this band. The values thus obtained are<br />

ψ 44 = 0.41 +0.03<br />

−0.04 , β L = −0.22 +0.04<br />

−0.05 and β z = 0.57 +0.12<br />

−0.10<br />

for the hard band, and<br />

ψ 44 = 0.22 ± 0.04 and β L = −0.45 +0.20<br />

−0.25<br />

for the ultrahard band (see Table 5.2).<br />

The 1σ errors correspond to ∆χ 2 = 1. The expected fraction of absorbed AGN<br />

in the hard band at z = 0 is ψ 44 = 0.29 ± 0.05, while the expected fraction in<br />

the ultrahard band coverted to the 2-10 keV band is ψ 44 = 0.27 ± 0.08, mutually<br />

consistent with each other.<br />

If we compare these results with those of [224] and [130], we find that our bestfit<br />

value at log L 2−10 = 44 (ψ 44 = 0.41 +0.03<br />

−0.04<br />

) is in excellent agreement with the<br />

119


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

log L < 43.5<br />

43.5 < log L < 44.5<br />

log L > 44.5<br />

All Sample<br />

0 10 20 30 40 10 20 30 40 50 10 20 30 40<br />

50 100<br />

20 20.5 21 21.5 22 22.5 23 23.5 24<br />

log N h<br />

(cm −2 )<br />

log L > 43.75<br />

43 < log L < 43.75<br />

0 10 20<br />

log L < 43<br />

All Sample<br />

5 10<br />

0 2 4 6 8<br />

0 10 20 30 40<br />

20 20.5 21 21.5 22 22.5 23 23.5 24<br />

log N h<br />

(cm −2 )<br />

Figure 5.2 Observed N H distributions of the joint XMS/AMSS/CDF-S sample in the<br />

Hard band (Top panel) and XMS/HBSS sample in the Ultrahard band (Bottom panel),<br />

respectively. The distribution is shown for the full sample and in different luminosity<br />

ranges. The first bin accounts for both the sources only observed in that bin (lower<br />

step) and these sources plus the sources with log N H < 20 or uncomputed absorptions<br />

(higher step). 120


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

Fraction of Absorbed AGNs<br />

0.2 0.4 0.6<br />

41 42 43 44 45 46 47<br />

log L X (2−10 keV)<br />

Fraction of Absorbed AGNs<br />

0.2 0.4 0.6 0.8<br />

43.5 < log L 2−10<br />

< 44.5<br />

0.5 1 1.5 2 2.5 3<br />

Redshift<br />

Figure 5.3 Top panel: Fraction of absorbed AGN as a function of the 2-10 keV luminosity.<br />

Bottom panel: Fraction of absorbed AGN as a function of redshift in the luminosity range<br />

log L X = 43.5 − 44.5. Dashed lines represent the best-fit model of the N H function using<br />

N H = 10 22 cm −2 as the dividing value between obscured and unobscured AGN.<br />

121


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

Fraction of Absorbed AGNs<br />

0.1 0.2 0.3 0.4 0.5<br />

Fraction of Absorbed AGNs<br />

0 0.2 0.4 0.6<br />

42 43 44 45 46<br />

log L X (4.5−7.5 keV)<br />

43 < log L < 43.75<br />

0.2 0.4 0.6 0.8<br />

Redshift<br />

Figure 5.4 Top panel: Fraction of absorbed AGN as a function of the 4.5-7.5 keV luminosity.<br />

Bottom panel: Fraction of absorbed AGN as a function of redshift in the luminosity<br />

range log L X = 43 − 43.75. Dashed lines represent the best-fit model of the N H function<br />

using N H = 10 22 cm −2 as the dividing value between obscured and unobscured AGN.<br />

122


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

ones calculated by [224] (ψ 44 = 0.41 ± 0.03) and [130] (ψ 44 = 0.42 +0.03<br />

−0.04<br />

). [100]<br />

found a strong linear decrease from ∼80% to ∼20% in the range log L X = 42 −<br />

46. A linear fit to the data in Table 5 of [100] yields a best-fit value at log L 2−10 =<br />

44 of 0.38 ± 0.04, and a slope of −0.226 ± 0.014, also in agreement with our<br />

results in the 2-10 keV band within the error bars.<br />

The decrease in the absorbed AGN fraction at high luminosities have been reported<br />

by many authors ([224], [130], [2]). In [130] is also reported that the<br />

absorbed fraction is also dependent on redshift. Redshift dependency is also<br />

found by [220] while [57] did not find dependence on either the luminosity and<br />

redshift. In a recent paper, [100] has found a strong decrease in the fraction of<br />

absorbed AGN with X-ray luminosity and a significant increase of that fraction<br />

with redshift. The result of [100] suggests that the evolution of this fraction at a<br />

fixed luminosity of log L 2−10 = 43.75 is faster than the result obtained by [220]<br />

(probably due to the fact that the latter used the Broad Line AGN classification<br />

only, which tends to overestimate the fraction of absorbed AGN at low redshifts)<br />

and is consistent with our best fit model at a typical redshift of z = 0.5<br />

(ψ(z = 0.5) = 0.37 ± 0.07 against our prediction of ψ(z = 0.5) = 0.39 ± 0.08).<br />

The lack of dependence on redshift in our ultrahard sample, however, must be<br />

handled with caution since it comes from two similar XMM-Newton surveys<br />

only and spans a limited redshift range. As stated in [178], who found no luminosity<br />

dependence in a sample of 117 sources in the 2-10 keV band, deeper<br />

X-ray surveys are needed to take into account those sources at higher redshifts<br />

and lower luminosities in order to fully investigate the true incidence of absorption.<br />

Only combining several surveys at different depths, as we have done<br />

in the hard band, and therefore covering wider regions in the L X − z plane is<br />

possible to unveil the real dependencies.<br />

In a recent work ([48]) the dependency on luminosity of the fraction of absorbed<br />

AGN detected in the 4.5-7.5 keV band is calculated using N H = 4 × 10 21 cm −2 as<br />

the dividing value between obscured and unobscured AGN. This corresponds<br />

to A V ∼ 2 magnitudes for a given Galactic A V /N H ratio of 5.27 × 10 22 mag<br />

cm −2 which is the best dividing value between optical Type-1 AGN and optical<br />

Type-2 AGN in the HBSS sample (see section 2.3) as shown in [31].<br />

In we repeat the calculations in the ultrahard band using this N H value as a<br />

dividing line we obtain ψ 44 = 0.30 +0.04<br />

−0.05 and β L = −0.42 +0.15<br />

−0.18<br />

as the best fit parameters<br />

for the N H function. They are slightly different but consistent within<br />

123


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

the 1σ error bars to those calculated with a dividing value of N H = 10 22 cm −2 .<br />

[48] also found a fraction of absorbed AGN with L 2−10 ≥ 3 × 10 42 erg s −1 of<br />

0.57 ± 0.11 which was in excellent agreement with the results obtained by a variety<br />

of SWIFT and INTEGRAL surveys (see Table 3 in [48]). The value predicted<br />

by our ultrahard N H function, converting 2-10 keV luminosities to 4.5-7.5 keV<br />

using Γ = 1.7, is 0.55 ± 0.18 using a dividing value of N H = 4 × 10 21 cm −2 ,<br />

which fully agrees with that of [48]. Considering a dividing value of N H =<br />

10 22 cm −2 , our predicted fraction decreases to 0.42 ± 0.18 which coincides with<br />

the INTEGRAL result of [201] (0.42 ± 0.09) who also used the latter dividing N H<br />

value between absorbed and unabsorbed AGN. The predicted fraction in the<br />

hard band at L 2−10 ≥ 3 × 10 42 erg s −1 is 0.39 ± 0.09, still consistent with the<br />

value of [201].<br />

5.3 Method<br />

In this section we will calculate the X-ray luminosity function of our sources<br />

in the soft, hard and ultrahard bands. Although the backbone of this study<br />

is the XMS sample, it is imperative to combine it with other complementary<br />

shallower and deeper surveys in order to cover a wider L x − z range. Shallower,<br />

wider area surveys will provide significant numbers of bright sources at low<br />

redshifts while deep pencil-beam surveys will probe fainter sources at greater<br />

distances. The accompanying surveys used in this chapter are summarized in<br />

Table 5.1 and they have been described in section 2.3.<br />

In order to have an optimal L X − z plane coverage (see Figure 5.5) we will constrain<br />

our analysis to those sources identified as AGN with redshifts in the<br />

range 0.01 < z < 3 in the soft and hard bands (spanning up to seven and<br />

six orders of magnitude in luminosity, respectively) and 0.01 < z < 2 in the<br />

ultrahard band (spanning up to four orders of magnitude in luminosity). The<br />

total sky area covered by these surveys is shown in Figure 5.6. For the purposes<br />

of this work and to avoid errors and biases caused by further classification, we<br />

have used the entire AGN population available (within the redshift limits stated<br />

above) irrespective of whether they were identified as Type-1 AGN or Type-2<br />

AGN.<br />

The differential XLF measures the number of AGN per unit of comoving vol-<br />

124


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

Table 5.1 Summary of surveys used in this work, along with their flux limits, sky coverage,<br />

identification completeness and the number of identified AGN for each energy<br />

band.<br />

Soft (0.5-2 keV)<br />

Survey Flux limit (erg cm −2 s −1 ) Area (deg 2 ) Completeness (%) N AGN<br />

RBS 2.5×10 −12 20300 100 310<br />

RIXOS8 8.4×10 −14 4.44 100 40<br />

RIXOS3 3.0×10 −14 15.77 94 182<br />

XMS 1.5×10 −14 3.33 96 178<br />

RDS 5.5×10 −15 0.30 92 39<br />

CDF-S 5.5×10 −17 0.12 100 a 226<br />

Hard (2-10 keV)<br />

Survey Flux limit (erg cm −2 s −1 ) Area (deg 2 ) Completeness (%) N AGN<br />

AMSS 3.0×10 −13 68 99 79<br />

XMS 3.3×10 −14 3.33 84 120<br />

CDF-S 4.5×10 −16 0.11 100 a 236<br />

Ultrahard (4.5-7.5 keV)<br />

Survey Flux limit (erg cm −2 s −1 ) Area (deg 2 ) Completeness (%) N AGN<br />

HBSS 7.0×10 −14 25.17 97 62<br />

XMS 6.8×10 −15 3.33 86 57<br />

a Including photometric redshifts (see section 2.3).<br />

125


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

log L x (0.5−2 keV)<br />

40 42 44 46<br />

XMS<br />

CDFS<br />

RIXOS<br />

RDS<br />

RBS<br />

0 1 2 3 4<br />

z<br />

log L x (2−10 keV)<br />

40 42 44 46<br />

XMS<br />

CDF−S<br />

AMSS<br />

0 1 2 3 4<br />

z<br />

log L x (4.5−7.5 keV)<br />

42 43 44 45<br />

XMS<br />

HBSS<br />

0 0.5 1 1.5<br />

z<br />

Figure 5.5 Luminosity-redshift plane of different X-ray surveys in the Soft (top panel),<br />

Hard (Center panel) and Ultrahard (Bottom panel) bands.<br />

126


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

Sky Area (deg 2 )<br />

10 −6 10 −5 10 −4 10 −3 0.01 0.1 1 10 100 1000 10 4<br />

Overall<br />

CDF−S<br />

XMS<br />

RIXOS<br />

RDS<br />

RBS<br />

10 −17 10 −16 10 −15 10 −14 10 −13 10 −12 10 −11<br />

Flux (cgs) (0.5−2 keV)<br />

Sky Area (deg 2 )<br />

10 −6 10 −5 10 −4 10 −3 0.01 0.1 1 10 100<br />

Overall<br />

CDF−S<br />

XMS<br />

AMSS<br />

10 −17 10 −16 10 −15 10 −14 10 −13 10 −12 10 −11<br />

Flux (cgs) (2−10 keV)<br />

Sky Area (deg 2 )<br />

0.01 0.1 1 10<br />

Overall<br />

HBS<br />

XMS<br />

10 −14 10 −13 10 −12<br />

Flux (cgs) (4.5−7.5 keV)<br />

Figure 5.6 Sky area of the survey as a function of flux in the Soft (top panel), Hard (Center<br />

panel) and Ultrahard (Bottom panel) bands.<br />

127


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

ume V and log L X , and is function of both luminosity and redshift:<br />

dΦ(L X , z)<br />

d log L X<br />

= d2 N(L X , z)<br />

dVd log L X<br />

(5.3.1)<br />

and it is assumed that is a continous function over the luminosity and redshift<br />

ranges over which it is defined.<br />

As a first approach we will estimate the XLF in fixed luminosity and redshift<br />

bins (section 5.3.1) which will allow us to have a general overview of the overall<br />

XLF behavior. In section 5.3.2 we will express the XLF in terms of an analytical<br />

model over which we will perform a Maximum Likelihood (ML) fit, using<br />

the full information available from each single source and thus avoiding biases<br />

coming from finite bin widths and the imperfect sampling of the L X − z plane.<br />

5.3.1 Binned luminosity function<br />

There are a variety of methods to estimate the binned XLF. The classical approach<br />

is the 1/V a method ([203]), which was later generalized by [9] for samples<br />

with multiple flux limits. This method has been widely used to compute<br />

the binned XLF in evolution studies of flux-limited samples (i.e. [140], [62])<br />

but it can lead to systematic errors, especially when dealing with sources very<br />

close to the flux limit. A number of techniques have been proposed to solve<br />

this problem ([129], [170], [158]). In this work, we have used the alternative<br />

method proposed by [170] to calculate the differential binned luminosity function,<br />

which avoids most of the biases that accompany the classic 1/V a method.<br />

The differential XLF can be approximated by:<br />

dΦ(L X , z)<br />

d log L X<br />

≈<br />

∫ Lmax<br />

L min<br />

∫ zmax (L)<br />

z min<br />

N<br />

dV<br />

dz dzd log L X<br />

(5.3.2)<br />

where N is the number of objects found in a given luminosity-redshift bin ∆L∆z<br />

which is surveyed by the double integral in the denominator. The differential<br />

comoving volume has been calculated using the expression from [110]:<br />

dV<br />

dz =<br />

c ( ) 2 DL ( ) −1/2<br />

Ω M (1 + z) 3 + Ω Λ Ω(LX , z) (5.3.3)<br />

H 0 1 + z<br />

where D L is the luminosity distance and Ω(L X , z) is the solid angle subtended<br />

by the survey. Note that some portion of the ∆L∆z bin may represent objects<br />

128


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

that are fainter than the survey flux limit. The upper limit of the integral in<br />

redshift z max (L) is hence either the top of the redshift shell ∆z or the redshift at<br />

the survey flux limit for a given luminosity: z max (L) = z lim (L X , S lim ).<br />

If the sample is composed of multiple flux-limited surveys (as it happens in<br />

this work), they are added ’coherently’ as explained in [9]. This means we will<br />

assume that each object could be found in any of the survey areas for which is<br />

brighter than the corresponding flux limit.<br />

The binned XLF thus obtained for our sample presents an overall double powerlaw<br />

shape (see Figures 5.7, 5.8 and 5.9), with a steeper slope at brighter luminosities<br />

beyond a given break luminosity that takes place somewhere in the<br />

range log L X = 43 − 44. The position of this break seems to change with redshift,<br />

moving to higher values of log L X as we go deeper, thus showing clear<br />

evidences of evolution in the AGN that compose our sample. The shape at<br />

higher redshifts is less constrained due to the limited statistics. An analytical<br />

model is therefore needed to account for these changes in shape as it evolves<br />

with z, and whose parameters could be compared with previous works.<br />

5.3.2 Analytical model<br />

In the light of the results obtained in section 5.3.1, we will express the XLF as<br />

an analytic function with an smooth behavior over the range of L X and z under<br />

study in this work. The parameters of this function will describe its overall<br />

shape and how it changes with luminosity and redshift, which will express<br />

physical properties of the population under study.<br />

As shown in many works in the soft and hard bands (i.e. [157], [224], [104],<br />

[211]), the XLF seems to be best described as a double power law modified by a<br />

factor for evolution. Here we implement two different models: the simpler Pure<br />

Luminosity Evolution (PLE) model in which the X-ray luminosity evolves with<br />

redshift, and the more complicated Luminosity-dependent Density Evolution<br />

(LDDE) model in which the evolution factor not only depends on the redshift<br />

but also on the luminosity. Both models can be expressed respectively as:<br />

dΦ(L X , z)<br />

d log L X<br />

= dΦ(L X/e(z), 0)<br />

d log L X<br />

(5.3.4)<br />

129


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

for the PLE model, and:<br />

dΦ(L X , z)<br />

d log L X<br />

= dΦ(L X, 0)<br />

d log L X<br />

e(z, L X ) (5.3.5)<br />

for the LDDE model. The first factor of the second term in equation 5.3.5<br />

describes the shape of the present-day XLF for which we adopt a smoothlyconnected<br />

double power law:<br />

dΦ(L X , 0)<br />

d log L X<br />

[( ) γ1<br />

( ) γ2<br />

] −1 LX LX<br />

= A +<br />

(5.3.6)<br />

L 0 L 0<br />

where γ 1 and γ 2 are the slopes, L 0 is the value of the luminosity where the<br />

change of slope occurs, and A is the normalization.<br />

The evolution factor of the PLE model is expressed as:<br />

⎧<br />

⎫<br />

⎨ (1 + z) p 1 ; z < z c ⎬<br />

e(z) =<br />

( )<br />

⎩ (1 + z c ) p p2<br />

1 1+z<br />

1+z c<br />

; z ≥ z c<br />

⎭<br />

(5.3.7)<br />

where the parameters p 1 and p 2 account for the evolution below and above, respectively,<br />

the cut-off redshift z c . However, in the LDDE model we will assume<br />

that the cut-off redshift depends on the X-ray luminosity:<br />

⎧<br />

⎫<br />

⎨ (1 + z) p 1 ; z < z c (L X ) ⎬<br />

e(z, L X ) =<br />

(<br />

⎩ (1 + z c (L X )) p p2<br />

1 1+z<br />

1+z c (L X ))<br />

; z ≥ z c (L X ) ⎭<br />

⎧<br />

⎨<br />

z c (L X ) =<br />

⎩<br />

z ∗ c<br />

z ∗ c<br />

(<br />

LX<br />

La<br />

) α<br />

; L X ≥ L a<br />

; L X < L a<br />

⎫<br />

⎬<br />

⎭<br />

(5.3.8)<br />

(5.3.9)<br />

where α measures the strength of the dependence of z c with luminosity.<br />

5.3.3 Model fitting to soft sources<br />

We have fitted our sample in the soft band to the PLE and LDDE models described<br />

above using a Maximum Likelihood (ML) method, which optimally<br />

exploits the information from each source without binning and is therefore free<br />

from all the biases commented in section 5.3.1.<br />

The ML technique we have implemented here is that of [145], in which the<br />

likelihood function is defined as the probability of observing exactly one object<br />

130


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

in the differential element dzd log L X at each (z i , L Xi ) for the N objects in the<br />

sample, multiplied by the product of the probability of observing zero objects<br />

in all other differential elements in the accessible regions of the z − L X plane.<br />

The expression to be minimized is hence:<br />

S = −2 ∑ N i=1 ln dΦ(Li X ,zi )<br />

+2 ∑ N sur<br />

j=1<br />

∫ z2<br />

z 1<br />

∫ L2<br />

L 1<br />

d log L X<br />

dΦ(L X ,z)<br />

d log L X<br />

C j (L X , z) dVj (L X ,z)<br />

dz<br />

dzd log L X<br />

(5.3.10)<br />

Here, the index i runs over all the sources present in the sample whereas the index<br />

j runs over all the surveys that compose the sample. The factor C j (L X , z) is a<br />

correction for spectroscopic incompleteness in the j-th survey, and is function of<br />

the X-ray flux. The integrals are calculated over the full redshift (0.01 < z < 3)<br />

and luminosity (40 log L X < 46) ranges spanned by our sample in the soft<br />

band taking into account the flux limits of the different surveys that compose<br />

the sample.<br />

The expression for S is minimized using the MINUIT software package ([116])<br />

from the CERN Program Library. 1σ errors for each parameter are calculated<br />

by fixing the parameter of interest at different values and leaving the other parameters<br />

to float freely until ∆S=1. Since the model is rather complex and some<br />

of the parameters are unconstrained when performing the fit using the 9 free<br />

parameters, we have fixed some of them. In particular we have fixed p 2 , z c and<br />

L a to the values obtained by [104] in the 0.5-2 keV band. This left us with 6 free<br />

parameters to fix: A, γ 1 , γ 2 , L 0 , p 1 and α (see Table 5.2).<br />

5.3.4 Model fitting to hard and ultrahard sources<br />

For the sources in the hard and ultrahard bands we will make use of the information<br />

obtained in section 5.2 when performing the ML fit. Given the coverage<br />

of the three-dimensional space L X − z − N H spanned by our Hard and Ultrahard<br />

samples, we were not able to fit simultaneously the N H function and the<br />

XLF (which would lead us to a functional form with 11 free parameters). Instead,<br />

we will add the N H function to the analytic expression to be minimized<br />

but fixing their parameters, ψ 44 , β L and β z , to those obtained in section 5.2.<br />

131


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

Table 5.2 Parameters of the X-ray luminosity function.<br />

Soft (0.5-2 keV) Hard (2-10 keV) Ultrahard (4.5-7.5 keV)<br />

PLE LDDE PLE LDDE PLE LDDE<br />

Present day XLF parameters<br />

A a 49.27±0.58 9.83±2.94 17.96 +9.97<br />

−6.09 4.78 +0.20<br />

−0.23 10.41 +1.87<br />

−1.46 1.32±0.20<br />

log L b 0 42.90 +0.04<br />

−0.03 43.36±0.11 43.60±0.13 43.91 +0.01<br />

−0.02 43.00 +0.18<br />

−0.21 43.43 +0.23<br />

−0.32<br />

γ 1 0.36±0.01 0.51±0.02 0.81±0.06 0.96±0.02 1.14±0.01 1.28 +0.08<br />

−0.16<br />

γ 2 2.01±0.04 2.13±0.10 2.37 +0.19<br />

−0.18 2.35±0.07 2.72 +0.29<br />

−0.28 2.53±0.32<br />

Evolution parameters<br />

p 1 1.80±0.06 3.65±0.21 2.04 +0.13<br />

−0.12 4.07 +0.06<br />

−0.07 3.03 +0.17<br />

−0.22 6.46 +0.69<br />

−0.29<br />

p 2 0.00 ( f ixed) -1.5 ( f ixed) 0.00 ( f ixed) -1.5 ( f ixed) 0.00 ( f ixed) -1.5 ( f ixed)<br />

z c 1.7 ( f ixed) 1.42 ( f ixed) 1.9 ( f ixed) 1.9 ( f ixed) 1.9 ( f ixed) 1.9 ( f ixed)<br />

log L a<br />

b ... 44.6 ( f ixed) ... 44.6 ( f ixed) ... 44.6 ( f ixed)<br />

α ... 0.108±0.005 ... 0.245±0.003 ... 0.245 ( f ixed)<br />

N H function parameters<br />

ψ 44 ... ... 0.41 +0.03<br />

−0.04 0.41 +0.03<br />

−0.04 0.22±0.04 0.22±0.04<br />

β L ... ... -0.22 +0.04<br />

−0.05 -0.22 +0.04<br />

−0.05 -0.45 +0.20<br />

−0.25 -0.45 +0.20<br />

−0.25<br />

β z ... ... 0.57 +0.12<br />

−0.10 0.57 +0.12<br />

−0.10 0.00 ( f ixed) 0.00 ( f ixed)<br />

P 2DKS (L X , z) c 2×10 −4 0.21 3×10 −4 0.18 0.28 0.84<br />

a In units of 10 −6 h 3 70 Mpc−3 .<br />

b In units of h −2<br />

70 erg s−1 .<br />

c 2D K-S test probability.<br />

Taking this into account, the expression for S is:<br />

S = −2 ∑i=1 N ln dΦ(Li X ,zi )<br />

d log L<br />

f (L i X X , zi ; NH i )<br />

+2 ∑ N sur<br />

j=1<br />

∫ z2<br />

z 1<br />

∫ L2<br />

L 1<br />

∫ NH2<br />

N H1<br />

×C j (L X , z, N H ) dVj (L X ,z,N H )<br />

f (L X , z; N H ) dΦ(L X,z)<br />

d log L X<br />

(5.3.11)<br />

dz<br />

dzd log L X d log N H<br />

Again, the integrals are calculated over the full L X − z − N H space in the ranges<br />

0.01 < z < 2, 20 < log N H < 24, 40 log L X(2−10) 46 (in the hard band)<br />

and 42 log L X(4.5−7.5) < 45 (in the ultrahard band). As we did in the previous<br />

section, we fixed the XLF parameters p 2 , z c and L a to those obtained by [224]<br />

for sources detected in the hard band and, in addition, given the poor coverage<br />

of the L x − z plane achieved by the ultrahard sample, we have also fixed the<br />

strength of the dependence of z c on luminosity α to the value we obtained when<br />

fitting the hard sample.<br />

132


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

5.4 Results<br />

5.4.1 Discussion<br />

The best-fit parameters of the XLF calculated by the ML method explained<br />

above are summarized in Table 5.2. Best fit results to the PLE and LDDE models<br />

along with data points from the binned XLF are shown in Figures 5.7, 5.8 and<br />

5.9.<br />

It has been shown in different works that the LDDE model for the XLF provides<br />

the best framework that describes the evolutionary properties of AGN,<br />

both in soft ([157], [104]) and hard X-rays ([224], [130], [211]), as well as in<br />

the optical range ([22]). We show in this paper that this analytical model also<br />

works for AGN detected at harder X-rays such as the ones comprised in the<br />

joint XMS/HBSS sample.<br />

At first sight our best-fit models in the soft band (0.5-2 keV) reveal strong differences<br />

between the PLE and LDDE models. Although the PLE best-fit seems<br />

to be a fair description of the XLF at low redshifts (z < 1) it clearly fails to reproduce<br />

the behaviour at high redshifts, where the LDDE model matches better<br />

with our binned XLF data points.<br />

The best-fit parameters obtained by our ML technique slightly differ from the<br />

soft XLF models reported in [157] and [104] using a variety of ROSAT, XMM-<br />

Newton and Chandra surveys. For instance, our best-fit PLE model requires less<br />

evolution (p 1 parameter) than that of [104] PLE model and the present-day luminosity<br />

break L 0 is displaced at fainter luminosities in our model (but consistent<br />

within 2σ with that of [104]).<br />

The LDDE model provides a better overall description of the XLF. However, our<br />

present day XLF power law shape has flatter slopes and the luminosity break L 0<br />

is also displaced at fainter luminosities than in [104] by more than 2σ, although<br />

it is consistent with the LDDE models of [157] within the error bars. Our best-fit<br />

evolution parameter p 1 reveals less evolution below the cut-off redshift (fixed<br />

to the value obtained by [104]: z c = 1.42) and a much weaker dependence of z c<br />

on luminosity α than in these works.<br />

Similarly, in the hard band (2-10 keV) the LDDE model outperforms the PLE<br />

model at z > 0.5, failing the latter to describe the evolution of the binned data<br />

points at the faint end of the XLF underestimating the data at log L X > 44 by<br />

133


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

d φ / d log L x<br />

10 −11 10 −10 10 −9 10 −8 10 −7 10 −6 10 −5 10 −4 10 −3 0.01<br />

0.01


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

a factor of several. Comparing our results in the hard band (2-10 keV) with<br />

previous major works in this band ([224], [130], [211]) we find an overall good<br />

agreement between our work and theirs, although with some differences in<br />

detail. The general shape of the present day XLF, described by the smooth double<br />

power law, is similar to that reported in other works within the error bars.<br />

The evolution below the redshift cut-off p 1 is also consistent with the results<br />

of [224] and [130] (both using samples corrected by the intrinsic absorption of<br />

their sources), and [211] well within the 1σ confidence level. The stronger difference<br />

between these models arises with the strength of the dependence of the<br />

cut-off redshift z c on luminosity, measured by the parameter α. The values of<br />

α obtained by [224] and [211] are consistent with each other while the ones calculated<br />

by [130] (α ∼ 0.2) and in this work (α ∼ 0.25) are sistematically lower.<br />

Overall, our best-fit LDDE parameters are more constrained than in the other<br />

works attending to the computed 1σ error bars. We have left the evolution parameter<br />

above the cut-off redshift p 2 fixed to that of [224] (the value obtained<br />

in [130] reveals less evolution beyond z c but consistent with [224] within 1σ).<br />

The model in [211] requires a much stronger evolution p 2 = −3.27 +0.31<br />

−0.34 than<br />

the others. It must be noted, nevertheless, that it is extremely difficult to properly<br />

constrain the faint end of the luminosity function given the necessity of<br />

highly complete deep pencil-beam surveys that account for the population of<br />

high-redshift low-luminosity AGN.<br />

In a recent work, [48] have computed the intrinsic present day XLF of the HBSS<br />

sample in the 4.5-7.5 keV (ultrahard) band for absorbed and unabsorbed AGN.<br />

In this work we have added the HBSS sample to ours, and performed the same<br />

analysis as in the soft and hard bands using both absorbed and unabsorbed<br />

sources. This way we have doubled the number of available AGN and improved<br />

the L X − z coverage, which is directly reflected in that the best-fit parameters<br />

are more constrained.<br />

Moreover, given the availability of detailed<br />

spectral information for the vast majority of the sources in both samples, we<br />

have carried out the XLF analysis by convolving the N H function with the XLF<br />

analytical model when performing the fit (see section 5.3.4). Like in the other<br />

energy bands, the PLE best-fit clearly underestimates the data at faint luminosities<br />

and high redshifts. From the binned data points it can be inferred that<br />

a very strong evolution p 1 is required to describe properly the behaviour of<br />

the ultrahard sources, which is achieved by the LDDE model. Our results are<br />

fully consistent with those reported in [48] albeit with smaller error bars. The<br />

135


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

d φ / d log L x<br />

10 −11 10 −10 10 −9 10 −8 10 −7 10 −6 10 −5 10 −4 10 −3 0.01<br />

0.01


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

shape of our best-fit present day XLF would correspond to the average of the<br />

absorbed and unabsorbed present day XLF calculated in [48] thus lying inside<br />

the 1σ confidendence levels spanned by their parameters. Our best-fit value for<br />

the evolution parameter p 1 = 6.46 +0.69<br />

−0.29<br />

is in excellent agreement with that of<br />

[48] (p 1 = 6.5) and also with that of [22] obtained from a selected AGN sample<br />

in the optical range from the VIMOS-VLT survey (p 1 = 6.54) and that of the<br />

bolometric quasar luminosity function of [112] (p 1 = 5.95 ± 0.23). This value<br />

represents a much stronger evolution below the cut-off redshift than that obtained<br />

by [224] (p 1 = 4.23 ± 0.39) and [130] (p 1 = 4.62 ± 0.26) in the 2-10 keV<br />

band. Some works (i.e. [157], [104], [22]) suggest that the evolution parameter<br />

p 1 might depend linearly on the X-ray luminosity, increasing as the luminosity<br />

increases. Given the lack of deep surveys, which mainly account for intrincally<br />

faint objects, in the ultrahard band, this result could be biased towards highluminosity<br />

sources and hence provides higher evolution rates below the cut-off<br />

redshift.<br />

5.4.2 Accretion history of the Universe<br />

The X-ray luminosity function of AGN, whose bolometric emission is directly<br />

linked to accretion power, constrains the history of the formation of the supermassive<br />

black holes that reside in the galactic centers along cosmic times.<br />

The comoving density of AGN as a function of redshift can be calculated straightforward<br />

from our best-fit XLF LDDE model:<br />

Φ(z) =<br />

∫ L2<br />

L 1<br />

dΦ(L X , z)<br />

d log L X<br />

d log L X (5.4.1)<br />

As we can see in Figure 5.10, the most luminous AGN are formed before the<br />

less luminous ones in all bands, although this is difficult to say in the ultrahard<br />

band due to the limited statistics. AGN with log L X > 44 reach a maximum in<br />

density at redshift ∼1.5 while fainter AGN (log L X < 44) peak at z ∼ 0.7.<br />

Similarly to the comoving density, we can derive the luminosity density as a<br />

function of redshift in all bands and, from that, calculate the accretion rate density.<br />

Here we will assume that the accretion rate onto a supermassive black<br />

hole is related to the bolometric luminosity by a constant factor ɛ which is the<br />

137


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

d φ / d log L x<br />

d φ / d log L x<br />

d φ / d log L x<br />

10 −11 10 −10 10 −9 10 −8 10 −7 10 −6 10 −5 10 −4 10 −3 0.01<br />

10 −11 10 −10 10 −9 10 −8 10 −7 10 −6 10 −5 10 −4 10 −3 0.01<br />

10 −11 10 −10 10 −9 10 −8 10 −7 10 −6 10 −5 10 −4 10 −3 0.01<br />

0.01


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

Number Density (Mpc −3 )<br />

Number Density (Mpc −3 )<br />

10 −7 10 −6 10 −5 10 −4 10 −3 0.01<br />

10 −7 10 −6 10 −5 10 −4 10 −3 0.01<br />

41.5 < log L 0.5−2 keV<br />

< 44.0<br />

44.0 < log L 0.5−2 keV<br />

< 46.5<br />

0 0.5 1 1.5 2 2.5 3<br />

z<br />

42.5 < log L 2−10 keV<br />

< 44.0<br />

44.0 < log L 2−10 keV<br />

< 46.0<br />

0 0.5 1 1.5 2 2.5 3<br />

z<br />

Number Density (Mpc −3 )<br />

10 −7 10 −6 10 −5 10 −4 10 −3 0.01<br />

43.0 < log L 4.5−7.5 keV<br />

< 44.0<br />

44.0 < log L 4.5−7.5 keV<br />

< 45.0<br />

0 0.5 1 1.5 2<br />

z<br />

Figure 5.10 Comoving density of AGN in different luminosity bins as a function of<br />

redshift in the soft (top panel), hard (Center panel) and ultrahard (Bottom panel) bands.<br />

Overplotted are the predictions from our best-fit LDDE model in the low luminosity<br />

(solid line) and high luminosity (dashed line) ranges<br />

139<br />

.


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

radiative efficiency of the accretion flow ([144], [130]):<br />

L bol =<br />

ɛ<br />

1 − ɛ Ṁaccc 2 (5.4.2)<br />

We can derive the bolometric luminosities by means of a bolometric correction<br />

factor K simply using L bol = KL X . The accretion rate density is hence:<br />

˙ρ acc (z) = 1 − ɛ<br />

ɛc 2<br />

∫ L2<br />

L 1<br />

KL X<br />

dΦ(L X , z)<br />

d log L X<br />

d log L X (5.4.3)<br />

We will assume a nominal radiative efficiency of ɛ = 0.1 ([244], [144], [16]). The<br />

values of the bolometric correction K are derived from the polynomic expressions<br />

of [144], that accounts for changes in the overall spectral energy distribution<br />

of AGN as a function of the optical luminosity. Hence, the total accreted<br />

mass onto supermassive black holes is:<br />

ρ(z) =<br />

∫ z0<br />

z<br />

˙ρ acc (z) dt dz (5.4.4)<br />

dz<br />

where we assume that the initial mass of seed black holes at z 0 is negligible with<br />

respect to the total accreted mass ([130]).<br />

In Figures 5.11 and 5.12 is clearly seen that the vast majority of the accretion<br />

rate density and the total mass density are produced by low-luminosity AGN<br />

(log L X < 44). The total accreted mass density at z = 0 obtained from our XLF<br />

model (∼ 3 × 10 5 M ⊙ Mpc −3 ) in the hard band is in good agreement with that<br />

of [144] (2.25 +0.94<br />

−0.68 × 105 M ⊙ Mpc −3 ) determined using the local black hole mass<br />

function.<br />

From these results it can be inferred that high-luminosity AGN grow and feed<br />

very efficiently in the early Universe and are fully formed at redshifts 1.5-2,<br />

whereas the low-luminosity AGN keep forming down to z ∼ 1, in agreement<br />

with an anti-hierarchical black hole growth scenario as shown by the LDDE<br />

model of the XLF.<br />

5.5 Summary<br />

We have discussed here the cosmic evolution of a sample of AGN in three X-<br />

ray bands: soft (0.5-2 keV), hard (2-10 keV) and ultrahard (4.5-7.5 keV). The<br />

140


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

Figure 5.11 Accretion rate density of matter onto supermassive black holes as a function<br />

of redshift in the soft (top panel), hard (Center panel) and ultrahard (Bottom panel)<br />

bands. Dashed lines correspond to the redshift range in which the model has been<br />

extrapolated.<br />

141


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

Figure 5.12 Total accreted mass onto supermasive black holes as a function of redshift<br />

in the soft (top panel), hard (Center panel) and ultrahard (Bottom panel) bands. Dashed<br />

lines correspond to the redshift range in which the model has been extrapolated.<br />

142


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

backbone of our sample is the XMS survey ([12]) which is a flux-limited highlycomplete<br />

sample at medium redshifts. We have combined the XMS with other<br />

shallower and wider highly complete X-ray surveys in all three bands to end up<br />

with a total sample of ∼1000, 435 and 119 AGN in the soft, hard and ultrahard<br />

bands, respectively.<br />

We have used the detailed spectral information on the sources that compose<br />

the ultrahard sample, XMS and HBSS ([48]) to model their instrinsic absorption<br />

(N H function). We find that the fraction of absorbed of AGN in the 4.5-7.5 keV<br />

band, assuming a dividing value of N H = 10 22 cm −2 , is dependent on the X-<br />

ray luminosity but not on the redshift. This could be motivated by the narrow<br />

redshift range spanned by the sample, with the bulk of the sources located at<br />

low redshifts (z < 1). The same analysis applied to the hard (2-10 keV) sources<br />

reveals dependency on both the X-ray luminosity and redshift. Our predictions<br />

on the behaviour of the fraction of absorbed AGN in this band is in excellent<br />

agreement with the results of [224] and [130] and with the more recent work<br />

of [100], who studied the evolution of the absorption properties of AGN in a<br />

compiled sample of 1290 AGN in the 2-10 keV band.<br />

We have calculated the X-ray luminosity function of AGN using two methods.<br />

First, a modified version of the 1/V a method ([203]) discussed in [170]<br />

to compute the binned XLF. Secondly, a fit to an analytic model using a Maximum<br />

Likelihood technique ([145]) that fully exploits the available information<br />

on each individual source without binning. The adopted model consists of a<br />

smoothly connected double power law that accounts for the present day (z = 0)<br />

XLF modified by an evolution factor that depends either on the redshift (PLE<br />

model) or on both redshift and luminosity (LDDE model). We have found that<br />

the LDDE model outperforms the PLE model at faint luminosities and high<br />

redshifts. The latter is ruled out as a possible description of the observations by<br />

a two-dimensional Kolmogorov-Smirnov test performed over the whole luminosity<br />

and redshift ranges spanned by our sample, whereas the LDDE model<br />

fit achieves better acceptance probabilities, especially in the ultrahard band (see<br />

Table 5.2).<br />

The best-fit XLF parameters in the soft sample show slight discrepancies in both<br />

the overall shape and evolution with respect previous works in that band ([157],<br />

[104]). In the hard band, where we have computed the intrinsic XLF taking into<br />

account the intrinsic absorption N H of each source, we are in good agreement<br />

143


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

with the results of [224], [130] or [211]. Our best-fit model shows weaker evolution<br />

of the AGN below the cut-off redshift than in these works albeit with<br />

smaller error bars. In the ultrahard band, we have also calculated the intrinsic<br />

XLF finding similar results as in [48] but with our best-fit parameters much<br />

more constrained. The results in this band show that ultrahard AGN present a<br />

significantly stronger evolution below the cut-off redshift than those detected at<br />

softer energies. This could be due to the lack of complementary deep surveys<br />

(which probe intrinsically faint sources) in the ultrahard band. If there exists<br />

a dependency of the evolution rate on the X-ray luminosity ([157], [104]) we<br />

could be biased towards high-luminosity objects thus measuring higher evolution<br />

rates. In all three bands, the high-luminosity AGN (log L X > 44) are<br />

formed before than the low-luminosity ones (log L X < 44), reaching the former<br />

a maximum in density at redshift z ∼ 1.5 whereas the comoving density of the<br />

latter peak at z ∼ 0.7.<br />

Finally, we have used our best-fit XLF to compute the accretion rate density<br />

and total accreted mass onto supermassive black holes as a function of redshift.<br />

The total black hole mass density at z = 0 predicted by our best-fit model is in<br />

agreement with the local supermassive black hole density of [144]. As shown by<br />

the XLF model, the high-luminosity AGN have a more efficient growth in the<br />

early stages of the Universe and are fully formed at z ∼ 1.5 − 2 while the less<br />

luminous AGN keep forming down to redshifts below 1. This behaviour is also<br />

found in the ultrahard sample thus confirming that the evolution of the XLF<br />

along cosmic time is not caused by changes in the environment absorption but<br />

by intrinsic variations in the accretion rate at different epochs of the Universe.<br />

144


CHAPTER 5: LUMINOSITY FUNCTION OF XMS SOURCES<br />

146


Chapter 6<br />

Summary of the results<br />

Throughout this thesis we have studied the large-scale, cosmological and evolutionary<br />

properties of Active Galactic Nuclei (AGN), detected in X-rays by<br />

the XMM-Newton observatory. The main source of energy of these sources is<br />

acrettion of matter onto a supermassive black hole (∼10 6 -10 9 M ⊙ ), which has<br />

allowed us to constrain the parameters that describe the accretion history of the<br />

Universe through cosmic time.<br />

This chapter is structured as follows: in section 6.1 we summarize the main<br />

conclusions obtained from this thesis, while in section 6.2 we discuss possible<br />

future lines of work that can be derived from the results presented here.<br />

6.1 Summary<br />

In order to carry out the investigations presented in this thesis, we have made<br />

use of two X-ray surveys at medium fluxes which were led by the X-ray Astronomy<br />

Group at the Instituto de Física de Cantabria, and in which the author of<br />

this thesis has decisively participated by reducing optical and X-ray data and<br />

identifying optical spectra:<br />

• AXIS (An XMM-Newton International Survey, [13], [38]): the backbone of<br />

the XMM-Newton Survey Science Center medium XID programme. It<br />

has a wide sky coverage (∼4.3 deg 2 ) with fluxes ranging from ∼10 −15 -<br />

10 −12 erg cm −2 s −1 in the 0.5-10 keV broadband. A total of 36 XMM-<br />

Newton pointings were selected for optical follow-up of sources at high<br />

Galactic latitudes (|b| > 20 ◦ ) with total exposure times above 15 ks. This<br />

147


CHAPTER 6: SUMMARY OF THE RESULTS<br />

survey was divided in four subsamples in the following energy ranges:<br />

soft (0.5-2 keV), hard (2-10 keV), XID (0.5-4.5 keV) and ultrahard (4.5-<br />

7.5 keV), in which we ended up with 1267, 397, 1359 and 91 X-ray sources,<br />

respectively.<br />

• XMS (XMM-Newton Medium Sensitivity Survey, [12]): a serendipitous X-<br />

ray source survey at medium fluxes built from the AXIS survey. It covers<br />

a geometric area of ∼3 deg 2 and also consists of four overlapping subsamples<br />

in the soft, hard, XID and ultrahard energy bands. The soft, hard<br />

and XID subsamples are flux limites while the ultrahard subsample is not<br />

artificially limited due to the scarcity of its sources. An extensive identification<br />

programme was carried out on the XMS survey, ending up with<br />

272 out of distinct 318 sources spectroscopically identified (86% overall,<br />

over 90% in the soft and XID bands). Most of the identified sources turned<br />

out to be Type-1 AGN (∼70% in the soft and XID bands, ∼55% in the<br />

hard and ultrahard bands) while the rest were identified as Type-2 AGN<br />

(∼20%), galaxies (∼5%) and stars (∼4%).<br />

These surveys contribute the main samples used in the investigations carried<br />

out in this thesis but, in order to gather broader statistics, we have combined<br />

them with other complementary XMM-Newton surveys as well as with surveys<br />

from other present and past X-ray observatories such as ROSAT, ASCA<br />

or Chandra. Shallower, wide area surveys provide brigh sources, while deep<br />

pencil-beam surveys probe sources at fainter fluxes.<br />

The main aims enumerated in Chapter 1 that we have addressed throughout<br />

this thesis are:<br />

1. The Cosmic X-ray Background (CXB) is a record of the accretion power<br />

history of the Universe and it is now accepted that it comes from the integrated<br />

emission of discrete extragalactic sources. Both absorbed (Type-<br />

2) and unabsorbed (Type-1) AGN are the main contributors to the CXB,<br />

with galaxy populations only contributing significantly at fainter X-ray<br />

fluxes. We have calculated the sky density or source counts of the AXIS<br />

sources, since these distributions are dependent on the cosmological properties<br />

that configure the CXB emission. The more accurate the shape of<br />

the source counts is constrained, the more precise will be the contribution<br />

from discrete sources to the CXB.<br />

148


CHAPTER 6: SUMMARY OF THE RESULTS<br />

We have therefore used our best fit logN − logS to calculate both the total<br />

resolved fraction of the Cosmic X-ray Background (including the contribution<br />

from bright sources) and the relative contribution in different flux<br />

bins to the total CXB). We have used the estimates of the average CXB intensity<br />

in the Soft and Hard bands from [160] and converted them to the<br />

XID and Ultrahard bands using a power law spectrum with Γ = 1.4. The<br />

total resolved fraction down to the lowest fluxes of our combined sample<br />

reaches 88% in the Soft band and 86% in the Hard band (where we have<br />

been able to reach deep fluxes usin pencil beam surveys [19]), but it is only<br />

60% in the XID band and 25% in the Ultrahard band. The total intensity<br />

produced extrapolating our logN − logS to zero flux does not saturate the<br />

CXB intensity. Assuming that another flux break occurs just below our detected<br />

minimum fluxes, we have estimated the minimum slope needed to<br />

reproduce the CXB with only discrete sources obtaining values of 1.85 in<br />

the Soft band and 1.84 in the Hard band. If we compare these values with<br />

those of the galaxy and AGN source counts from the Chandra Deep Field<br />

(CDF, [19]), we find that galaxies could easily provide this steepening and,<br />

among the AGN population, only the absorbed ones could do so as well.<br />

The maximum fractional contribution to the CXB in the Soft, Hard and<br />

XID bands comes from sources within a decade around ∼ 10 −14 cgs (where<br />

the break in the source counts power law approximately occurs). This is<br />

almost 50% of the total contribution in the Soft and Hard bands, which<br />

means that medium deep surveys such as AXIS are essential to understand<br />

the evolution of the X-ray emission in the Universe up to 10 keV.<br />

2. The dominant population at medium and low X-ray fluxes are AGN, which<br />

are known to cluster strongly (i.e. [242], [84], [161], [37]). Since to detect<br />

source clustering it is needed a survey that it is both deep (to achieve high<br />

angular density) and wide (to prevent a single structure to bias the overall<br />

average), the AXIS and XMS surveys are ideal to investigate the large scale<br />

structure of AGN.<br />

Angular clustering studies the excess probability of finding a pair of sources<br />

at a given angular distance projected in the sky. We have studied it in<br />

two ways. The first is the standard two-point angular correlation function<br />

w(θ). Using this method we have detected evidence of clustering at about<br />

the 99% level in the Soft and XID bands but not in the Hard and Ultrahard<br />

149


CHAPTER 6: SUMMARY OF THE RESULTS<br />

bands. The strength of the clustering signal we have found is intermediate<br />

between those of previous results ([72], [18], [17], [229]). A more appropiate<br />

method using Poisson statistics detects clustering at the 99-99.9% level<br />

in the Soft and XID bands, but again not in the Hard and Ultrahard bands.<br />

We have checked out that if we remove at random two out every three<br />

sources in the soft sample, the clustering signal disappears which implies<br />

that angular clustering may also exist in the hard band but we have failed<br />

to detect it due to the smaller number of sources in that band. Dividing the<br />

soft sample into several subsamples reveals that the signal is widespread<br />

over the sky and it is not limited to a few deep fields. This means that if<br />

this structure has a cosmic origin it must come from z ≤ 1.5, the peak of<br />

the redshift distribution of medium flux surveys ([12]). Additionally, we<br />

cannot confirm the detection of clustering signal among hard-spectrum<br />

sources reported by [72].<br />

We have used the angular amplitudes θ 0 obtained in the angular correlation<br />

function and some assumptions in the redshift distribution via the<br />

luminosity function of these sources to deproject these values by inverting<br />

Limber’s equation in order to find an estimate of their spatial clustering.<br />

The deprojected spatial amplitudes r 0 that we have found for the AXIS<br />

sources are in agreement with those calculated by [18] and [161] in the<br />

Soft band for different clustering models. In spite of the lack of signal in<br />

the hard band, the values of r 0 obtained in this band are also in excellent<br />

agreement with those of [159] using sources from the COSMOS survey,<br />

but not with the ones of [17] which are sistematically larger than any other<br />

spatial correlation lengths reported for AGN.<br />

The XMS sample has very high completeness of secure spectroscopic redshifts<br />

in all bands. Since we can compute the real physical distances between<br />

these sources, we have studied the possible spatial clustering among<br />

them. For that, we have calculated the two-point spatial correlation function<br />

in the Soft and Hard bands, without finding evidences of large scale<br />

structure in the redshift space. Previous works with number of sources<br />

comparable to ours ([161], [84] or [242]) have reported strong spatial clustering<br />

for AGN. This discrepancy can be explained in terms of the dillution<br />

of the clustering signal since the XMS fields are widespread over the sky<br />

leading to a very low source density (about 7 and 5 sources per field in the<br />

Soft and Hard bands, respectively).<br />

150


CHAPTER 6: SUMMARY OF THE RESULTS<br />

3. Other of the aims of this thesis is to study the evolutionary properties of<br />

AGN, as they are strongly linked to the accretion history of the Universe<br />

and the growth and formation of the supermassive black holes that are believed<br />

to reside in the centre of all galaxies, active or not (i.e. [126], [141],<br />

[190]). This can be achieved by computing the X-ray luminosity function<br />

(XLF) of these sources. Hard (>2 keV) X-ray surveys are indeed essential<br />

to describe the luminosity function of the whole AGN population, including<br />

obscured AGN which are the main contributors to the CXB.<br />

Hence, we have used the detailed spectral information on the sources that<br />

compose the ultrahard sample, XMS and HBSS ([48]) to model their instrinsic<br />

absorption (N H function). We find that the fraction of absorbed<br />

of AGN in the 4.5-7.5 keV band, assuming a dividing value of N H =<br />

10 22 cm −2 , is dependent on the X-ray luminosity but not on the redshift.<br />

This could be motivated by the narrow redshift range spanned by the sample,<br />

with the bulk of the sources located at low redshifts (z < 1). The same<br />

analysis applied to the hard (2-10 keV) sources reveals dependency on<br />

both the X-ray luminosity and redshift. Our predictions on the behaviour<br />

of the fraction of absorbed AGN in this band is in excellent agreement with<br />

the results of [224] and [130] and with the more recent work of [100], who<br />

studied the evolution of the absorption properties of AGN in a compiled<br />

sample of 1290 AGN in the 2-10 keV band.<br />

4. We have calculated the X-ray luminosity function of AGN using two methods.<br />

First, a modified version of the 1/V a method ([203]) discussed in [170]<br />

to compute the binned XLF. Secondly, a fit to an analytic model using a<br />

Maximum Likelihood technique ([145]) that fully exploits the available information<br />

on each individual source without binning. The adopted model<br />

consists of a smoothly connected double power law that accounts for the<br />

present day (z = 0) XLF modified by an evolution factor that depends<br />

on the redshift (PLE model) or on both redshift and luminosity (LDDE<br />

model). We have found that the LDDE model outperforms the PLE model<br />

at faint luminosities and high redshifts. In all three bands (soft, hard and<br />

ultrahard), the high-luminosity AGN (log L X > 44) are formed before than<br />

the low-luminosity ones (log L X < 44), reaching the former a maximum<br />

in density at redshift z ∼ 1.5 whereas the comoving density of the latter<br />

peak at z ∼ 0.7.<br />

151


CHAPTER 6: SUMMARY OF THE RESULTS<br />

Finally, we have used our best-fit XLF to compute the accretion rate density<br />

and total accreted mass onto supermassive black holes as a function of<br />

redshift. The total black hole mass density at z = 0 predicted by our bestfit<br />

model is in agreement with the local supermassive black hole density of<br />

[144]. As shown by the XLF model, the high-luminosity AGN have a more<br />

efficient growth in the early stages of the Universe and are fully formed at<br />

z ∼ 1.5 − 2 while the less luminous AGN keep forming down to redshifts<br />

below 1. This behaviour is also found in the ultrahard sample thus confirming<br />

that the evolution of the XLF along cosmic time is not caused by<br />

changes in the environment absorption but by intrinsic variations in the<br />

accretion rate at different epochs of the Universe.<br />

We can conclude that we have achieved to show the importance of mediumdeep<br />

serendipitous X-ray surveys for clustering and evolutionary studies. We<br />

have shown that these sources are essential to trace the large-scale structure of<br />

the cosmic web and to resolve the discrete origin of the Cosmic X-ray Background<br />

in the 0.5-10 kev energy broadband. With them, we have been able to<br />

accurately constrain the evolutionary history of AGN thus showing that obscured<br />

accretion power plays a key role in the understanding of the cosmic<br />

emission and evolution of the X-ray Universe.<br />

6.2 Future prospects<br />

Although the results presented in this thesis have contributed substantially to<br />

the undestanding of evolutionary and clustering properties of AGN, there are<br />

some aspects in which would be desirable to go deeper into.<br />

• Although we have obtained evidence that extragalactic sources at medium<br />

fluxes detected in soft X-rays cluster strongly forming cosmological largescale<br />

structures, we were not able to confirm this for sources detected in<br />

hard X-rays. We can find in the literature discrepant results on this subject,<br />

being the main reason the low density of sources per field. In an ongoing<br />

work using a large sample of serendipitous X-ray sources described in<br />

[151], we intend to compute the most precise angular correlation function<br />

of extragalactic sources detected in both soft and hard X-rays to date. Additionally,<br />

we pursue to study the possible dependence on the flux limit of<br />

152


CHAPTER 6: SUMMARY OF THE RESULTS<br />

the correlation length of these sources, which could point to the fact that<br />

different populations of sources have different clustering properties.<br />

• The second catalogue of XMM-Newton sources, 2XMM 1 , the largest catalogue<br />

of X-ray sources to date (2XMM is five times larger than the previous<br />

available catalogue, 1XMM 2 ), which contains more than 250,000 sources<br />

(150,000 unique detections) over a sky area of ∼600 sq. degrees, has been<br />

recently released. Cross-correlating 2XMM with already existing catalogues<br />

in other wavelengths such as the Sloan Digital Sky survey (SDSS)<br />

will make possible to isolate potentially interesting sources for posterior<br />

studies (i.e. high-redshift AGN, heavily obscured AGN, etc...)<br />

• Cosmological properties of sources detected in very hard X-rays (4.5-7.5 keV<br />

band), which are very efficient to select samples with small biases due to<br />

absorption, are subject to high uncertainties. Specifically, the dependence<br />

on redshift of the fraction of absorbed AGN in this band is not well constrained,<br />

and some evolution parameters in their luminosity function need<br />

a more detailed study. It is required, in order to address these issues, to<br />

improve the sample statistics. Given that absorbed AGN have very low<br />

angular density, it is required to survey larger sky areas so as to obtain a<br />

significant amount of them to study their evolution at moderate redshifts<br />

( ∼ < 1).<br />

¢¡£¡¥¤§¦©¨¨£¢¡£©£! "#¨£$¥¡%¥&¥ ¢¨'¥(£))#¨<br />

1<br />

¢¡£¡¥¤§¦©¨¨£¢¡£©£! "#¨£$¥¡%¥&¥ ¢¨+*(£))#¨<br />

2<br />

153


Appendix A<br />

Sensitivity maps<br />

The value of the sensitivity map at a given point is defined as the minimum<br />

count rate that must have a source to be detected with the desired likelihood at<br />

that point. As explained in section 2.1.2, the detection likelihood provided by<br />

;D=2?§;2@B;BAD@<br />

takes into account the number of counts within the detection box, the<br />

fit to the PSF shape, and the variation of the exposure map over the detection<br />

box.<br />

While, in principle, the likelihood is not trivially related to the Poisson probability<br />

of having an excess in the number of detected counts over the expected<br />

background rate in the detection box, we have found that the count rate assigned<br />

by the software to a source (observed count rate) is proportional to the<br />

count rate expected from a Poisson distribution for the same likelihood. This<br />

way, for a given likelihood (_0/.B`7a§b L in the SAS source lists), radius of the detection<br />

region (^7c§.,§-_<br />

cutrad in the SAS source lists, in units of 4” pixels), total<br />

value of the background map bgdim and average value of the exposure map in<br />

that region expim, the expected Poisson count rate crpoisim can be calculated<br />

at each point of the detector using the proportionality constants for the corresponding<br />

observed count rates (,0-./ cr in the SAS source lists).<br />

The values of cutrad are different for each source in each field since ;D=7?B;2@B;BAC@<br />

chooses them to maximize the signal-to-noise ratio for source detection. We<br />

have adopted a value of cutrad in each band by averaging the values from all<br />

sources having L between 8 and 20. We have checked that assuming this average<br />

value instead of the individual ones does not significatively affect the<br />

subsequent calculations.<br />

expim is the average of the exposure map within a circle of radius cutrad around<br />

155


APPENDIX A: SENSITIVITY MAPS<br />

Table A.1 Summary of the results of the linear fits of crpoisim dDe7f7g to : Band is the band<br />

used for the fit, cutrad is the source extraction radius (in units of 4 ′′ pixels), LI is the<br />

best fit multiplying constant, N is the number of sources in the fit, χ 2 LI is the χ2 value<br />

of the best linear fit, χ 2 0 is the χ2 of the fit to LI ≡ 1 and P(F) the F-test probability of<br />

allowing LI ̸= 1 and not being a significant improvement to the fit.<br />

Band cutrad LI N χ 2 LI<br />

χ 2 0<br />

P(F)<br />

(pixels)<br />

1 5.12 1.14 284 104.5 185.5 < 10 −6<br />

2 (soft) 5.08 1.10 653 140.1 226.0 < 10 −6<br />

3 5.15 1.14 414 105.7 208.4 < 10 −6<br />

4 (ultrahard) 5.49 1.14 127 25.7 56.2 < 10 −6<br />

5 5.86 1.15 16 1.7 5.5 < 10 −4<br />

9 (XID) 5.04 1.02 722 141.0 145.2 < 10 −5<br />

3-5 (hard) 5.18 0.89 243 176.2 237.8 < 10 −6<br />

the X-ray source positions. The exposure maps for the single bands are PPS<br />

products, whereas for the composite bands were created using the SAS task<br />

;;LF N) is expressed in 4.3.9.<br />

P λ (≥ N) =<br />

∞<br />

∞<br />

λ<br />

∑ P λ (l) = ∑<br />

l e −λ<br />

l!<br />

l=N<br />

l=N<br />

(A.0.2)<br />

If we plot the values ,0-./ of versus crpoisim for the single and composites<br />

bands, it can be seen that there is a linear ,0-.0/<br />

proportionality = LI × crpoisim<br />

for sources with L between 8 and 20 in each band (see Figure A.1). The best fit<br />

values of LI along with the adopted cutrad values are listed in Table A.1. It is<br />

156


APPENDIX A: SENSITIVITY MAPS<br />

clearly seen from the values of χ 2 that this model is fairly good and therefore<br />

more sophisticated models are not needed. The values of LI are within ∼10%<br />

of 1 (equivalent ,0-./ to = crpoisim) but the difference is highly significant in<br />

terms of χ 2 . The empirical estimate of the typical count rate of a source at the<br />

position (X, Y) in image coordinates, detected with likelihood L, is therefore<br />

sens(X, Y) = crpoisim × LI. The areas around the chip edges and around the<br />

bright targets and the OOT regions have also been excluded in the sensitivity<br />

maps.<br />

To summarize, the recipe for creating a sensitivity map is then:<br />

1. Create background and exposure maps<br />

2. Chose a likelihood value L for the significance of the detections. This<br />

recipe has been only tested in the likelihood range 8 ≤ L ≤ 20<br />

3. Choose a source extraction radius cutrad appropriate for the band you are<br />

interested in (Table A.1)<br />

4. For each pixel (X, Y) in your input image:<br />

(a) Calculate the sum of the values of the pixels in the background map<br />

whose centres are within cutrad of (X, Y): bgdim<br />

(b) Calculate the average of the values of the pixels in the exposure map<br />

whose centres are within cutrad of (X, Y): expim<br />

(c) Find ctspoisim such that log(P bgdim (> (ctspoisim + bgdim))) = L<br />

(d) Calculate sens(X, Y) = ctspois/expim×LI.<br />

5. sens(X, Y) is our empirical estimate of the typical count rate of a source at<br />

(X, Y) detected with likelihood L<br />

157


APPENDIX A: SENSITIVITY MAPS<br />

Figure A.1 dDe7f7g<br />

vs. crpoisim for all bands. Also shown are the best linear fits.<br />

158


APPENDIX A: SENSITIVITY MAPS<br />

160


References<br />

[1] Akiyama, M., Ueda, Y., Ohta, K., Takahashi, T. & Yamada, T., 2003, ApJS,<br />

148, 275<br />

[2] Akylas, A., Georgantopoulos, I., Georgakakis, A. et al., 2006, A&A, 459, 693<br />

[3] Akylas, A., Georgantopoulos, I. & Plionis, M., 2000, MNRAS, 318, 1036<br />

[4] Alexander, D. M., Bauer, F. E., Brandt, W. N. et al., 2003, AJ, 126, 539<br />

[5] Antonucci, R., 1993, ARA&A, 31, 473<br />

[6] Antonucci, R. & Miller, J. S., 1985, ApJ, 297, 621<br />

[7] Arnaud, K., 1996, Astronomical Data Analysis Software and Systems V., ed.<br />

G. Jacoby & J. Barnes, ASP Conf. Ser., 101, 17<br />

[8] Arnaud, K., Branduardi-Raymont, G., Culhane, J. L. et al., 1985, MNRAS,<br />

217, 105<br />

[9] Avni, Y. & Bahcall, J. N., 1980, ApJ, 235, 694<br />

[10] Baldi, A., Molendi, S., Comastri, A. et al., 2002, ApJ, 564, 190<br />

[11] Barcons, X., 1992, ApJ, 396, 460<br />

[12] Barcons, X., Carrera, F. J., Ceballos, M. T., Page, M. J. et al., 2007, A&A, 476,<br />

1191<br />

[13] Barcons, X., Carrera, F. J., Watson, M. G. et al., 2002, A&A, 382, 522<br />

[14] Barcons, X., Mateos, S. & Ceballos, M. T., 2000, MNRAS, 311, 456<br />

[15] Barger, A. J., Cowie, L. L., Mushotzky, R. F. & Richards, E. A., 2001, AJ, 121,<br />

662<br />

161


REFERENCES<br />

[16] Barger, A. J., Cowie, L. L., Mushotzky, R. F. et al., 2005, AJ, 129, 578<br />

[17] Basilakos, S., Georgakakis, A., Plionis, M. & Georgantopoulos, G., 2004,<br />

ApJ, 607, L79<br />

[18] Basilakos, S., Plionis, M., Georgakakis, A. & Georgantopoulos, G., 2005,<br />

MNRAS, 356, 183<br />

[19] Bauer, F. E., Alexander, D. M., Brandt, W. N. et al., 2004, AJ, 128, 2048<br />

[20] Benítez, N., 2000, ApJ, 536, 571<br />

[21] Bolzonella, M., Miralles, J. M. & Pelló, R., 2000, A&A, 363, 476<br />

[22] Bongiorno, A., Zamorani, G., Gavignaud, I. et al., 2007, A&A, 472, 443<br />

[23] Boyle, B. J., Griffiths, R. E., Shanks, T. et al., 1993, MNRAS, 260, 49<br />

[24] Boyle, B. J. & Mo, H. J., 1993, MNRAS, 260, 925<br />

[25] Boyle, B. J., Shanks, T., Georgantopoulos, I. et al., 1994, MNRAS, 271, 639<br />

[26] Boyle, B. J., Shanks, T. & Peterson, B. A., 1988, MNRAS, 235, 935<br />

[27] Braccesi, A., Ficarra, A. & Formiggini, L., 1970, A&A, 6, 268<br />

[28] Brandt, W. N., Alexander, D. M., Hornschemeier, A. E. et al., 2001, AJ, 122,<br />

2810<br />

[29] Budavari, T., Connolly, A. J., Szalay, A. S. et al., 2003, ApJ, 595, 59<br />

[30] Caccianiga, A., Severgnini, P., Braito, V. et al., 2004, A&A, 416, 901<br />

[31] Caccianiga, A., Severgnini, P., Della Ceca, R. et al., 2008, A&A, 477, 735<br />

[32] Cagnoni, I., Della Ceca, R. & Maccacaro, T., 1998, ApJ, 493, 54<br />

[33] Campana, S., Moretti, A., Lazzati, D. & Tagliaferri, G., 2001, ApJ, 560, L19<br />

[34] Cappelluti, N., Böhringer, H., Schuecker, P. et al., 2007a, A&A, 465, 35<br />

[35] Cappelluti, N., Cappi, M., Dadina, M. et al., 2005, A&A, 430, 39<br />

[36] Cappelluti, N., Hasinger, G., Brusa, M. et al., 2007b, ApJS, 172, 341<br />

[37] Carrera, F. J., Barcons, X., Fabian, A. C. et al., 1998, MNRAS, 299, 229<br />

162


REFERENCES<br />

[38] Carrera, F. J., Ebrero, J., Mateos, S., Ceballos, M. T. et al., 2007, A&A, 469,<br />

27<br />

[39] Colless, M., Dalton, G., Maddox, S. et al., 2001, MNRAS, 328, 1039<br />

[40] Comastri, A., Fiore, F., Vignali, C. et al., 2001, MNRAS, 327, 781<br />

[41] Comastri, A., Mignoli, M., Ciliegi, P. et al., 2002, ApJ, 571, 771<br />

[42] Comastri, A., Setti, G., Zamorani, G. & Hasinger, G., 1995, A&A, 196, 1<br />

[43] Croom, S. M., Boyle, B. J., Loaring, N. S. et al., 2002, MNRAS, 335, 459<br />

[44] Croom, S. M., Boyle, B. J., Shanks, T. et al., 2005, MNRAS, 356, 415<br />

[45] Croom, S. M., Shanks, T. & Boyle, B. J., 2001, MNRAS, 325, 483<br />

[46] Czerny, B. & Elvis, M., 1987, ApJ, 321, 305<br />

[47] D’Elia, V., Fiore, F., Elvis, M. et al., 2004, A&A, 422, 11<br />

[48] Della Ceca, R., Caccianiga, A., Severgnini, P. et al., 2008, A&A, 487, 119<br />

[49] Della Ceca, R., Castelli, G., Braito, V., Cagnoni, I. & Maccacaro, T., 1999,<br />

ApJ, 524, 674<br />

[50] Della Ceca, R., Maccacaro, T., Caccianiga, A. et al., 2004, A&A, 428, 383<br />

[51] Della Ceca, R., Maccacaro, T., Gioia, I. et al., 1992, ApJ, 389, 491<br />

[52] De Luca, A. & Molendi, S., 2004, A&A, 419, 837<br />

[53] De Zotti, G., Persic, M., Franceschini, A. et al., 1990, ApJ, 351, 22<br />

[54] Den Herder, J. W., Brinkman, A. C., Kahn, S. M. et al., 2001, A&A, 365, L7<br />

[55] Dickey, J. M. & Lockman, F. J., 1990, ARA&A, 28, 215<br />

[56] Dunlop, J. S. & Peacock, J. A., 1990, MNRAS, 247, 19<br />

[57] Dwelly, T. & Page M. J., MNRAS, 372, 1755<br />

[58] Eddington, A. S., 1913, MNRAS, 73, 359<br />

[59] Efstathiou, G., Bernstein, G., Tyson, J. A., Katz, N. & Guhathakurta, P.,<br />

1991, ApJ, 380, L47<br />

[60] Eggen, O. J. & Iben, I., 1989, AJ, 97, 431<br />

163


REFERENCES<br />

[61] Ehle, M., Breitfellner, M., González-Riestra, R. et al., 2005, XMM-Newton<br />

Users Handbook, ESA<br />

[62] Ellis, R. S., Colless, M., Broadhurst, T. el al., 1996, MNRAS, 280, 235<br />

[63] Elvis, M., Schreier, E. J., Tonry, J. et al., 1981, ApJ, 246, 20<br />

[64] Fabian, A. C. & Barcons, X., 1992, ARA&A, 30, 429<br />

[65] Fabian, A. C., Barcons, X., Almaini, O. & Iwasawa, K., 1998, MNRAS, 297,<br />

L11<br />

[66] Fabian, A. C. & Iwasawa, K., 1999, MNRAS, 303, L34<br />

[67] Fanaroff, B. L. & Riley, J. M., 1974, MNRAS, 167, 31<br />

[68] Ferrarese, L., Ford, H. C. & Jaffe, W., 1996, ApJ, 470, 444<br />

[69] Fiore, F., La Franca, F., Giommi, P. et al., 1999, MNRAS, 306, L55<br />

[70] Fischer, J. U., Hasinger, G., Schwope, A. D., Brunner, H. et al., 1998, AN, 6,<br />

347<br />

[71] Gabriel, C., Denby, M., Fyfe, D. J. et al., 2004, ASPC, 314, 759<br />

[72] Gandhi, P., Garcet, O., Loisseau, L. et al., 2006, A&A, 457, 393<br />

[73] Geller, M. J. & Suchra, J. P., 1989, Science, 246, 897<br />

[74] Gendreau, K. C., Barcons, X. & Fabian, A. C., 1998, MNRAS, 297, 41<br />

[75] George, I. M. & Fabian, A. C., 1991, MNRAS, 249, 352<br />

[76] George, I. M., Turner, T. J., Mushotzky, R. et al., 1998, ApJ, 503, 174<br />

[77] Giacconi, R., Bechtold, J., Branduardi, G. et al., 1979, ApJ, 234, L1<br />

[78] Giacconi, R., Gursky, H., Paolini, F. R. & Rossi, B. B., 1962, Phys. Rew. Lett.,<br />

9, 439<br />

[79] Giacconi, R., Rosati, P., Tozzi, P. et al., 2001, ApJ, 551, 624<br />

[80] Giacconi, R., Zirm, A., Wang, J. et al., 2002, APJS, 139, 369<br />

[81] Gierlinski, M. & Done, C., 2004, MNRAS, 349, L7<br />

[82] Gilli, R., 2004, AdSpR, 34, 2470<br />

164


REFERENCES<br />

[83] Gilli, R., Comastri, A. & Hasinger, G., 2007, A&A, 463, 79<br />

[84] Gilli, R., Daddi, E., Zamorani, G. et al., 2005, A&A, 430, 811<br />

[85] Gilli, R., Salvati, M. & Hasinger, G., 2001, A&A, 366, 407<br />

[86] Goodrich, R. W., Veilleux, S. & Hill, G. J., 1994, ApJ, 422, 521<br />

[87] Granato, G. L. & Danese, L., 1994, MNRAS, 268, 235<br />

[88] Granato, G. L., Danese, L. & Franceschini, A., 1997, ApJ, 486, 197<br />

[89] Grazian, A., Negrello, L., Moscardini, L. et al., 2004, AJ, 127, 592<br />

[90] Griffiths, R. E., Georgantopoulos, I., Boyle, B. J. et al., 1995, MNRAS, 275,<br />

77<br />

[91] Griffiths, R. E., Murray, S. S., Giacconi, R. et al., 1983, ApJ, 269, 375<br />

[92] Guainazzi, M., Perola, G. C., Matt, G. et al., 1999, A&A, 346, 407<br />

[93] Guilbert, P. W. & Rees, M. J., 1988, MNRAS, 233, 475<br />

[94] Haardt, F. & Maraschi, L., 1991, ApJ, 380, L51<br />

[95] Haardt, F. & Maraschi, L., 1993, ApJ, 413, 507<br />

[96] Hamilton, T. T., Helfand, D. J. & Wu, X., 1991, ApJ, 476, 620<br />

[97] Harrison, F. A., Eckart, M. E., Mao, P. H. et al., 2003, ApJ, 596, 944<br />

[98] Hasinger, G., 1993, AdSpR, 13, 241<br />

[99] Hasinger, G., 1998, AN, 319, 37<br />

[100] Hasinger, G., 2008, A&A, in press, astro-ph/0808.0260<br />

[101] Hasinger, G., Altieri, B., Arnaud, M. et al., 2001, A&A, 365, L45<br />

[102] Hasinger, G., Burg, R., Giacconi, R. et al., 1998, A&A, 329,482<br />

[103] Hasinger, G., Cappelluti, N., Brunner, H. et al., 2007, ApJS, 172, 29<br />

[104] Hasinger, G., Miyaji, T. & Schmidt, M., 2005, A&A, 441, 417<br />

[105] Hawkins, E., Maddox, S., Cole, S. et al., 2003, MNRAS, 346, 78<br />

[106] Hawkins, M. R. S. & Veron, P., 1995, MNRAS, 275, 1102<br />

165


REFERENCES<br />

[107] Hawkins, M. R. S. & Veron, P., 1996, MNRAS, 281, 348<br />

[108] Hewett, P. C., Foltz, C. B. & Chaffee, F. H., 1993, ApJ, 406, L43<br />

[109] Hickox, R. C. & Markevitch, M., 2006, ApJ, 645, 95<br />

[110] Hogg, D. W., 1999, astro-ph/9905116<br />

[111] Holt, S. S., Mushotzky, R. F., Boldt, E. A. et al., 1980, ApJ, 241, L13<br />

[112] Hopkins, P. F., Richards, G. T. & Hernquist, L., 2007, ApJ, 654, 731<br />

[113] Hornschemeier, A. E., Bauer, F. E., Alexander, D. M. et al., 2003, AJ, 126,<br />

575<br />

[114] Hornschemeier, A. E., Brandt, W. N., Garmire, G. P. et al., 2000, ApJ, 541,<br />

49<br />

[115] Hornschemeier, A. E., Brandt, W. N., Garmire, G. P. et al., 2001, ApJ, 554,<br />

742<br />

[116] James, F., 1994, MINUIT reference manual, CERN Program Library<br />

Writeup D506, CERN, Geneva<br />

[117] Jansen, F., Lumb, B., Altieri, B. et al., 2001, A&A, 365, L1<br />

[118] Jones, L. R., McHardy, I. M., Merrifield, M. R. et al., 1996, MNRAS, 285,<br />

547<br />

[119] Kaastra, J. S., 1992, An X-Ray Spectral Code for Optically Thin Plasmas<br />

(Internal <strong>SRON</strong>-Leiden Report, updated version 2.0)<br />

[120] Kaastra, J. S., Steenbrugge, K. C., Raassen, A. J. J. et al., 2002, A&A, 386,<br />

427<br />

[121] Kaspi, S., Brandt, W. N., Netzer, H. et al., 2001, ApJ, 554, 216<br />

[122] Kendziorra, E., Bihler, E., Grubmiller, W. et al., 1997, SPIE Proc., 3114, 155<br />

[123] Kendziorra, E., Colli, M., Kuster, M. et al., 1999, SPIE Proc., 3765, 204<br />

[124] Kim, D. W., Cameron, R. A., Drake, J. J. et al., 2004, ApJS, 150, 19<br />

[125] Kim, M., Wilkes, B. J., Kim, D., Green, P. J. et al., 2007, ApJ, 659, 29<br />

[126] Kormendy, J. & Richstone, D., 1995, ARA&A, 33, 581<br />

166


REFERENCES<br />

[127] Krolik, J. H. & Begelman, M. C., 1988, ApJ, 329, 702<br />

[128] Kuster, M., Benlloch, S., Kendziorra, E. & Briel, U. G., 1999, SPIE Proc.,<br />

3765, 673<br />

[129] La Franca, F. & Cristiani, S., 1997, AJ, 113, 1517<br />

[130] La Franca, F., Fiore, F., Comastri, A., Perola, G. C. et al., 2005, ApJ, 635,<br />

864<br />

[131] Landy, S. D. & Szalay, A., 1993, ApJ, 412, 64<br />

[132] Laor, A., 1990, MNRAS, 246, 369<br />

[133] Laor, A. & Netzer, H., 1989, MNRAS, 238, 897<br />

[134] Lehmann, I., Hasinger, G., Böhringer, H. et al., 2004, AAS Meeting Abstracts,<br />

205<br />

[135] Liedahl, D. A., Osterheld, A. L. & Goldstein, W. H., 1995, ApJL, 438, 115<br />

[136] Lightman, A. P. & White, T. R., 1988, ApJ, 335, 57<br />

[137] Loaring, N., Dwelly, T., Page, M. J. et al., 2005, MNRAS, 362, 1371<br />

[138] Lockman, F. J., Jahoda, K. & McCammon, D., 1986, ApJ, 302, 432<br />

[139] Lumb, D. H., Warwick, R. H., Page, M. J. & De Luca, A., 2002, A&A, 389,<br />

93<br />

[140] Maccacaro, T., Della Ceca, R., Gioia, I. M. et al., 1991, ApJ, 374, 117<br />

[141] Magorrian, J., Tremaine, S., Richstone, D. et al., 1998, AJ, 115, 2285<br />

[142] Maiolino, R., Comastri, A., Gilli, R. et al., 2003, MNRAS, 344, L59<br />

[143] Malizia, A., Bassani, L., Stephen, J. B. et al., 1997, ApJS, 113, 311<br />

[144] Marconi, A., Risaliti, G., Gilli, R. et al., 2004, MNRAS, 351, 169<br />

[145] Marshall, H. L., Avni, Y., Tananbaum, H. & Zamorani G., 1983, ApJ, 269,<br />

35<br />

[146] Marshall, F. E., Boldt, E. A., Holt, S. S. et al., 1980, ApJ, 235, 4<br />

[147] Mason, K. O., Carrera, F. J., Hasinger, G., Andernach, H. et al., 2000, MN-<br />

RAS, 311, 456<br />

167


REFERENCES<br />

[148] Matarrese, S., Coles, P., Lucchin, F. & Moscardini, L., 1997, MNRAS, 286,<br />

115<br />

[149] Mateos, S., Barcons, X., Carrera, F. J. et al., 2005a, A&A, 433, 855<br />

[150] Mateos, S., Barcons, X., Carrera, F. J. et al., 2005b, A&A, 444, 79<br />

[151] Mateos, S., Warwick, R. S., Carrera, F. J. et al., 2008, A&A, in press, astroph/0809.1939<br />

[152] Mather, J. C., Cheng, E. S., Cottingham, D. A. et al., 1994, ApJ, 420, 439<br />

[153] McCammon, D. & Sanders, W. T., 1990, ARA&A, 28, 657<br />

[154] Mewe, R., Gronenschild, E. H. B. M. & Van Den Oord, G. H. J., 1985,<br />

A&AS, 62, 197<br />

[155] Mewe, R., Lemen, J. R. & Van Den Oord, G. H. J., 1986, A&AS, 65, 511<br />

[156] Miyaji, T. & Griffiths, R. E., 2002, ApJ, 564, L5<br />

[157] Miyaji, T., Hasinger, G. & Schmidt, M., 2000, A&A, 353, 25<br />

[158] Miyaji, T., Hasinger, G. & Schmidt, M., 2001, A&A, 369, 49<br />

[159] Miyaji, T., Zamorani, G., Cappelluti, N., et al., 2007, ApJSS, 172, 396<br />

[160] Moretti, A., Campana, S., Lazzati, D. & Tagliaferri, G., 2003, ApJ, 588, 696<br />

[161] Mullis, C. R., Henry, J. P., Gioia, I. M. et al., 2004, ApJ, 617, 192<br />

[162] Mushotzky, R. F., 1982, ApJ, 256, 92<br />

[163] Mushotzky, R. F., Cowie, L. L., Barger, A. J. & Arnaud, K. A., 2000, Nature,<br />

404, 459<br />

[164] Nandra, K. & Pounds, K. A., 1994, MNRAS, 268, 405<br />

[165] Nandra, K., Pounds, K. A., Stewart, G. C. et al., 1991, MNRAS, 248, 760<br />

[166] Nicastro, F., Piro, L., De Rosa, A. et al., 2000, ApJ, 536, 718<br />

[167] Osmer, P. S., 1981, ApJ, 247, 762<br />

[168] Overzier, R. A., Röttgering, H., Rengelink, R. B. & Wilman, R. J., 2003,<br />

A&A, 405, 33<br />

168


REFERENCES<br />

[169] Padovani, P., 1993, MNRAS, 263, 461<br />

[170] Page, M. J. & Carrera, F. J., 2000, MNRAS, 311, 433<br />

[171] Page, M. J., Carrera, F. J., Hasinger, G. et al., 1996, MNRAS, 281, 576<br />

[172] Page, M. J., Mason, K. O., McHardy, I. M. et al., 1997, MNRAS, 291, 324<br />

[173] Page, M. J., Mittaz, J. P. D. & Carrera, F. J., 2000, MNRAS, 318, 1073<br />

[174] Peebles, P. J. E., 1980, The large-scale structure of the Universe (Princeton<br />

University Press)<br />

[175] Peebles, P. J. E., 1993, Principles of Physical Cosmology (Princeton University<br />

Press)<br />

[176] Perola, G. C., Matt, G., Cappi, M. et al., 1999, A&A, 351, 937<br />

[177] Perola, G. C., Matt, G., Fiore, F. et al., 2000, A&A, 358, 117<br />

[178] Perola, G. C., Puccetti, S., Fiore, F. et al., 2004, A&A, 421, 491<br />

[179] Peterson, B. M., 2001, Advanced Lectures in Starburst-AGN Connection,<br />

Page 3, astro-ph/0109495<br />

[180] Piccinotti, G., Mushotzky, R. F., Boldt, E. A. et al., 1982, ApJ, 253, 485<br />

[181] Pier, E. A. & Krolik, J. H., 1992, ApJ, 399, L23<br />

[182] Pier, E. A. & Krolik, J. H., 1993, ApJ, 418, 673<br />

[183] Pierre, M., Valtchanov, I., Altieri, B. et al., 2004, Cosm. Astropart. Phys., 9,<br />

11<br />

[184] Pounds, K. A., Nandra, K., Stewart, G. C. et al., 1990, Nature, 344, 132<br />

[185] Pounds, K. A., Warwick, R. S., Culhane, J. L. & De Korte, P. A. J., 1986,<br />

MNRAS, 218, 685<br />

[186] Primini, F. A., Murray, S. S., Huchra, J. et al., 1991, ApJ, 374, 440<br />

[187] Puccetti, S., Fiore, F., D’Elia, V. et al., 2006, A&A, 457, 501<br />

[188] Ratcliffe, A., Shanks, T., Parker, Q. A. & Fong, R., 1998, MNRAS, 296, 173<br />

[189] Reynolds, C. S., 1997, MNRAS, 286, 513<br />

169


REFERENCES<br />

[190] Richstone, D., Ahjar, E. A., Bender, R. et al., 1998, Nature, 395, 14<br />

[191] Risaliti, G. & Elvis, M., 2004, ASSL, 308, 187<br />

[192] Rix, H.-W., Rieke, G., Rieke, M. & Carleton, N. P., 1990, ApJ, 363, 480<br />

[193] Roche, N. D., Dunlop, J. & Almaini, 0., 2003, MNRAS, 346, 803<br />

[194] Rosati, P., Tozzi, P., Giacconi, R. et al., 2002, ApJ, 566, 667<br />

[195] Ross, R. R. & Fabian, A. C., 1993, MNRAS, 258, 189<br />

[196] Ross, R. R., Fabian, A. C. & Mineshige, S., 1992, MNRAS, 258, 189<br />

[197] Röttgering, H., Daddi, E., Overzier, R. A. & Wilman, R. J., 2003, NewAR,<br />

47, 309<br />

[198] Rowan-Robinson, M., Benn, C. R. & Lawrence, A., 1993, MNRAS, 263,<br />

123<br />

[199] Ruiz, M., Rieke, G. H. & Sheilds, J. D., 1994, AAS Bulletin, 26, 1356<br />

[200] Sanders, D. B., Phinney, E. S., Neugebauer, G. et al., 1989, ApJ, 347, 29<br />

[201] Sazonov, S., Revnivtsev, M., Kryvonos, R. et al., 2007, A&A, 462, 57<br />

[202] Scheuer, P. A. G., 1974, MNRAS, 166, 329<br />

[203] Schmidt, M., 1968, ApJ, 151, 393<br />

[204] Schmidt, M., Hasinger, G., Gunn, J., Schneider, D. et al., 1998, A&A, 329,<br />

495<br />

[205] Schmidt, M., Schneider, D. P., Gunn, J. E. et al., 1995, AJ, 110, 68<br />

[206] Schuecker, P., Böhringer, H., Reiprich, T. H. & Feretti, L., 2001, A&A, 378,<br />

408<br />

[207] Schwope, A., Hasinger, G., Lehmann, I. et al., 2000, AN, 321, 1<br />

[208] Severgnini, P., Caccianiga, A., Braito, V. et al., 2003, A&A, 406, 483<br />

[209] Shakura, N. I. & Sunyaev, R. A., 1973, A&A, 24, 337<br />

[210] Shastri, P., Wilkes, B. J., Elvis, M. & McDowell, J., 1993, ApJ, 410, 29<br />

[211] Silverman, J. D., Green, P. J., Barkhouse, W. A. et al., 2008, ApJ, 679, 118<br />

170


REFERENCES<br />

[212] Simpson, C., 2005, MNRAS, 360, 565<br />

[213] Spergel, D. N., Verde, L., Peiris, H. V. et al., 2003, ApJS, 148, 175<br />

[214] Springel, V., Frenk, C. S. & White, D. M., 2006, Nature, 440, 1137<br />

[215] Steffen, A. T., Barger, A. J., Cowie L. L. et al., 2003, ApJ, 569, L23<br />

[216] Strüdel, L., Briel, U., Dennerl, K. et al., 2001, A&A, 365, L18<br />

[217] Szokoly, G. P., Bergeron, J., Hasinger, G. et al., 2004, ApJS, 155, 271<br />

[218] Totsuji, H. & Kihara, T., 1969, PASJ, 21, 221<br />

[219] Tozzi, P., Rosati, P., Nonino, M., Bergeron, J. et al., 2001, ApJ, 562, 42<br />

[220] Treister, E. & Urry, C. M., 2006, ApJ, 652, L79<br />

[221] Tucker, D. L., Oemler, A., Kirshner, R. P. et al., 1997, MNRAS, 285, L5<br />

[222] Turner, M. J. L., Abbey, A., Arnaud, M. et al., 2001, A&A, 365, L27<br />

[223] Turner, T. J. & Pounds, K. A., 1989, MNRAS, 240, 833<br />

[224] Ueda, Y., Akiyama, M., Ohta, K. & Miyaji, T., 2003, ApJ, 598,886<br />

[225] Ueda, Y., Ishisaki, Y., Takahashi, T., Makishima, K. & Ohashi, T., 2005,<br />

ApJS, 161, 185<br />

[226] Ueda, Y., Takahashi, T., Inoue, H. et al., 1998, Nature, 391, 866<br />

[227] Ueda, Y., Watson, M. G., Stewart, I. M. et al., 2008, ApJS, in press, astroph/08062846<br />

[228] Urry, C. M. & Padovani, P., 1995, PASP, 107, 803<br />

[229] Vikhlinin, A. & Forman, W., 1995, ApJ, 455, L109<br />

[230] Vikhlinin, A., Forman, W., Jones, C. & Murray, S., 1995a, ApJ, 451, 564<br />

[231] Vikhlinin, A., Forman, W., Jones, C. & Murray, S., 1995b, ApJ, 451, 553<br />

[232] Voges, W., Aschenbach, B., Boller, T. et al., 1999, A&A, 349, 389<br />

[233] Voges, W., Henry, J. P., Briel, U. G. et al., 2001, ApJ, 553, L119<br />

[234] Watson, M. G., Auguères, J. L., Ballet, J. et al., 2001, A&A, 365, L51<br />

[235] Wilkes, B. J. & Elvis, M., ApJ, 323, 243<br />

171


REFERENCES<br />

[236] Wisotzky, L., 1998, AN, 319, 257<br />

[237] Wolf, C., Dye, S., Kleinheinrich, M. et al., 2001, A&A, 377, 442<br />

[238] Wolf, C., Meisenheimer, K., Kleinheinrich, M. et al., 2004, A&A, 421, 913<br />

[239] Wolf, C., Meisenheimer, K., Rix, H. W., Borch, A., Dye, S. & Kleinheinrich,<br />

M., 2003, A&A, 401, 73<br />

[240] Worsley, M. A., Fabian, A. C., Barcons, X. et al., 2004, MNRAS, 352, 128<br />

[241] Yang, Y., Mushotzky, R. F., Barger, A. J. et al., 2003, ApJ, 585, L85<br />

[242] Yang, Y., Mushotzky, R. F., Barger, A. J. & Cowie, L. L., 2006, ApJ, 645, 68<br />

[243] York, D. G., Adelman, J., Anderson, J. E. et al., 2000, AJ, 120, 1579<br />

[244] Yu, Q. & Tremaine, S., 2002, MNRAS, 335, 965<br />

[245] Zamorani, G., Gioia, I. M., Maccacaro, T. & Wolter, A., 1988, A&A, 196, 39<br />

[246] Zheng, W., Mikles, V. J., Mainieri, V., Hasinger, G. et al., 2004, ApJS, 155,<br />

73<br />

172

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!