Simulación atomística de la producción y evolución de ...

132
Simulación atomística de la producción y evolución de defectos en aleaciones basadas en Fe debido a la irradiación María José Aliaga Gosálvez

Transcript of Simulación atomística de la producción y evolución de ...

Page 1: Simulación atomística de la producción y evolución de ...

Simulación atomística de la producción y evolución de defectos en aleaciones basadas en Fe debido a la irradiación

María José Aliaga Gosálvez

Page 2: Simulación atomística de la producción y evolución de ...

Departamento de Física Aplicada – Instituto Universitario de Materiales Facultad de Ciencias

Simulación atomística de la producción y evolución de defectos en aleaciones basadas en Fe debido a

la irradiación

María José Aliaga Gosálvez

Tesis presentada para aspirar al grado de

DOCTORA POR LA UNIVERSIDAD DE ALICANTE

MENCIÓN DE DOCTORA INTERNACIONAL

DOCTORADO EN CIENCIA DE MATERIALES

Dirigida por la doctora: María José Caturla Terol

Financiada por el programa de Ayudas para becas y contratos destinados a la formación

de doctores del Vicerrectorado de Investigación, Desarrollo e Innovación (Resolución 22

de diciembre de 2011)

Page 3: Simulación atomística de la producción y evolución de ...
Page 4: Simulación atomística de la producción y evolución de ...

A toda mi familia

A todos mis amigos/as

A mi directora de tesis María José

A Ben

y a Max

Page 5: Simulación atomística de la producción y evolución de ...
Page 6: Simulación atomística de la producción y evolución de ...

i

Resumen

Esta tesis doctoral se centra en el estudio a nivel atómico de la producción de

defectos en aleaciones basadas en Fe, particularmente aleaciones de Fe-Cr,

debido a la interacción con partículas energéticas. Uno de los grandes retos a

los que se enfrenta la energía de fusión es el desarrollo de materiales

resistentes a los altos niveles de radiación a los que se verán sometidos [1]. La

falta de fuentes de radiación que reproduzcan las condiciones que existirán en

los reactores de fusión hace necesario el desarrollo de modelos que permitan

extrapolar los resultados experimentales obtenidos con fuentes de radiación

como iones o neutrones de fisión, a las condiciones de fusión. Tales modelos

sólo son posibles si se conoce el comportamiento del material bajo irradiación a

nivel fundamental, desde la formación de defectos a nivel atómico y en tiempos

de picosegundos, hasta la evolución de estos defectos a tiempos de horas o

incluso años. Este tipo de simulaciones requiere de la unión de varios métodos

de cálculo distintos en lo que se denomina modelización multiescala [2].

Además, el desarrollo de estos modelos requiere de una validación de los

resultados obtenidos a través de comparaciones con experimentos. La

radiación de los materiales con iones y su caracterización mediante

microscopía electrónica de transmisión (TEM por sus siglas en inglés,

Transmission Electron Microscopy), son las técnicas de referencia para validar

los modelos. Es importante tener en cuenta las diferencias entre irradiación con

iones e irradiación con neutrones, especialmente cuando la irradiación con

iones se lleva a cabo mientras el material está siendo observado con TEM

(TEM in-situ), ya que requiere que la muestra sea una lámina muy fina, en

comparación con la irradiación con neutrones que sucede en el interior del

material (bulk).

Uno de los materiales de mayor interés para la fusión son las aleaciones de

Fe, consideradas como el principal candidato para los materiales estructurales

de estos reactores [3]. Debemos así pues comprender los cambios en las

propiedades mecánicas de estos materiales debido a la producción de

defectos.

Page 7: Simulación atomística de la producción y evolución de ...

ii

En esta tesis doctoral se han utilizado tres técnicas de simulación para estudiar

los procesos de formación y de evolución de defectos en hierro y en Fe-Cr,

centrándonos en estudiar las diferencias entre irradiación en capas finas y en

muestras en volumen . En primer lugar se ha utilizado la técnica de Dinámica

Molecular con potenciales empíricos [4] para estudiar cómo se produce el daño

en los primeros picosegundos después de la irradiación. Con esta técnica

hemos calculado el número y tipo de defectos que se producen al irradiar con

iones de Fe láminas finas de este mismo material para TEM in-situ, así como

muestras de bulk. La base de datos obtenida de cascadas de desplazamiento

se ha utilizado a continuación como entrada al método de simulación de Monte

Carlo Cinético para analizar la evolución del daño primario por la interacción y

difusión de los defectos. Finalmente hemos utilizado algunas de las cascadas

para simular imágenes de microscopía electrónica de transmisión (TEM). Los

resultados obtenidos con cada técnica se han comparado con resultados

experimentales de otros grupos de investigación del mismo proyecto y con

experimentos diseñados y realizados por este mismo grupo.

El informe que sigue a continuación consiste en un resumen global del trabajo

realizado, los artículos publicados y los que se encuentran en vía de

publicación. Su estructura es la siguiente. El capítulo 1 es una introducción al

daño por radiación en materiales para fusión y su modelización. El capítulo 2

resume la metodología de la modelización multiescala utilizada en esta tesis.

En el capítulo 3 se discuten los resultados más relevantes de los trabajos

publicados y no publicados que se encuentran en los capítulos 4 y 5. El

capítulo 6 concluye el informe con las conclusiones globales de la tesis.

Los trabajos presentados en los capítulos 4 y 5 son:

M.J. Aliaga, A. Prokhodtseva, R. Schaeublin, M.J. Caturla (2014),

Molecular dynamics simulations of irradiation of α-Fe thin films with

energetic Fe ions under channeling conditions, Journal of Nuclear

Materials, 452, 453.

M.J. Aliaga, M.J. Caturla, R. Schaeublin (2015), Surface damage in TEM

thick α-Fe samples by implantation with 150 keV Fe ions, Nuclear

Instruments and Methods in Physics Research B, 352, 217.

Page 8: Simulación atomística de la producción y evolución de ...

iii

M.J. Aliaga, R. Schaeublin, J.F. Löffler, M.J. Caturla (2015), Surface-

induced vacancy loops and damage dispersión in irradiated Fe thin

films, Acta Materialia, 101, 22.

M.J. Aliaga, I. Dopico, I. Martin-Bragado, M.J. Caturla (2016), Influence

of free surfaces on microstructure evolution of radiation damage in Fe

from molecular dynamics and object kinetic Monte Carlo calculations,

Physica Status Solidi A, enviado.

M.J. Aliaga, I. Dopico, I. Martin-Bragado, M. Hernández-Mayoral, L.

Malerba, M.J. Caturla, Insights on loop nucleation and growth in α-Fe

thin films under ion implantation fron atomistic models.

A.E. Sand, M.J. Aliaga, M.J. Caturla, K. Nordlund, Surface effects and

statistical laws of defects in primary radiation damage: tungsten vs. Iron.

S. García-González, A. Rivera, M.J. Aliaga, M.J. Caturla, I. Martín-

Bragado, OKMC study of differences between MD and BCA cascades in

neutron irradiated Fe simulations.

Page 9: Simulación atomística de la producción y evolución de ...

iv

Page 10: Simulación atomística de la producción y evolución de ...

v

Abstract

This thesis is focused on the study, at the atomic level, of the production of

defects in Fe based alloys, particularly FeCr alloys, due to the interaction with

energetic particles. One of the biggest challenges of fusion energy is the

development of resistant materials to the high level of radiation that they will

have to withstand in the nuclear reactor [1]. The lack of radiation sources able

to reproduce the exact conditions in a future fusion plant makes it necessary the

development of models that will permit the extrapolation of experimental results

of radiation with ions or fission neutrons, to fusion conditions. These models are

only possible if we know the behaviour of the material under irradiation at a

fundamental level, from the formation of defects at the atomic level and

picoseconds time scale, to the evolution of these defects at a level of hours or

even years. These kinds of simulations require the union of several methods in

what is known as multiscale modelling [2]. Importantly, these models need to be

validated with experiments. The reference technique to validate the models and

characterize the damage produced by irradiation is the Transmission Electron

Microscopy (TEM). It is important to have into account the differences between

ion and neutron irradiation, particularly when the ion irradiation takes place

while the material is being observed under the TEM, what is called in-situ TEM,

because the sample needs to be a thin film, in comparison with neutron

irradiation, that takes place in the bulk.

Some of the materials of major interest to fusion are Fe base alloys,

considered as the main candidates for structural materials of the reactor [3]. In

this way, it is essencial to know the changes in the mechanical properties of

these materials due to production of defects.

In this thesis three simulation techniques have been used to study the formation

and evolution of defects in Fe and FeCr. The focus has been in studying the

differences between irradiation in thin films and irradiation in bulk. First, we

have used molecular dynamics with empirical potentials [4] to study how

damage is produced during the first picoseconds after irradiation. With this

technique we have calculated the number and type of defects produced while

irradiating Fe thin films with ions for in-situ TEM, as well as bulk samples. The

Page 11: Simulación atomística de la producción y evolución de ...

vi

obtained database has then been used as input for the Monte Carlo technique

to analyse the evolution of the primary damage by interaction and diffusion of

defects. Finally, we have used some of the cascades to simulate TEM images.

Results obtained with this technique have been compared to experimental

results of other groups in the same project and with experiments designed and

performed by this group.

This report consists of a global summary of the objectives, methods and main

results of the thesis. It is structured as follows. Chapter 1 is an introduction to

radiation damage in materials for fusion and its modelling. Chapter 2 gives a

description of the multiscale methodology used in this thesis. Chapter 3

discusses the most relevant results of the papers, which are in chapters 4 and

5. Finally, chapter 6 closes the report with the conclusions.

The papers in chapters 4 and 5 are:

M.J. Aliaga, A. Prokhodtseva, R. Schaeublin, M.J. Caturla (2014),

Molecular dynamics simulations of irradiation of α-Fe thin films with

energetic Fe ions under channeling conditions, Journal of Nuclear

Materials, 452, 453.

M.J. Aliaga, M.J. Caturla, R. Schaeublin (2015), Surface damage in TEM

thick α-Fe samples by implantation with 150 keV Fe ions, Nuclear

Instruments and Methods in Physics Research B, 352, 217.

M.J. Aliaga, R. Schaeublin, J.F. Löffler, M.J. Caturla (2015), Surface-

induced vacancy loops and damage dispersión in irradiated Fe thin

films, Acta Materialia, 101, 22.

M.J. Aliaga, I. Dopico, I. Martin-Bragado, M.J. Caturla (2016), Influence

of free surfaces on microstructure evolution of radiation damage in Fe

from molecular dynamics and object kinetic Monte Carlo calculations,

Physica Status Solidi A, enviado.

Page 12: Simulación atomística de la producción y evolución de ...

vii

M.J. Aliaga, I. Dopico, I. Martin-Bragado, M. Hernández-Mayoral, L.

Malerba, M.J. Caturla, Insights on loop nucleation and growth in α-Fe

thin films under ion implantation fron atomistic models.

A.E. Sand, M.J. Aliaga, M.J. Caturla, K. Nordlund, Surface effects and

statistical laws of defects in primary radiation damage: tungsten vs. Iron.

S. García-González, A. Rivera, M.J. Aliaga, M.J. Caturla, I. Martín-

Bragado, OKMC study of differences between MD and BCA cascades in

neutron irradiated Fe simulations.

Page 13: Simulación atomística de la producción y evolución de ...

viii

Page 14: Simulación atomística de la producción y evolución de ...

ix

ÍNDICE

RESUMEN i

ABSTRACT v

1 INTRODUCCIÓN 1

1.1. Daño por radiación en materiales para fusión 4

1.2. Modelización del daño por radiación: modelos multiescala 6

1.3. Propósito de este trabajo 9

2 METODOLOGÍA 11

2.1 Dinámica Molecular 11

2.1.1. Potenciales de interacción 12

2.1.2. El modelo del átomo embebido 13

2.1.3. Métodos de integración 15

2.1.4. Condiciones de contorno 17

2.1.5. El código MDCASK 20

2.1.6. Simulación de irradiación de Fe en Fe: antecedentes 21

2.2 Monte Carlo Cinético 22

2.2.1. La teoría del estado de transición 25

2.2.2. El código MMonCa 26

2.3. Simulación de imágenes TEM 28

Page 15: Simulación atomística de la producción y evolución de ...

x

2.3.1. El método multicapa 28

2.3.2. El código EMS 30

3 RESULTADOS Y DISCUSIÓN 32

3.1. Daño en escala de picosegundos: iones frente a neutrones 32

3.2. Evolución del daño primario en Fe y en FeCr 35

3.3. Comparación con imágenes TEM 36

4 TRABAJOS PUBLICADOS O ACEPTADOS 41

I. Molecular dynamics simulations of irradiation of α-Fe thin films with

energetic Fe ions under channeling conditions

43

II. Surface damage in TEM thick α-Fe samples by implantation with

150 keV Fe ions

49

III. Surface-induced vacancy loops and damage dispersión in

irradiated Fe thin films

55

IV. Influence of free surfaces on microstructure evolution of radiation

damage in Fe from molecular dynamics and object kinetic Monte

Carlo calculations

67

5 TRABAJOS NO PUBLICADOS 75

V. Surface effects and statistical laws of defects in primary radiation

damage: tungsten vs. iron

77

VI. OKMC study of differences between MD and BCA cascades in

neutron irradiated Fe simulations

87

VII. Insights on loop nucleation and growth in α-Fe thin films under ion 95

Page 16: Simulación atomística de la producción y evolución de ...

xi

implantation fron atomistic models

CONCLUSIONES 107

CONCLUSIONS 109

BIBLIOGRAFÍA 111

Page 17: Simulación atomística de la producción y evolución de ...

xii

Page 18: Simulación atomística de la producción y evolución de ...

1

1. INTRODUCCIÓN

El cambio climático es un tema que preocupa a la mayoría de la población y a

sus gobiernos, y se están tomando muchas medidas desde hace años para

intentar frenarlo o atenuar al menos sus efectos. El principal objetivo de estas

medidas es la sustitución del petróleo, principal causante del gas de efecto

invernadero CO2, por energías alternativas que no contribuyan al cambio

climático. Una de estas energías es la energía de fusión.

Llegar a utilizar comercialmente la energía de fusión para producir electricidad

es un gran reto, ya que los requerimientos para que se produzcan reacciones

de fusión son muy exigentes, y a diferencia de las reacciones de fisión, si estas

condiciones no se cumplen el proceso simplemente se detiene, haciendo

inexistente el riesgo de las reacciones en cadena. Estos requerimientos

específicos son parte de la razón por la cual no existe todavía a día de hoy un

reactor de fusión que produzca más energía de la que consume.

Para conseguir la fusión se debe forzar a los núcleos de los átomos de deuterio

(D) y tritio (T) a unirse formando helio (He) y neutrones con una energía de 14

MeV. Esto se consigue calentando el combustible a unos 100 millones de

grados para que alcance el estado de plasma. El confinamiento de este plasma

es el parámetro más crítico y el más desafiante. Hoy día existen dos métodos

para conseguir el confinamiento, uno de ellos es mediante la utilización de

láseres de alta potencia, conocido como fusión por confinamiento inercial,

como NIF en los EEUU [25] o el LaserMegaJoule en Francia [26]. Un segundo

método consiste en la utilización de campos magnéticos, conocido como fusión

por confinamiento magnético. Este es el sistema en el que se enmarca el

proyecto de investigación de esta tesis.

El dispositivo utilizado en el confinamiento magnético es el llamado tokamak

(que en ruso significa “cámara toroidal con espirales magnéticas”). En un

reactor tokamak, el plasma se mantiene confinado mediante líneas de campo

magnético en una cámara de vacío con forma de toroide [6].

Page 19: Simulación atomística de la producción y evolución de ...

2

Los neutrones altamente energéticos que se forman en la reacción de fusión

del D+T serán absorbidos en agua y el vapor de agua generado moverá una

turbina como sucede en los reactores de fisión convencionales. Pero partículas

del plasma escaparán inevitablemente al confinamiento y golpearán las

paredes del reactor. Los materiales que conforman la primera pared del reactor

en contacto directo con el plasma sufrirán erosión, pero además, los altamente

energéticos neutrones pueden penetrar en los materiales estructurales del

reactor.

Las propiedades mecánicas de estos materiales se alteran debido a los

defectos inducidos por la radiación, siendo la fragilización y el aumento de

volumen las principales consecuencias [9]. Las condiciones de fusión son

mucho más agresivas para los materiales del reactor que las condiciones de

fisión, por lo que se hace necesaria la utilización de nuevos materiales que

alargue el tiempo de vida del reactor. Con esta finalidad, diferentes escenarios

y materiales están siendo testeados tanto en reactores existentes como en

instalaciones experimentales. Por ejemplo, diferentes tipos de aceros

ferríticos/martensíticos especiales de baja activación se están desarrollando

por considerarse los mejores candidatos para ser utilizados como materiales

estructurales. Sin embargo, una evaluación experimental completa es difícil de

conseguir, debido por un lado a la gran variedad de materiales, y por otro a la

imposibilidad de reproducir las condiciones exactas del plasma en el reactor de

la próxima generación ITER [6]. A día de hoy no existe ningún dispositivo

experimental capaz de producir neutrones de 14 MeV como en un reactor de

fusión. Por esta razón, las técnicas de simulación por ordenador, tales como las

empleadas en este trabajo, son una herramienta vital para proporcionar

información y comprensión fundamentales del comportamiento de un material

específico para un futuro reactor de fusión.

En este sentido comenzó en el año 2002, dentro del programa europeo de

fusión conocido como EFDA (European Fusion Development Agreement),

transformado hoy en EUROfusion, una tarea de modelización de materiales

para fusión. Esta tarea, dentro de la cual se enmarca esta tesis, incluye la

Page 20: Simulación atomística de la producción y evolución de ...

3

validación de los modelos desarrollados a través de distintos métodos

experimentales, en particular, la irradiación con iones.

Figura 1: Una ilustración del reactor de fusión de próxima generación ITER [6]

resaltando sus principales componentes. El volumen del plasma es de 850 m3 y

el radio principal del toroide es de 6,2 m. El objetivo de ITER es conseguir un

factor de ganancia energética de Q = 10 y producir 500 MW para un pulso de

unos 100 s. Está siendo construido en Cadarache, Francia. Cortesía de ITER.

1.1. Daño por radiación en materiales para fusión

En los reactores de fusión tanto inercial como magnética, la radiación procede

de las partículas emergentes de la reacción de fusión: neutrones, iones y rayos

Page 21: Simulación atomística de la producción y evolución de ...

4

X, siendo la contribución neutrónica la más importante. De los neutrones

emitidos, sólo una pequeña fracción será absorbida por la primera pared, lo

que significa que la mayor parte de los 14MeV de energía serán transferidos a

los materiales estructurales. Aquí producirán defectos debidos a la radiación

que pueden alterar profundamente las propiedades de los materiales, tanto por

procesos inelásticos como por colisiones elásticas [7]. Los procesos inelásticos

incluyen interacciones con los electrones o transmutaciones de (n,α) o (n,p), en

los cuales se forman He o H (o renio en W). La cantidad de He e H que se

produce de este modo en un reactor de fusión se espera que sea alrededor de

un orden de magnitud mayor que en los reactores de fisión [8], provocando

problemas de hinchamiento debido a la formación de burbujas, así como

fragilización por la segregación de He en los bordes de grano. Por ejemplo, la

producción de He en Be cuando ITER sea sometido a un flujo de neutrones de

1 MWa/m2 es de 3500 partes atómicas por millón (appm) y en los aceros de

150 appm [9].

Otro gran problema son las colisiones elásticas entre los neutrones y los

núcleos atómicos, que resultan en cascadas de desplazamiento. El daño

resultante en la red cristalina consiste en defectos puntuales, clusters de

defectos, loops de dislocaciones, precipitados y/o agujeros (agregados de

vacantes). El daño se expresa a menudo en desplazamientos por átomo (dpa),

que describe con qué frecuencia cada átomo en el material es en promedio

desplazado de su posición en la red. Debido a la recombinación de defectos, el

número final de desplazamientos es a menudo sólo una pequeña fracción del

valor inicial de dpa.

Como ejemplo de defectos inducidos por la radiación, se muestran en la figura

2 las imágenes de microscopía electrónica de transmisión (TEM) del daño

producido en aleaciones de Fe con Cr por bombardeo de iones de Fe3+.

Page 22: Simulación atomística de la producción y evolución de ...

5

Antes de irradiar

Después de irradiar

Figura 2: Imágenes de TEM de una aleación de Fe-5Cr antes de ser irradiada y

después de la irradiación con iones de Fe3+ con una dosis de 0,45 dpa.

Cortesía de Anya Prokhodtseva, École Polytechnique Fédérale de Lausanne,

Centre de Reserches en Physique des Plasmas, 2011.

El aumento de defectos debido a la radiación muestra claramente cómo la

irradiación con partículas tiene un gran impacto en la microestructura del

material.

Puesto que las propiedades físicas y mecánicas de un material están

gobernadas por su microestructura, éstas pueden sufrir grandes cambios

debido al daño en la red. Por ejemplo, la agrupación de defectos formando

nanoestructuras de mayor tamaño, así como los precipitados, dificultan el

movimiento de las dislocaciones, provocando fragilización inducida por la

radiación [10]. Esto incluye endurecimiento, pérdida de ductilidad, resistencia a

Fe-5Cr 0.45 dpa 790 appm He Fe-5Cr 0.45 dpa

Page 23: Simulación atomística de la producción y evolución de ...

6

la fractura y creep por irradiación. Un cambio en la temperatura de transición

dúctil-frágil hacia temperaturas más altas es también común.

El conocimiento a nivel atómico de los defectos de la red cristalina es así pues

de vital importancia para poder predecir la respuesta a la irradiación de los

materiales estructurales.

1.2. Modelización del daño por radiación: modelos multiescala

Los fenómenos que ocurren en los materiales debido a la radiación tienen lugar

en diferentes escalas de tiempo y de espacio, son fenómenos jerárquicos y

típicos multiescala, desde la producción de defectos a escala atómica (unos 10-

15 s, 10 -10 m) hasta la evolución de la microestructura durante los años de

operación (unos 109 s, 10-3 m). Esto hace que la predicción a largo término de

los cambios inducidos por la radiación en las propiedades mecánicas de los

materiales se convierta en un desafío importante. Una aproximación consiste

en encadenar modelos existentes en las diferentes escalas de manera que los

resultados de un modelo sirven como datos de entrada para el siguiente

modelo de orden superior.

Seguidamente se analizan brevemente los procesos que tienen lugar a las

diferentes escalas arriba mencionadas y el modelo computacional utilizado

para su estudio, mostrados también en la figura 3. Para una información más

detallada de los procesos de daño por irradiación se pueden consultar las

siguientes referencias [11-16].

Page 24: Simulación atomística de la producción y evolución de ...

7

Figura 3: Diagrama de los fenómenos relevantes en daño por irradiación y los

métodos computacionales típicamente empleados [17]

1. Átomo primario de retroceso o PKA (Primary knock-on atom): El proceso

de daño comienza con la creación de un átomo de alta energía desplazado de

su posición original por la partícula incidente. Las secciones eficaces y los

modelos cinemáticos necesarios para el cómputo del espectro de PKAs están

implementados en códigos como SPECTER. SPECTER permite calcular el

espectro energético de PKAs producido por un neutrón de una determinada

energía.

2. Estructura de defectos: La estructura electrónica de defectos puntuales

se estudia con modelos de mecánica cuántica, denotados como primeros

principios, los cuales resuelven por aproximaciones la ecuación de Schrödinger

para un conjunto de átomos. Este método es computacionalmente muy costoso

por lo que el número de átomos que es posible estudiar es reducido. Códigos

como VASP [18] o SIESTA [19] son los más utilizados en el campo del daño

por radiación. Estos métodos de simulación se utilizan fundamentalmente para

obtener información sobre las energías de formación y de migración de

defectos puntuales y pequeñas agrupaciones de defectos.

3. Producción de defectos primarios en cascadas de desplazamiento: La

alta energía del PKA es rápidamente transferida a otros átomos mediante una

cadena de colisiones atómicas generando una cascada de vacantes y átomos

Page 25: Simulación atomística de la producción y evolución de ...

8

intersticiales. Este proceso se simula mediante códigos de dinámica molecular

(MD) que integran las ecuaciones del movimiento de millones de átomos. El

tiempo máximo de simulación que puede alcanzarse, sin embargo, es sólo de

nanosegundos.

4. Evolución de cascadas y migración de defectos de largo rango: Estudia

como los defectos migran e interactúan entre ellos modificando así la

microestructura. En general los átomos intersticiales son muy móviles migrando

a gran velocidad desde la zona de daño, mientras que las vacantes son más

lentas, agrupándose en la mayoría de los casos y permaneciendo cerca del

lugar en el que la cascada fue creada. Esta difusión se expande a lo largo de la

vida del material irradiado. En general, los coeficientes de difusión de las

especies que difunden son hallados mediante técnicas ab initio o con MD, y

utilizados por modelos de cinética de Monte Carlo (KMC) como el object kinetic

Monte Carlo (OKMC) para estudiar la evolución y acumulación de dichos

defectos. Del mismo modo es posible utilizar modelos cinéticos de ecuaciones

de tasas para estudiar la evolución de la microestructura y la difusión de

defectos. Grandes conjuntos de ecuaciones diferenciales se integran para

predecir la evolución de los defectos. Este método es mucho más eficiente que

los métodos Monte Carlo pero pierden la resolución atómica de estos.

5. Incremento en el stress de fluencia inducido por irradiación: las

heterogeneidades de la microestructura provocan un aumento substancial en el

estrés de fluencia (Δσy ) actuando como obstáculos dispersos para el

movimiento de las dislocaciones. Dicho estudio de la interacción entre una

dislocación y un obstáculo se han realizado tradicionalmente con modelos de

dinámica molecular [20].

6. Dinámica de Dislocaciones (DD): Estudia el movimiento de las

dislocaciones y su interacción entre ellas y con la posible introducción de

obstáculos como precipitados que modifican el mapa de estreses e impiden el

movimiento de dichas dislocaciones libremente. En este modelo las

dislocaciones son discretizadas resolviendo la formulación elástica y

avanzando la estructura a través de una función de movilidad que relaciona la

Page 26: Simulación atomística de la producción y evolución de ...

9

fuerza en un segmento de dislocación con su velocidad. Normalmente, esta

función de movilidad se obtiene mediante simulaciones con MD.

7. Curvas de cambio de la resistencia a la fractura con la temperatura:

modificaciones en las propiedades constitutivas resultan en una degradación

de la resistencia a la fractura del material que se manifiesta como un cambio

(ΔT) en la temperatura de transición dúctil-frágil.

Los modelos teóricos utilizados necesitan de la validación experimental de

cada uno de ellos, en las diferentes fases del proceso de daño que se

describen, para garantizar que ofrecen resultados fiables. Los datos

experimentales para validar los procesos a nivel microscópico proceden en

gran medida de la irradiación de los materiales con iones, ya que es mucho

menos costosa y más versátil que la irradiación con neutrones, y la posterior

caracterización microestructural del daño con diferentes técnicas, siendo la

Microscopía Electrónica de Transmisión (TEM) una de las más utilizadas.

1.3. Propósito de este trabajo

En el marco de la modelización multiescala del daño por irradiación descrita en

el apartado anterior un punto importante que ha sido resaltado por varios

autores [21] es que la evolución del daño a largo plazo es especialmente

sensible a la distribución inicial de defectos obtenida en la cascada de

desplazamiento. Esta importancia de la distribución del daño primario ya fue

resaltada en los años 80 en los primeros cálculos de dinámica molecular de

cascadas de desplazamiento [22, 23].

El objetivo principal de este estudio ha sido obtener, utilizando dinámica

molecular, Monte Carlo cinético y simulación de imágenes TEM, una

descripción detallada del daño primario producido al irradiar películas finas de

hierro puro con iones de hierro de entre 50 y 500 kev de energía. Las películas

finas tienen un espesor de entre 40 y 80 nm para reproducir el espesor de las

muestras utilizadas en las medidas de TEM durante los experimentos de

irradiación que se están llevando a cabo en Jannus (Francia) y los ya

Page 27: Simulación atomística de la producción y evolución de ...

10

publicados de Mercedes Hernández Mayoral y colaboradores [24]. En las

simulaciones se ha estudiado principalmente el efecto de las superficies en la

producción y evolución de daño y su comparación con la radiación en volumen

producida por neutrones. El estudio de las diferencias entre el daño que

produce un neutrón y el daño que produce un ión a nivel fundamental es clave

para la construcción de modelos predictivos de más alto nivel en el marco de la

simulación multiescala.

Page 28: Simulación atomística de la producción y evolución de ...

11

2. METODOLOGÍA

Para la realización de este trabajo se han utilizado tres métodos de simulación:

dinámica molecular con potenciales empíricos, Monte Carlo cinético y cálculo

de imágenes de TEM. En los siguientes apartados se describen brevemente

cada una de estas técnicas.

2.1. DINÁMICA MOLECULAR

La técnica de dinámica molecular consiste en el estudio de la evolución en el

tiempo de un sistema de N-cuerpos. Está basado en una interpretación

determinista de la naturaleza donde el comportamiento de un sistema se puede

realizar de forma computacional si conoces las condiciones iniciales y las

fuerzas de interacción.

La metodología de una simulación de dinámica molecular consiste en la

integración de las ecuaciones de movimiento para todos los átomos en una

celda computacional. Las trayectorias de las partículas se obtienen de la

integración de la ecuación de Newton:

(1)

Para un potencial conservativo la fuerza es una función de las coordenadas y

se puede obtener a partir del gradiente del potencial:

(2)

La integración numérica de estas ecuaciones nos proporciona la trayectoria de

las partículas, siendo el tiempo el paso de integración. El paso de tiempo se

elige para minimizar los errores y maximizar la eficiencia computacional.

Page 29: Simulación atomística de la producción y evolución de ...

12

Normalmente es del orden de 0.5-1.0 femtosegundo (fs), ya que debe ser

menor a la frecuencia de vibración de los átomos del material.

Aunque en principio es un método muy simple de implementar, sus

fundamentos residen en:

Mecánica clásica,

Dinámica clásica no lineal,

Teoría cinética,

Mecánica estadística,

Principios de conservación y

Física del estado sólido.

Las partes fundamentales de un código de dinámica molecular son el potencial

interatómico, el algoritmo de integración y las condiciones de contorno.

2.1.1. Potenciales de interacción

En la simulación de un material real por dinámica molecular el potencial

interatómico es la parte fundamental. Es en él donde está incluida toda la

información sobre la interacción entre las partículas y la precisión con la que

reproduzca las propiedades del material estudiado determina en su mayor

parte la fiabilidad de la simulación. El principal reto en la construcción de

potenciales es hacerlos suficientemente simples para permitir una velocidad de

cálculo aceptable, y a su vez capaces de reproducir fielmente las propiedades

del material. Los potenciales más extendidos en la actualidad son de carácter

semiempírico.

El desarrollo empieza buscando una forma del potencial adecuada que

depende de una cierta cantidad de parámetros. A continuación estos

parámetros se ajustan a una base de datos relativamente extensa. La base de

datos puede contener propiedades calculadas experimental o teóricamente. El

cálculo teórico de propiedades se lleva a cabo por primeros principios.

Dependiendo del fenómeno que se quiera simular, se valora en mayor o menor

Page 30: Simulación atomística de la producción y evolución de ...

13

medida la calidad del ajuste a cada propiedad. Es importante destacar que, el

hecho de reproducir fielmente una gran cantidad de propiedades no garantiza

que el valor calculado de una propiedad nueva sea el correcto.

2.1.2. El modelo del átomo embebido

En el caso de los metales, el modelo del átomo embebido (EAM de embedded

atom method) es el potencial de uso más extendido [27]. Se basa en la idea de

que la energía de un átomo en un sólido cristalino es igual a la energía

necesaria para embeberlo en la nube electrónica de la red. En otras palabras,

la energía de cohesión de un átomo en la red es un funcional de la densidad

electrónica en ese punto de la red. Si ρi es la densidad electrónica en torno al

átomo i, su valor se obtiene por la superposición de las densidades electrónicas

de los N átomos del sistema en la posición del átomo i.

(3)

Existen varias parametrizaciones del EAM. Originalmente, la forma funcional de

las densidades electrónicas ρa se tomaba de cálculos previos de las funciones

de onda de Hartree-Fock:

(4)

Donde ns y nd son el número de electrones s y d externos, y ρs y ρd son las

densidades asociadas con las funciones de onda de los orbitales s y d.

Conocida la densidad electrónica en la posición de cada átomo, la expresión de

la energía total de un sistema de N átomos es la suma de las energías de

embebimiento de todos los átomos, más la suma de la suma de las repulsiones

nucleares (que se representa mediante un potencial de pares)

(5)

Page 31: Simulación atomística de la producción y evolución de ...

14

En su forma más general, suponiendo la repulsión electrostática entre dos

núcleos, uno de un átomo tipo A y otro de un átomo tipo B, φ toma la forma

(6)

Donde Z es la carga efectiva, una función de la distancia R. La forma funcional

de la energía de embebimiento F no se puede expresar de forma analítica.

Para determinarla, se usa la ecuación de estado universal, descrita por Rose,

en la que la energía de cohesión de un átomo puede expresarse en función de

la energía de sublimación Esub y la constante de red a

(7)

Donde

(8)

Siendo B el bulk modulus, Ω el volumen atómico y Esub la energía de

sublimación, todos ellos valores de equilibrio. Entonces, basándose en la

fórmula (7), la función de embebimiento se puede tabular variando el parámetro

de red.

Otra variante de los potenciales EAM es el potencial interatómico de Finnis-

Sinclair [28]. Esta familia de potenciales parte del modelo EAM y lo deriva

utilizando una aproximación de segundo orden de la teoría de tight-binding de

sólidos [31]. Desde el punto de vista de cálculo, es más simple todavía, puesto

que la función F es una raíz cuadrada.

Page 32: Simulación atomística de la producción y evolución de ...

15

Durante los últimos años, el EAM ha demostrado ser un potencial con

excelentes prestaciones en el caso de los metales fcc [29].

Desafortunadamente, la forma funcional del potencial predice

empaquetamientos compactos, por lo que no es un buen candidato para las

simulaciones de metales bcc. EAM es incapaz de describir la fuerte

direccionalidad de los enlaces en los metales bcc. En cualquier caso, a falta de

mejores alternativas, sigue siendo enormemente utilizado en metales bcc

también.

Los potenciales utilizados en este trabajo, en el que el metal simulado (α-Fe)

tiene un empaquetamiento bcc, han sido el potencial de Dudarev y Derlet (DD)

[30] y el potencial de Ackland, Mendelev y Srolovitz et al. (AM) [33]. Ambos

potenciales fueron modificados en el rango de distancias cortas, esto es, en la

parte repulsiva del potencial, de acuerdo con el procedimiento descrito en [32].

Fueron elegidos porque reproducen con bastante exactitud la estabilidad de

diferentes defectos puntuales en comparación con cálculos obtenidos con la

teoría del funcional de la densidad (DFT). El potencial AM es un potencial multi-

cuerpo de tipo Finnis-Sinclair y está ajustado a valores de ab initio de energías

de defectos puntuales y de propiedades de bulk [33]. El potencial DD es

también un potencial multi-cuerpo de tipo Finnis-Sinclair, pero incluye los

efectos del magnetismo en la energía de interacción de los átomos de Fe

utilizando los modelos de Stoner y de Ginzburg-Landau [30].

2.1.3. Métodos de integración

La estabilidad del sistema, exactitud del cálculo y eficacia del código

determinan la elección de uno u otro algoritmo. En general, estos algoritmos

resuelven las ecuaciones de movimiento de un sistema calculando las

posiciones de las partículas como una función del tiempo. Suelen asumir que la

fuerza aplicada sobre cada partícula se mantiene constante durante el

incremento temporal Δt. Las condiciones ideales de un buen algoritmo son:

Rápido en el cálculo,

Page 33: Simulación atomística de la producción y evolución de ...

16

Estable con incrementos temporales relativamente grandes,

Buena conservación de energía en el conjunto microcanónico.

El algoritmo utilizado en este trabajo es el predictor-corrector de cuarto grado.

Dadas la posición, velocidades, y demás información en el instante t, este

algoritmo calcula las nuevas posiciones, velocidades y demás en el instante t +

δt. Como la trayectoria es continua, estas cantidades se pueden expresar en

forma de serie de Taylor.

(9)

(10)

(11)

(12)

Donde el superíndice p significa predicho. El siguiente paso es calcular las

nuevas fuerzas para los valores predichos de las posiciones, con lo que se

obtienen las aceleraciones corregidas . Por lo tanto, el error en la

predicción es

(13)

Entonces, este error se incluye en los valores predichos de las posiciones,

velocidades y demás, para calcular los valores corregidos

(14)

(15)

(16)

Page 34: Simulación atomística de la producción y evolución de ...

17

(17)

Los valores óptimos de los coeficientes c0, c1, c2 y c3 para conseguir máximas

estabilidad y precisión en las trayectorias fueron obtenidos por Gear. En la

tabla 1 se muestran los valores para diferentes órdenes de truncamiento de la

serie de Taylor.

Orden C0 C1 C2 C3 C4 C5

3 0 1 1

4 1/6 5/6 1 1/3

5 19/20 ¾ 1 ½ 1/12

6 3/20 251/360 1 11/18 1/6 1/60

Tabla 1: Coeficientes para el algoritmo predictor-corrector

El número de pasos correctores puede ser el que se desee aunque, por

cuestiones de eficiencia computacional, sólo se suele usar un paso corrector.

Este es el algoritmo que suele ofrecer la mejor conservación de energía en el

conjunto microcanónico. Por lo tanto, es muy utilizado siempre que el número

de pasos de integración sea alto. Un criterio para la selección del paso

temporal es considerar que Δt << 1/ donde es la frecuencia de Debye (1013

s-1). El valor elegido, exceptuando aplicaciones de alta energía, suele ser del

orden del femtosegundo. De esta forma, el error en la conservación de energía

se mantiene inferior a 0.1 eV/ns.

2.1.4. Condiciones de contorno

Las condiciones de contorno más simples son considerar superficies libres.

Pero por lo general, interesa simular sólidos macroscópicos, por lo que se

recurre a las condiciones periódicas de contorno. Estas consisten en la

repetición de la celda MD básica infinitas veces en las tres direcciones del

espacio. Cada una de las réplicas de la celda básica recibe el nombre de celda

imagen. Las celdas imagen son del mismo tamaño y contienen las mismas

Page 35: Simulación atomística de la producción y evolución de ...

18

partículas que la celda original. Considerando una celda MD de lado L con N

partículas, las ecuaciones básicas que definen las condiciones periódicas de

contorno son

(18)

Donde i´ es la partícula imagen de i en la celda de índices (n1,n2,n3).

Centrándose ahora en cómo afectan las condiciones periódicas a la celda

original, suponiendo una celda de simulación cúbica, toda partícula que

abandone la celda reentra por la cara opuesta con la misma velocidad (figura

4).

Por lo tanto, las condiciones periódicas de contorno conservan el número de

partículas y la cantidad de movimiento, pero no el momento cinético (aunque sí

su valor medio).

Figura 4: esquema del funcionamiento de las condiciones periódicas [34].

Page 36: Simulación atomística de la producción y evolución de ...

19

Además, las condiciones de contorno periódicas suponen que para calcular las

fuerzas sobre un átomo en el borde de la celda principal, se consideran las

celdas imágenes, de forma que no hay átomos de superficie, todos los átomos

está rodeados de otros átomos en las tres dimensiones. De esta forma se

simula un sólido infinito.

En este trabajo, para simular la irradiación con iones de las películas finas

analizadas con TEM, se han utilizado condiciones periódicas únicamente en las

direcciones del espacio x e y, dejando en el eje z las superficies libres.

2.1.5. El código MDCASK

El código MDCASK, dedicado específicamente a simular cascadas de

desplazamiento, ha sido el empleado en este estudio. Está basado en el código

MOLDY desarrollado por Finnis y colaboradores en el laboratorio Harwell del

Reino Unido [35]. Este programa ha sido implementado con características

especiales para estudios de implantación de iones de energías entre eV y keV

en metales y blancos de silicio. Además ha sido modificado para funcionar en

todo tipo de plataformas en paralelo.

Su implementación en paralelo fue realizada en el laboratorio Lawrence

Livermore National Laboratory de California en los EEUU. Para el cálculo de

vecinos de cada uno de los átomos utiliza el algoritmo de las listas enlazadas

[36] que se explica a continuación.

En cualquier código de dinámica molecular, la mayor parte del tiempo se

consume en calcular las distancias de cada átomo a sus vecinos. Suponiendo

potenciales de alcance infinito, el problema sería de orden N2.

Afortunadamente, al ser los potenciales de alcance limitado, no es necesario el

cálculo de todas las distancias. Por potencial de alcance limitado se entiende

aquel en el que la posición de un átomo i no influye en la fuerza sobre el átomo

j si la distancia que los separa (rij) es mayor que un cierto radio de corte rc.

Page 37: Simulación atomística de la producción y evolución de ...

20

Entonces, para calcular la fuerza en el átomo j sólo se necesita calcular su

distancia con todos los átomos dentro de una esfera de radio rc. En la figura 4,

la energía del átomo 1 depende sólo de la posición de los átomos en el círculo

oscuro. El algoritmo de las listas enlazadas es el algoritmo que menos

distancias innecesarias calcula y organiza las listas de vecinos de forma

eficiente. Este método divide la celda de simulación en pequeñas celdas de

arista ligeramente superior a rc. De esta forma, se garantiza que dos átomos

que no estén en celdas adyacentes no interaccionan entre sí. Entonces,

cuando se quiera calcular los vecinos de un átomo, basta con calcular las

distancias con los átomos de su propia celda y las adyacentes.

2.1.6. Simulación de irradiación de Fe con iones de Fe: antecedentes

El concepto de cascada de desplazamiento como proceso fundamental en la

producción de daño por irradiación data de los años 50, cuando se introdujeron

por primera vez los términos displacement spike [37] y thermal spike [38]. En

esos años ya se conocían bastante bien los aspectos básicos de este

fenómeno, aunque de forma cualitativa [39]. Pronto se reconoció que para

obtener más información sobre este proceso – que experimentalmente es

imposible de observar – el uso de métodos numéricos era especialmente

adecuado. Las primeras simulaciones de colisiones en cristales mediante

técnicas computacionales se hicieron en cobre con estructura fcc [40] y poco

después, debido a su importancia en aplicaciones prácticas, en hierro bcc [41].

La aproximación de colisiones binarias (BCA), desarrollada en paralelo a lo que

ahora conocemos como dinámica molecular (MD), era en esos tiempos la única

posibilidad para explorar eventos de energías relativamente altas [42], por lo

que permaneció como técnica de referencia hasta que los avances en la

ciencia computacional de los años 80 permitió el uso generalizado de la MD,

demostrado el hecho de que BCA predecía una cantidad de defectos

demasiado alta comparada con los experimentos [43].

Desde entonces se han realizado una gran cantidad de simulaciones con MD

de irradiación de iones en α-Fe, pero debido a su intención de simular el daño

Page 38: Simulación atomística de la producción y evolución de ...

21

por irradiación de neutrones en los materiales del reactor, prácticamente todos

los cálculos se han realizado en el interior del material aplicando condiciones

periódicas en las tres direcciones del espacio. Los efectos de la irradiación en

el interior de un material son sin embargo muy difíciles de observar

experimentalmente, y de hecho, la mayoría de técnicas como la microscopía

electrónica de transmisión (TEM), examinan defectos cercanos a las

superficies. Trabajos recientes [44-50] indican que la presencia de superficies

cercanas puede influir en la formación del daño primario, por lo que las

simulaciones con sólidos infinitos no se corresponden con los experimentos de

irradiación con iones de las películas finas observadas por TEM. De estos

trabajos, los únicos realizados en Fe al comienzo de la tesis [48, 49] habían

utilizado potenciales interatómicos que han demostrado tener una configuración

errónea para los átomos intersticiales, además de emplear energías muy bajas

(10kev). La motivación de este trabajo de tesis ha sido por tanto estudiar

mediante simulación multiescala las diferencias entre la irradiación en láminas

finas de Fe y en el interior del material de manera detallada y cuantitativa,

utilizando potenciales interatómicos recientes y condiciones homólogas a las

experimentales.

2.2. MONTE CARLO CINÉTICO

Las simulaciones de Monte Carlo se caracterizan por utilizar números

aleatorios en sus algoritmos. Su nombre se debe precisamente al famoso

casino de la ciudad de Mónaco. El desarrollo sistemático del método comenzó

en el Laboratorio Nacional de Los Alamos en los años 40, con los trabajos de

Ulam y Metropolis [51], desarrollándose ampliamente desde entonces. Hoy en

día los métodos de Monte Carlo más usados son el Metrópolis Monte Carlo [52]

y el Monte Carlo Cinético [53]. El método de Metrópolis Monte Carlo se

desarrolló en los cincuenta para estudiar propiedades en equilibrio de un

sistema. Diez años más tarde, en los sesenta, se introdujo el método de Monte

Carlo Cinético que, a diferencia del Metrópolis Monte Carlo, permite estudiar

propiedades dinámicas de un sistema.

Page 39: Simulación atomística de la producción y evolución de ...

22

En nuestro problema de simular una cascada de desplazamiento hemos

utilizado dinámica molecular clásica. La dinámica molecular sigue de manera

natural la evolución de un sistema de átomos a través de la propagación en el

tiempo de las ecuaciones clásicas del movimiento. Sin embargo, como se ha

comentado anteriormente, la integración exacta de las ecuaciones de Newton

requiere pasos de tiempo lo suficientemente cortos (~10-15 s) como para

resolver las vibraciones atómicas. Esto hace que el tiempo total de simulación

que se puede alcanzar con Dinámica Molecular Clásica sea menor de un

microsegundo. El método de Monte Carlo Cinético salva esta limitación

utilizando el hecho de que la dinámica de este tipo de sistemas consiste

típicamente en procesos de difusión (saltos atómicos) que ocurren en escalas

de tiempo mucho mayores. De este modo, se puede asumir que la evolución de

un sistema se caracteriza por transiciones ocasionales de un estado a otro,

donde cada estado corresponde a un pozo de potencial o mínimo local y dos

estados adyacentes están separados por una barrera de energía, tal como

indica la figura 5.

Figura 5: Contorno esquemático de una superficie de energía potencial para la

transición de un sistema entre dos estados. Después de muchos períodos de

vibración, el sistema pasa de un estado a otro. Los puntos indican las barreras

de energía [54].

Este tipo de sistemas se catalogan dentro de los llamados “infrequent-event

systems”. Para que un suceso sea considerado como “rare event” la barrera de

Page 40: Simulación atomística de la producción y evolución de ...

23

energía potencial tiene que ser mucho mayor que KBT. La propiedad clave de

este tipo de sistemas es que, después de una transición, y debido al largo

tiempo de permanencia en cada estado, la partícula pierde la memoria de cómo

llegó hasta allí. Esto hace que la probabilidad de escape hacia un nuevo estado

sea independiente de los anteriores eventos y sólo dependa del estado actual y

del posible nuevo estado adyacente. Esta característica es lo que define a los

procesos Markovianos [55] y permite obtener a priori todas las probabilidades

de transición. En los procesos estocásticos las transiciones se describen a

menudo con procesos de Poisson [55, 56]. Si consideramos un suceso con una

frecuencia r, la densidad de probabilidad de la transición f(t) que nos dará la

probabilidad de que la transición ocurra a un tiempo t será:

(19)

Generalizando a N procesos independientes de Poisson, con frecuencia ri,

obtenemos un proceso de Poisson con una frecuencia acumulada

(20)

y una densidad de probabilidad de la transición igual a:

(21)

Podemos determinar el tiempo medio entre eventos sucesivos:

(22)

La evolución de un sistema se puede describir eligiendo al azar eventos con

una probabilidad proporcional a su frecuencia tal como muestra la figura 6.

Page 41: Simulación atomística de la producción y evolución de ...

24

Figura 6. Esquema del procedimiento para elegir el camino de reacción.

Los eventos se colocan uno al lado del otro con un tamaño proporcional a su

frecuencia. A continuación se genera un número aleatorio s en el intervalo (0,1)

y se elige el evento cuya frecuencia cumple:

(23)

Para actualizar el reloj del sistema, el tiempo se debe incrementar en (el

tiempo medio entre eventos sucesivos). De manera general, el reloj del sistema

se puede incrementar en un tiempo aleatorio , de acuerdo a la ecuación:

(24)

Donde s’ es un número aleatorio en el intervalo (0,1). El uso de números

aleatorios para evolucionar el tiempo de un sistema da una mejor descripción

de la naturaleza estocástica de los procesos que intervienen y se justifica

porque . Este procedimiento es el más usado en Monte Carlo

Cinético y se conoce como el algoritmo del tiempo de residencia [57].

r1

r2

r3

rn

...

sR

Page 42: Simulación atomística de la producción y evolución de ...

25

2.2.1. La teoría del estado de transición

Para calcular las frecuencias de transición entre un estado i y un estado j, el

método de Monte Carlo Cinético se basa en la aproximación armónica del

formalismo de la teoría del estado de transición [58, 59]:

(25)

Donde Pij es el llamado prefactor, y Ea es la energía de activación del proceso

tal como muestra la figura 7. Los prefactores normalmente se encuentran en el

rango de 1012 s-1 – 1013 s-1, por lo que una práctica común es utilizar un valor

fijo en este rango para ahorrar tiempo de computación. Las energías de

activación se obtienen de cálculos de DFT o en su defecto de dinámica

molecular.

Figura 7. Esquema de la cinética de transición de un sistema entre dos

estados. En la teoría del estado de transición la reacción se produce siempre

del estado A al estado B.

Page 43: Simulación atomística de la producción y evolución de ...

26

2.2.1. El código MMonCa

En el presente trabajo se ha utilizado el código de Monte Carlo Cinético

MMonCa [60]. MMonCa es un simulador multi-material creado por el Dr.

Ignacio Martín-Bragado en el Instituto IMDEA Materiales de Madrid. Está

escrito en C++ y ha sido integrado en la librería TCL para la interfaz de usuario.

Contiene dos módulos independientes para conseguir diferentes niveles de

simulación:

Lattice kinetic Monte Carlo (LKMC). Este módulo utiliza la red del

material y se usa por ejemplo para simular cambios de fase tales como la

recristalización epitaxial en estado sólido, donde la orientación cristalina es

fundamental.

Object kinetic Monte Carlo (OKMC). Este módulo no utiliza la red del

material (off-lattice en inglés) y se usa para estudiar la evolución de defectos en

un sólido.

La estructura simplificada de MMonCa se muestra en la figura 8. El usuario

lanza una simulación a través de un script escrito con comandos TCL y otros

comandos especiales para inicializar una simulación, simular un recocido, etc.

Page 44: Simulación atomística de la producción y evolución de ...

27

Figura 8. Estructura por bloques del simulador MMonCa.

En esta tesis hemos utilizado el módulo OKMC para estudiar la evolución del

daño implantado en las simulaciones de dinámica molecular en láminas finas y

en bulk de Fe irradiando con iones de Fe de 100 keV. En OKMC los defectos

se definen como objetos. En particular, en MMonCa se definen como

interfases, partículas libres, clusters, defectos extendidos y multiclusters. Para

cada uno de estos objetos es necesario definir el número de eventos asociados

con el objeto, la frecuencia asociada a cada evento, y las funciones para llevar

a cabo el evento una vez es elegido por el algoritmo de MMonCa.

User input script

TCL

library

KMC kernel:Space managerTime manager

LKMC OKMC

Operating System

User

interface C+

+

Page 45: Simulación atomística de la producción y evolución de ...

28

2.3. SIMULACIÓN DE IMÁGENES TEM

La simulación de imágenes de microscopía electrónica de transmisión (TEM)

se utiliza, entre otros motivos, para relacionar las imágenes de dinámica

molecular con las imágenes reales de experimentos de irradiación en

materiales observados por TEM. Como se ha comentado anteriormente, la

mayor limitación de la dinámica molecular es la escala de tiempo en la que

opera, la cual alcanza como máximo los nanosegundos, mientras que la escala

real de laboratorio comienza en los segundos. Esto hace imposible una

correlación directa entre la dinámica molecular y los experimentos. La

simulación de imágenes TEM de las configuraciones obtenidas con dinámica

molecular permite relacionar ambas técnicas. Así, es posible, por ejemplo,

averiguar si los clusters de defectos obtenidos en dinámica molecular serán

estables y visibles en la escala real de los experimentos, identificar la

naturaleza (intersticial o vacante) de defectos observados en el microscopio, o

saber si los clusters observados se han formado en la cascada de

desplazamiento o mediante migración y recombinación de defectos.

2.3.1. El método multicapa

El método multicapa [61] es el más usado en los programas de simulación de

imágenes TEM y consiste básicamente en seccionar la muestra en múltiples

capas perpendiculares al haz incidente. Para ello existen varias

aproximaciones entre las que destacan:

El formalismo del espacio recíproco.

El formalismo FFT.

La aproximación del espacio real.

La aproximación de las funciones de Bloch.

El fundamento de todos ellos es el mismo. En primer lugar, el potencial del

cristal se divide en capas y se proyecta en un plano perpendicular a la dirección

Page 46: Simulación atomística de la producción y evolución de ...

29

de observación. Al potencial proyectado en cada capa se le denomina “phase

grating” o “rejilla de fase”. A continuación se calculan las amplitudes y fases de

todos los haces generados por la interacción de la función de onda del haz de

electrones incidente con el primer plano proyectado. Estos haces se propagan

por el vacío (el microscopio virtual) hasta la segunda capa y se repite el

proceso. Así hasta la última capa de la muestra (figura 9).

Figura 9 Representación gráfica de la metodología multicapa para obtener

imágenes TEM simuladas [62].

Principalmente el método multicapa considera tres componentes:

Ψ describe la función de onda del electrón.

P es la propagación de la función de onda electrónica en el vacío: el

microscopio.

Q es el “fase grating” o “rejilla de fase”: la muestra.

En el formalismo del espacio recíproco este proceso se puede describir por la

ecuación:

(26)

Page 47: Simulación atomística de la producción y evolución de ...

30

Donde es la función de onda en el espacio recíproco a la salida de la

capa n+1, es la función de propagación, es la función de

transmisión, y (x) es la convolución. Las tres funciones descritas son funciones

del espacio recíproco, de ahí el nombre del formalismo. En el formalismo FFT

la ecuación (26) se transforma en:

(27)

Donde es la transformada de Fourier de la función en el

espacio real.

2.3.2. El código EMS

El programa que hemos utilizado en esta tesis, el código EMS [63] utiliza el

formalismo FFT para maximizar la eficiencia computacional del método

multicapa. La figura 10 representa esquemáticamente la implementación del

método en el programa EMS.

Page 48: Simulación atomística de la producción y evolución de ...

31

Figura 10. Esquema de la implementación del método multicapa en el software

EMS [64].

Con este software hemos simulado imágenes TEM de campo claro de clusters

de defectos a 200 kV. La muestra de dinámica molecular se divide

perpendicularmente al haz de electrones en láminas de 0.2 nm de espesor. Las

láminas han de ser periódicas para realizar la transformada de Fourier. Para

cumplir con este requisito se elige una dirección cristalográfica para realizar el

corte, la cual determina la condición de difracción. A continuación se obtiene la

función de onda de cada una de las láminas, y finalmente, se obtiene la imagen

TEM simulada haciendo pasar cada función de onda por el “microscopio

virtual”, un subprograma del código EMS.

Page 49: Simulación atomística de la producción y evolución de ...

32

3. RESULTADOS Y DISCUSIÓN

En este apartado se resumen brevemente los principales resultados del trabajo

realizado durante la tesis. Estos resultados se discuten en mayor profundidad

en los trabajos presentados en los apartados 4 y 5.

3.1. Daño en escala de picosengundos: iones frente a neutrones

En los trabajos I y II se muestra que el daño por irradiación en thin films es

totalmente diferente al daño en bulk. En el trabajo III se cuantifican las

diferencias y se muestra la formación de loops visibles de vacantes <100>

directamente en la cascada de desplazamiento. Estos resultados explican los

experimentos de irradiación in-situ de láminas finas de Fe y FeCr a bajas dosis

donde se observa la formación de este tipo de loops.

En el trabajo I estudiamos con dinámica molecular el daño producido al irradiar

láminas finas de Fe puro con iones de Fe de 50, 100 y 150 keV bajo

condiciones de channeling y con el potencial de DD. Los resultados se

comparan con los resultados de Björkas [65] en cascadas de bulk. En este

trabajo obtenemos dos conclusiones importantes. En primer lugar mostramos

que, a diferencia de las cascadas en bulk que producen el mismo número de

vacantes e intersticiales, en las cascadas en láminas finas, aún en condiciones

de channeling, el número de vacantes es siempre mayor al número de

intersticiales. Esto es debido a la morfología del daño producido en las

cascadas en láminas finas. La segunda conclusión importante es que, debido a

esta morfología que produce diferentes tipos de eventos según dónde colisione

el PKA con mayor energía, la dispersión en el número de defectos en láminas

finas es mucho mayor que en las simulaciones en el bulk debido a la influencia

de las superficies. Además esta dispersión aumenta con la energía del PKA.

En el trabajo II nos centramos en las cascadas de mayor energía, 150 keV, y

estudiamos el efecto del ángulo de incidencia del ión y la morfología de la

cascada, también con el potencial de DD. El resultado más relevante es que,

por el contrario de lo ocurre a menores energías donde sólo se daña la lámina

Page 50: Simulación atomística de la producción y evolución de ...

33

superior, a esta energía y por extrapolación a energías superiores, tanto la

lámina superior como la inferior sufren un daño considerable (Figura 11).

Figura 11. Las figuras (a-c) muestran tres cascadas representativas de

irradiación de Fe por un ión de Fe de 150 keV al cabo de 25 ps. Las esferas

verdes son vacantes, las esferas rojas son intersticiales y las amarillas átomos

de superficie. Se representan ambas superficies. La flecha indica la posición de

implantación del átomo de Fe. En (a) y (c) el átomo de Fe impacta con un

ángulo de 10º y en (b) el ángulo de implantación es de 22º respecto a la

normal. La figura (d) es una imagen aumentada del daño producido en la

superficie inferior del caso (c). Del trabajo II.

En el trabajo III hacemos un estudio estadístico detallado comparando en

paralelo cascadas en láminas finas y cascadas de bulk con dos potenciales

distintos, el de DD y el AM, dos ángulos distintos, 10º y 22º, y dos energías

distintas, 50 keV y 100 keV. En este trabajo cuantificamos las diferencias entre

las dos geometrías y estudiamos en detalle la morfología de las cascadas. La

diferencia más remarcable es la creación de loops de vacantes <100>. Tanto

su tamaño como su frecuencia son mucho mayores en las cascadas en

láminas finas, con loops que alcanzan los 4 nm de tamaño. Además, los

Page 51: Simulación atomística de la producción y evolución de ...

34

clusters de intersticiales producidos en la cascada son en promedio más

pequeños en las láminas finas, especialmente en los casos en los que se

produce un cluster de vacantes de gran tamaño (Figura 12).

Figura 12. Microestructura de dos cascadas de 50 keV de Fe en Fe a los 25 ps.

En ambos casos el ángulo de impacto es de 10º pero en (a) el ángulo acimutal

es de 10º y en (b) de 80º. La distribución inicial de velocidades es también

distinta. Las esferas rojas son intersticiales, las verdes vacantes y las amarillas

átomos de superficie. Del trabajo III.

En el trabajo V comparamos los resultados de dinámica molecular de daño en

Fe con el daño que se produce en el otro material clave en fusión, el W.

Encontramos que en ambos materiales el tamaño de los defectos sigue una ley

de escala tanto en bulk como en láminas finas. Sin embargo, la pendiente de la

ley de escala en Fe bulk es marcadamente diferente, y se ve afectada por la

presencia de superficies en las irradiaciones en láminas finas. Esto explica las

diferencias en los experimentos de TEM en Fe y W. En W bulk las cascadas

son más compactas y con mayor densidad de energía que las cascadas en Fe

bulk, por lo que en W bulk se producen clusters grandes tanto de vacantes

como de intersciales, que no se producen en Fe bulk. El efecto de las

superficies en ambos materiales es el mismo, pero debido al diferente

comportamiento en bulk, este efecto es mucho más evidente en Fe.

En el trabajo VI estudiamos las diferencias entre cascadas de dinámica

molecular y cascadas de colisiones binarias (BCA) en simulaciones de

Page 52: Simulación atomística de la producción y evolución de ...

35

irradiación de Fe con neutrones, y la evolución de la microestructura con

OKMC. En cuanto a la morfología de las cascadas, en ambos casos el número

de pares de Frenkel generado aumenta con la energía del PKA de manera

proporcional, y las cascadas tienen el mismo rango de longitud. Sin embargo,

las cascadas de BCA, debido a que no sufren recombinación, producen un

mayor número de defectos que las cascadas de dinámica molecular, y

muestran una distribución espacial diferente. Las consecuencias de estas

diferencias se discuten en el siguiente apartado.

3.2. Evolución del daño primario en Fe y en FeCr

La influencia de las superficies en la evolución del daño en Fe irradiado se

muestra en el trabajo IV. Utilizamos como datos de entrada las cascadas de

dinámica molecular de 100 keV en láminas finas y en bulk irradiadas con iones

de Fe con un ángulo de 22º. Para las energías de enlace y de migración de

defectos utilizamos datos de cálculos de DFT y asumimos que los loops de

intersticiales <100> se forman por reacción de dos loops de intersticiales <111>

de tamaño similar. En este trabajo nos centramos principalmente en los efectos

de los obstáculos y de la distribución inicial del daño en la cascada. Los

principales resultados son, por un lado, que el efecto de la superficie depende

en gran medida de la concentración de obstáculos, y por otro, y aún más

interesante, que la distribución del daño inicial es clave tanto para la evolución

de la concentración de defectos como para su naturaleza. En este modelo la

irradiación cercana a la superficie favorece la formación de loops <100>.

En el trabajo VII abordamos el problema de la nucleación y crecimiento de

loops en Fe y en FeCr bajo irradiación mediante OKMC. Utilizamos como en el

trabajo IV datos de DFT para la energética de los defectos y las cascadas de

100 keV en láminas finas con el potencial de AM. Para la formación de los

loops de intersticiales <100> cosideramos dos posibles modelos. Uno de ellos

es el utilizado en el trabajo IV, según el cual los loops de intersticiales <100> se

producen por reacción de dos loops <111> se tamaño similar (modelo A), y un

segundo modelo, según el cual los clusters pequeños de intersticiales pueden

transformarse en loops <111> o <100> al crecer (modelo B). La principal

Page 53: Simulación atomística de la producción y evolución de ...

36

conclusión es que el modelo B reproduce más fielmente los resultados

experimentales, ya que reproduce la tendencia de crecimiento de los loops con

la dosis y la proporción de loops <100> y <111>.

3.3. Comparación con imágenes TEM

Finalmente la comparación con imágenes TEM no ha dado lugar a un trabajo

preparado para publicar en la tesis, de manera que se discutirán en este

apartado los principales resultados obtenidos hasta el momento.

Hemos realizado imágenes TEM simuladas de las cascadas producidas por la

irradiación de Fe en Fe en láminas finas con energías de 30 keV, 50 keV y 100

keV. Como se ha explicado en la metodología el software empleado ha sido el

código EMS desarrollado por Robin Schaublin y las simulaciones se han

llevado a cabo en el cluster Atsimat del departamento de Física Aplicada de la

Universidad de Alicante. Cada imagen simulada de toda la muestra, unos 50

nm, necesita hasta 1500 horas de CPU.

La figura 13 muestra la imagen TEM simulada de una cascada de Fe en Fe de

30 keV. También se muestra la posición de todos los átomos en la superficie,

así como la localización de los defectos. Los cálculos revelan que incluso para

esta energía tan baja, el daño producido puede observarse experimentalmente

ya que sus dimensiones alcanzan unos pocos nanómetros.

Page 54: Simulación atomística de la producción y evolución de ...

37

Figura 13. Imagen TEM simulada (izquierda) de una cascada de dinámica

molecular de 30 keV de Fe en Fe (derecha). La figura muestra la posición de

todos los defectos y de ambas superficies a los 25 ps del inicio de la cascada.

Las esferas verdes son vacantes, las rojas intersticiales y las amarillas átomos

de la superficie.

Hemos realizado el mismo tipo de cálculo para energías más altas, 50 keV y

100 keV. En la figura 14 se muestra el caso de una cascada de 50 keV. En este

caso se forman dos loops de vacantes bajo la superficie, uno mucho mayor que

el otro. En la imagen TEM simulada a los 7 nm de espesor se observan dos

puntos brillantes que muestran la localización de estos loops de vacantes, con

dimensiones de unos pocos nanómetros. Sin embargo, en la imagen TEM

simulada de toda la muestra sólo se observa un único punto, lo cual significa

que el contraste del loop de mayor tamaño eclipsa el contraste del de menor

tamaño, hecho corroborado experimentalmente.

Page 55: Simulación atomística de la producción y evolución de ...

38

Figura 14. Imágenes TEM simuladas (superior) de una cascada de dinámica

molecular de 50 keV de Fe en Fe (inferior). La figura muestra la posición de

todos los defectos y de la superficie superior a los 35 ps del inicio de la

cascada. Las esferas verdes son vacantes, las rojas intersticiales y las

amarillas átomos de la superficie. El loop de vacantes tiene unos 5 nm de

diámetro y el daño en la celda se extiende hasta los 11 nm de espesor.

Además de las imágenes TEM simuladas, realizamos experimentos de

irradiación de Fe ultrapuro con átomos de Ga de 30 keV utilizando un sistema

FIB (Focused Ion Beam) y una dosis entre 0.02 iones/nm y 0.32 iones/nm. El

análisis de TEM de las muestras muestran la formación de defectos de tamaño

considerable cerca de la superficie, como los mostrados en la Figura 15.

Page 56: Simulación atomística de la producción y evolución de ...

39

Figura 15. Imagen de microscopía electrónica de transmisión de una lámina de

Fe ultrapuro irradiado con iones de Ga+ de 30 keV mediante un sistema FIB.

La forma de estos defectos, con tres lóbulos oscuros alrededor de un punto

brillante, es semejante a la forma que observamos en las imágenes TEM

simuladas de nuestras cascadas, por lo que podrían ser loops de vacantes. Sin

embargo, estos resultados son preliminares y deberán ser estudiados en mayor

profundidad en el futuro.

20 nm

Page 57: Simulación atomística de la producción y evolución de ...

40

Page 58: Simulación atomística de la producción y evolución de ...

4. TRABAJOS PUBLICADOS O ACEPTADOS

Page 59: Simulación atomística de la producción y evolución de ...
Page 60: Simulación atomística de la producción y evolución de ...

I

Page 61: Simulación atomística de la producción y evolución de ...
Page 62: Simulación atomística de la producción y evolución de ...

Journal of Nuclear Materials 452 (2014) 453–456

Contents lists available at ScienceDirect

Journal of Nuclear Materials

journal homepage: www.elsevier .com/ locate / jnucmat

Molecular dynamics simulations of irradiation of a-Fe thin films withenergetic Fe ions under channeling conditions

http://dx.doi.org/10.1016/j.jnucmat.2014.05.0770022-3115/ 2014 Published by Elsevier B.V.

⇑ Corresponding author. Tel.: +34 965903400x2056; fax: +34 965909726.E-mail address: [email protected] (M.J. Caturla).

M.J. Aliaga a, A. Prokhodtseva b, R. Schaeublin b, M.J. Caturla a,⇑a Departamento de Física Aplicada, Facultad de Ciencias, Fase II, Universidad de Alicante, Alicante E-03690, Spainb Ecole Polytechnique Féderale de Lausanne (EPFL), Centre de Recherches en Physique des Plasmas, Association Euratom-Confédération Suisse, 5232 Villigen PSI, Switzerland

a r t i c l e i n f o

Article history:Received 18 November 2013Accepted 31 May 2014Available online 10 June 2014

a b s t r a c t

Using molecular dynamics simulations with recent interatomic potentials developed for Fe, we havestudied the defects in thin films of pure bcc Fe induced by the displacement cascade produced by Featoms of 50, 100, and 150 keV impinging under a channeling incident angle of 6 to a [001] direction.

The thin films have a thickness between 40 and 100 nm, to reproduce the thickness of the samples usedin transmission electron microscope in situ measurements during irradiation. In the simulations we focusmostly on the effect of channeling and free surfaces on damage production. The results are compared tobulk cascades. The comparison shows that the primary damage in thin films of pure Fe is quite differentfrom that originated in the volume of the material. The presence of near surfaces can lead to a variety ofevents that do not occur in bulk collisional cascades, such as the production of craters and the glide ofself-interstitial defects to the surface. Additionally, in the range of energies and the incident angle used,channeling is a predominant effect that significantly reduces damage compared to bulk cascades.

2014 Published by Elsevier B.V.

1. Introduction

One of the main challenges facing the use of nuclear fusion as afuture energy source is related to the reactor materials. In thefuture fusion reactors, a large amount of 14 MeV neutrons will beproduced, which will deteriorate the reactor vessel as theycontinuously collide with the first wall or plasma facing compo-nents. In order to have long-lived nuclear reactors that permitfusion to be an economically competitive energy source, it isnecessary to understand radiation damage in the harshest condi-tions expected for these materials. However, experiments withhigh energy neutrons are very expensive and there are only afew facilities around the world where they can be performed. Inaddition, neutrons produce radioactive isotopes in the irradiatedmaterial, which complicates its analysis. For these reasons, ionirradiation appears to be an excellent tool for understanding defectproduction in materials. However, the correlation to neutron irra-diation induced damage is not trivial. Therefore, substantial effortis devoted to modeling these effects, mainly using moleculardynamics, and to model-oriented experiments using ion implanta-tion. This would increase our knowledge of the correlation toneutron irradiation and, eventually, it will expand our capacity topredict these effects.

Experiments are currently being performed in JANNuS, France,where while a target is being irradiated with Fe ions it is observedin situ in a transmission electron microscope (TEM). For theseexperiments to be performed the samples can only be a few tensof nanometers thick to be transparent to electrons. There are manystudies of ion irradiation in a-Fe using molecular dynamics simu-lations (see Ref. [1] for a review), but the majority of these calcula-tions are performed in bulk specimens in view of reproducingneutron damage in reactor materials. The conditions in an ion irra-diation experiment are significantly different. On the one hand, forlow ion irradiation energies, the damage will occur near thesurface. On the other hand, in TEM in situ irradiation, the effectof surfaces cannot be neglected because the specimen thicknessis between 40 and 200 nm. The influence of surfaces on theprimary damage was studied in detail during the 1990s usingmolecular dynamics simulations for low energies in f.c.c. materials[2–4]. However, the only studies in Fe including surfaces are thosefrom Refs. [5,6]. Those calculations were performed for very lowion energies (10–20 keV) and using interatomic potentials thatare now known to provide the wrong self-interstitial configuration.Nonetheless, the influence of the surface in terms of damageproduction and defect distribution was already shown.

Our objective is to further study how damage is produced inthese thin films and how it relates to bulk specimens. In this articlewe study, using molecular dynamics simulations with recent inter-atomic potentials, the very first stages of the damage produced by

Page 63: Simulación atomística de la producción y evolución de ...

Fig. 1. Defect distribution for a series of 20 simulations of cascades in thin filmsproduced by irradiation with a 100 keV Fe atom.

454 M.J. Aliaga et al. / Journal of Nuclear Materials 452 (2014) 453–456

Fe ions with energies between 50 and 150 keV in thin films of purea-Fe under channeling conditions.

2. Method

Calculations were performed using the molecular dynamicscode MDCASK, developed at Lawrence Livermore National Labora-tory, with the interatomic potential of Dudarev and Derlet [7] fora-Fe. This interatomic potential was developed to include themagnetic character of Fe by adding a new term in the embeddingfunction of the potential [7]. For short range interactions thepotential is connected to the Universal potential as described inRef. [8]. Displacement cascades were simulated injecting an exter-nal Fe atom, or primary knock-on atom (pka), in the top freesurface of a thin film with a [001] normal. The energy of theincident Fe atom is 50, 100, and 150 keV. The Fe atom incident

Fig. 2. Three snapshots of the time evolution of the damage produced by a 150 keV Finterstitials, (a) 0.06 ps after the initiation of the recoil, (b) 0.5 ps and (c) 10 ps. (For interweb version of this article.)

angle was tilted 6 from normal incidence, lower than the criticalangle for channeling which, according to the Lindhard expression[9] is 21, 15, and 12 for the 50, 100, and 150 keV respectively.Cell dimensions are 40 nm 40 nm 40 nm (140 140 140lattice parameters) with a total of 5,154,801 atoms. We used theLindhard model to include a friction force proportional to thevelocity for all atoms with a kinetic energy greater than 5 eV tomimic the inelastic energy losses produced by collisions with theelectrons. Periodic boundary conditions were applied to directionsperpendicular to that of the incident ion to emulate an infinite sys-tem, whereas in the direction of the incident ion free surfaces areconsidered. The conditions used are intended to reproduce theexperiments already published by Yao et al. [10].

The simulations were run for the duration needed for the eventto be completed and the number of defects stabilized, which occursin less than 15 ps in most cases. The temperature was kept near 0 Kto minimize thermal atomic vibrations and thus facilitate the iden-tification of defects. Before the recoil atom was started, the cell wasequilibrated for 1 ps to the desired temperature of the simulation.A thermal bath was imposed to dissipate the energy deposited bythe ion in the solid by rescaling the velocity of two atom layerssituated at the end of the [100] and [010] directions.

In order to identify the defects, vacancies and interstitials, weused Wigner–Seitz cells centered in each (perfect) lattice positionso that an empty cell corresponds to a vacant and a doubleoccupied cell corresponds to an interstitial defect. Variability wasintroduced to obtain statistical results by changing the value forthe impact angle orientation from 0 to 180 for up to a total of20 cases for the 50 and 100 keV pka, and 30 cases for the150 keV pka.

3. Results and discussion

Results are presented here and compared to the results byBjörkas for bulk cascades with energies of 50 keV and 100 keV [11].

Fig. 1 represents the final number of defects (vacancies andinterstitials) that results for the series of 20 simulations ofirradiation of thin films with a pka of 100 keV. It appears that the

e ion in a 40 nm thick Fe sample. Light dots are vacancies and dark dots are self-pretation of the references to color in this figure legend, the reader is referred to the

Page 64: Simulación atomística de la producción y evolución de ...

Fig. 3. Average number of defects produced in thin films (vacancies and intersti-tials) compared to bulk cascades (Frenkel pairs) [11]. (For interpretation of thereferences to color in this figure legend, the reader is referred to the web version ofthis article.)

M.J. Aliaga et al. / Journal of Nuclear Materials 452 (2014) 453–456 455

dispersion in the number of defects is much wider in cascadesaffected by free surfaces than in cascades of the same energy pro-duced in the bulk of the material [11]. In the latter case, the meanvalue of Frenkel pairs produced is 159 with a standard deviation of7 [11]. In our case, we observe, on the one hand, that the averagenumber of vacancies does not match that of interstitials, with 52vacancies and 34 self-interstitials. On the other hand, the standarddeviation for vacancies is 17 and for interstitials is 11. This is inagreement with the results of Stoller et al. [5,6] that show both ahigher production of vacancies than self-interstitials as well as lar-ger standard deviations in their results for surface damage. It isinteresting to note that recent simulations of cascades close tograin boundaries by Bai et al. also show a larger production ofvacancies than interstitials close to the grain boundary [12]. In thatcase, with the use of temperature accelerated dynamics (TAD), theauthors show how at later times self-interstitials trapped at grainboundaries can annihilate those remaining vacancies in the bulk.

The higher number of vacancies is mainly the result of the migra-tion of interstitials to the surface. Differences in the energy spatialdeposition account for the dispersion in the results, with some casesin which the energy is deposited very close to the surface, creatingsurface damage [2], and other cases in which damage is deeper,being more similar to the results obtained in bulk cascades.

Fig. 2 shows three snapshots of the damage produced by a150 keV Fe ion in a 40 nm thick Fe sample. Dark dots show the loca-tions of self-interstitials ions while light dots are the locations ofvacancies. Fig. 2(a) shows the distribution of defects at 0.06 ps,before the ion reached the back surface. Fig. 2(b) shows thedamage at 0.5 ps which corresponds to the time of maximum num-ber of displacements in the lattice, that is the peak of the collisioncascade. Much of the damage is recovered during the next stage,reaching an almost constant number of defects. The final distribu-tion at 10 ps is shown in Fig. 2(c). Note that the damage producedin the top and bottom surfaces is mostly vacancies since the intersti-tials are ejected to the surface and stay as adatoms (not shown in thefigure), again because of the creation of surface damage. It should benoted that in this case the pka traveled through the sample andescaped through the bottom surface, producing little damage, whichshows that the impact angle used (6) is within channeling condi-tions. Therefore the damage produced consists of Frenkel pairs andsmall vacancy and self-interstitial clusters. In all cases studied inthese conditions there was transmission of the pka. We have ana-lyzed the damage produced as a function of the distance to the frontand back surfaces. For the case of Fig. 2(c), the total number of vacan-cies produced in this cascade is 137 while the total number ofself-interstitials produced is 95. The total number of vacancies at adistance of 3 nm from either the back or the front surface is 56, whileonly 6 interstitials are found at this distance. Below 3 nm we thenfind 89 self-interstitials and 81 vacancies, that is, similar values forthe two types of defects, getting closer to bulk calculations.

Comparing the different energies studied, we observe sometrends. The maximum production of defects occurs when eventsvery near the surface take place, leading to the formation of surfacedamage either at the top or the bottom surface. On the other hand,in the cases with the minimum number of defects, events very nearthe surface are not important and the damage created is moresimilar to that produced in bulk material.

Table 1Average number of stable defects and their standard deviation produced in thin films and

Energy (keV) Thin films vacancies Standard deviation Thin films inter

50 47 8 32100 52 17 34150 73 28 46

a Ref. [11].

The results for our simulations in thin films and in bulk cas-cades [11] are summarized in Table 1. Fig. 3 graphically representsthese values. Unlike the results of Stoller et al. [5,6], the averagenumber of defects is significantly lower in our simulations thanin the case of bulk cascades. This is due to the channeling condi-tions used in the calculations, which allow the ion to travelthrough the whole sample thickness without depositing all itsenergy in the target. The total energy deposited by the ion can becalculated from the difference between the initial energy and theenergy of the ion after it crosses the film. On average, the percent-age of the energy deposited for the 50 keV, 100 keV, and 150 keVcascades is 44%, 32%, and 29% respectively. However, even underthese conditions, where little damage should be produced, thereare cases where a high number of vacancies is created, as can beseen in Fig. 1. These are related to cases where surface damage isformed in the front or back surfaces, or in both.

4. Conclusions

In this paper we have studied damage produced by Fe ions withenergies between 50 and 150 keV in thin films of pure bcc Fe usingmolecular dynamics simulations under channeling conditions. Theresults are also compared with those obtained in bulk cascades.Conclusions can be summarized as follows:

1. Unlike cascades simulated in the bulk that produce the samenumber of vacancies and interstitials, cascades in thin filmssimulated by external ion irradiation produce more vacanciesthan interstitials, even under channeling conditions such asthose studied here.

2. Dispersion in the number of defects in ion irradiated thin filmsis greater than in simulations in bulk materials due to the vari-ety of events that can occur because of the surface influence,and this dispersion increases with the pka energy.

bulk cascades.

stitials Standard deviation Bulka Frenkel pairs Standard deviation

7 91 511 159 714 – –

Page 65: Simulación atomística de la producción y evolución de ...

456 M.J. Aliaga et al. / Journal of Nuclear Materials 452 (2014) 453–456

These conclusions show that to be able to reproduce andunderstand ion irradiation experiments in thin films analyzed byTEM, a detailed description of the primary damage is needed. Thisimplies the inclusion of surfaces in the simulations because thedamage produced is completely different from that originated inbulk materials. The time scale in MD simulations is too short tobe able to make a direct comparison with TEM experimentalmeasurements. Kinetic Monte Carlo (kMC) simulations using theinformation obtained from these MD calculations could be usedto make such comparison. If the initial damage distribution hasan impact on microstructure evolution under irradiation, asshown in previous works, the concentration of defects with doseas well as the cluster size distribution obtained by kMC shouldbe different when using bulk cascades or surface cascades.These results could then be contrasted with the experimentallymeasured values.

Acknowledgements

We thank Cristian Denton from the UA and Carolina Björkasfrom the University of Helsinki for helpful discussions. This workwas supported by the FPVII projects FEMaS, GETMAT and PERFECT

and by the MAT-IREMEV program of EFDA. We acknowledge thesupport of the European Commission, the European Atomic EnergyCommunity (Euratom), the European Fusion Development Agree-ment (EFDA) and the Forschungszentrum Jülich GmbH, jointlyfunding the Project HPC for Fusion (HPC-FF), Contract NumberFU07-CT-2007-00055. The views and opinions expressed hereindo not necessarily reflect those of the European Commission.

References

[1] L. Malerba, J. Nucl. Mater. 351 (2006) 28–38.[2] M. Ghaly, R.S. Averback, Phys. Rev. Lett. 72 (1994) 364.[3] K. Nordlund, J. Keinonen, M. Ghaly, R.S. Averback, Nature 398 (1999) 48.[4] K. Nordlund et al., Nucl. Instrum. Methods Phys. Res. B 148 (1999) 74–82.[5] R.E. Stoller, S.G. Guiriec, J. Nucl. Mater. 329 (2004) 1238.[6] R.E. Stoller, J. Nucl. Mater. 307–311 (2002) 935.[7] S. Dudarev, P. Derlet, J. Phys. Condens. Matter. 17 (2005) 1–22.[8] C. Björkas, K. Nordlund, Nucl. Instrum. Methods Phys. Res. B 259 (2007) 853–

860.[9] L.-P. Zheng et al., Nucl. Instrum. Methods Phys. Res. B 268 (2010) 120–122.

[10] Z. Yao, M. Hernández Mayoral, M.L. Jenkins, M.A. Kirk, Philos. Mag. 88 (2008)2851.

[11] C. Björkas, (Private communication).[12] X.-M. Bai, A.F. Voter, R.G. Hoagland, M. Nastasi, B.P. Uberuaga, Science 327

(2010) 1631–1634.

Page 66: Simulación atomística de la producción y evolución de ...

II

Page 67: Simulación atomística de la producción y evolución de ...
Page 68: Simulación atomística de la producción y evolución de ...

Nuclear Instruments and Methods in Physics Research B 352 (2015) 217–220

Contents lists available at ScienceDirect

Nuclear Instruments and Methods in Physics Research B

journal homepage: www.elsevier .com/locate /n imb

Surface damage in TEM thick a-Fe samples by implantationwith 150 keV Fe ions

http://dx.doi.org/10.1016/j.nimb.2014.11.1110168-583X/ 2014 Elsevier B.V. All rights reserved.

⇑ Corresponding author.

M.J. Aliaga a,⇑, M.J. Caturla a, R. Schäublin b

a Dept. Física Aplicada, Facultad de Ciencias, Fase II, Universidad de Alicante, Alicante E-03690, Spainb Metal Physics and Technology, Department of Materials, ETH Zürich, HCI G515, Vladimir-Prelog-Weg 5, 8093 Zürich, Switzerland

a r t i c l e i n f o

Article history:Received 11 July 2014Received in revised form 21 October 2014Accepted 29 November 2014Available online 17 December 2014

Keywords:Molecular dynamicsDefectsIon irradiationSurface damageTransmission electron microscopy

a b s t r a c t

We have performed molecular dynamics simulations of implantation of 150 keV Fe ions in pure bcc Fe.The thickness of the simulation box is of the same order of those used in in situ TEM analysis of irradiatedmaterials. We assess the effect of the implantation angle and the presence of front and back surfaces. Thenumber and type of defects, ion range, cluster distribution and primary damage morphology are studied.Results indicate that, for the very thin samples used in in situ TEM irradiation experiments the presenceof surfaces affect dramatically the damage produced. At this particular energy, the ion has sufficientenergy to damage both the top and the back surfaces and still leave the sample through the bottom. Thisprovides new insights on the study of radiation damage using TEM in situ.

2014 Elsevier B.V. All rights reserved.

1. Introduction

Ion irradiation experiments are being used extensively tounderstand the fundamental aspects of the damage produced inmetals and alloys by irradiation [1]. In the current nuclear powerplants and experimental fusion reactors the damage is producedby neutrons. However, the study of neutron irradiation is difficultsince conditions cannot be easily controlled, samples are activatedand experiments are very costly. For this reason, ion irradiation isnowadays being used to gain basic understanding of the effects ofradiation in the structural materials of the reactors.

Iron is the main element of the reactor vessel, and although ithas been studied for many years, there are still many issues underdebate considering its radiation damage. Both neutrons and Fe ionsproduce damage in cascades, but their damage profile is very dif-ferent. Neutrons have a long range of penetration and producedamage quite homogeneously, whereas ion damage is more super-ficial. However, 100 keV Fe ions have been used in the last fewyears to approach indirectly the study of neutron irradiation inFe, since the first Fe atom that is hit by a neutron in a lattice (pri-mary knocked-on atom o PKA) has around 100 keV energy [2]. Theirradiated samples can then be examined by in situ TEM [3] infacilities such as Jannus at CEA in France [4] or the IVEM-TandemFacility at Argonne National Laboratory [5]. Using this character-ization technique the sample can be observed while it is being

irradiated. The requirement is that it has to be between 40 and100 nm thick to be transparent to electrons.

As it has been demonstrated in previous works [6–13] the pres-ence of surfaces in metals affect the damage produced in the mate-rial, being quite different from the damage in bulk. Earlier, it hasbeen observed that irradiation at room temperature of pure Fe inthe form of a transmission electron microscopy (TEM) thin filmleads to a0 h100i dislocation loops while in the bulk form irradi-ated Fe exhibits mainly ½a0 h111i [14]. This has been attributedto so-called elastic ‘image forces’ due to the free surfaces but neverquantified. This effect is confirmed in recent works [15–18]. In thispaper we continue our work from an atomistic point of view withthe study of the primary damage produced by Fe ions of 150 keV inthin films of pure Fe using molecular dynamics.

2. Methodology

Calculations were performed using the molecular dynamicscode MDCASK with the interatomic potential of Dudarev and Der-let [19]. This potential was modified for short range interactionsfollowing the procedure described in [20]. This potential wasselected since it reproduces fairly well the energetics of pointdefects compared to density functional theory calculations [20].It is also the only interatomic potential for Fe that includes anexplicit term for the magnetic contribution to the interatomicpotential energy. Displacement cascades were simulated sendingan Fe atom with an energy of 150 keV towards the top free surface

Page 69: Simulación atomística de la producción y evolución de ...

218 M.J. Aliaga et al. / Nuclear Instruments and Methods in Physics Research B 352 (2015) 217–220

of a [001] thin film of a-Fe. Two impinging angles have been used,10 and 22, being this second angle the one used in the TEM in situanalysis of ion irradiation experiments at Orsay JANNUS facility [4].20 cases were run for each angle. Variability was introducedchanging the azimuthal angle from 0 to 200. Simulation cellscontained 10,076,401 atoms with a size of 180a0 180a0 180a0,where a0 is the lattice parameter for Fe (a0 = 2.8665 Å). This sizecorresponds to thin films of about 50 nm. The setup described cor-responds to the energies used in the experiments for low doses ofYao et al. [15]. Temperature was kept close to 0 K in order to avoidthermal vibrations and thus facilitate the identification of defects.The excess energy deposited by the injected atom is dissipated byadding a thermal bath that scales the velocity of two atom layers atthe border of the simulation cell. Inelastic energy losses wereincluded by the Lindhard model [21], which introduces a frictionforce proportional to the velocity. This force was introduced onlyfor those atoms with a kinetic energy larger than 5 eV. Periodicboundary conditions are imposed in two axes, while free surfacesare considered in the third axis. Simulations were run until thenumber of defects reached a stable population (25 ps in mostcases). Wigner–Seitz cells were used to identify the defects. Thenthe defects are grouped in clusters considering that two defectsbelong to the same cluster when the distance between them isbetween first and second nearest neighbours.

3. Results

Fig. 1 shows the primary damage of three representative casesof the events found in the simulations after the irradiation with a150 keV Fe atom. The location of vacancies (light circles) andself-interstitials (dark circles) are shown for the three cases afterthe simulation had run for 25 ps. The arrow indicates approxi-mately the initial position of the energetic atom. Both surfacesand adatoms are also represented in the figure. In Fig. 1(a) the Fe

Fig. 1. (a–c) show three representative 150 keV cascades of Fe implantation in Fe after 2also represented. The arrow indicates the implanted Fe atom. In (a) and (c) the Fe atom imbottom surface damage produced in (c) is shown.

atom is launched with an angle of 10 impacting heavily anddepositing most of its energy near the top surface. The damage isdivided into 3 displacement subcascades. The first and most ener-getic one occurs at the surface, creating a large vacancy cluster of1070 vacancies, another vacancy cluster of 115 vacancies and493 adatoms and 764 sputtered atoms (not shown in the figure)above the surface. The other two subscascades are around the cen-tre of the simulation cell where the ion stops. The total number ofvacancies is 1557 and the total number of interstitials is 155. InFig. 1(b) the Fe atom is sent with an angle of 22 and travelsthrough the whole sample leaving the film at the bottom surfacewith 20% of its initial energy, so depositing 80% of the total150 keV energy. In this case both top and bottom surfaces are dam-aged, but the back surface more strongly, with the creation of a 196vacancy cluster. Finally Fig. 1(c) shows an event in which the atomimpinges with a 10 angle and again goes through the entire sam-ple, but in this case it stops just before it escapes the film due to astrong collision with the back surface. In this case the top surfacebarely suffers any damage, but in the bottom surface a huge craterof 3441 vacancies is created, as well as large islands of adatoms.Fig. 1(d) shows a close-up of the crater created at the bottom sur-face of the case represented in Fig. 1(c).

Fig. 2 represents the histograms of the total number of vacan-cies and interstitials that results from the 20 cases simulated foran impact angle of 10 (two of these cases shown in Fig. 2(a) and(c)) and the other 20 cases for the impact angle of 22 (one exam-ple in Fig. 2(b)). As already shown in previous works [6,12] thenumber of vacancies for cascades with near surfaces is greater thanthe number of self-interstitials. This is due to the attraction theself-interstitials suffer by the surface. Also, the dispersion of resultsis larger than in bulk cascades. The main difference with [6] for thesame energy is that the increase in the angle results in some events(3 for 10 and 2 for 22) with a huge number of vacancies. Two ofthese events for the impact angle of 10 are the ones represented inFig. 1(a) and (c). These type of events correspond to cases where

5 ps. Light circles are vacancies and darker circles are interstitials. Both surfaces arepact angle is 10 and in (b) the impact angle is 22 off normal. In (d) a close-up of the

Page 70: Simulación atomística de la producción y evolución de ...

Fig. 2. Defect distribution for a series of 20 simulations of cascades in thin films produced by irradiation with a 150 keV Fe atom impacting with an angle of (a) 10 and (b) 22of the normal.

Fig. 3. Statistical analysis of ion ranges for the 150 keV Fe atom impacting with an angle of (a) 10 and (b) 22 of the normal.

Table 1Average number of vacancies and interstitials and their cluster fractions after the 150 keV cascades. Large clusters contain more than 55 defects. The ion range is also shown.

Angle () Number ofvacancies

Number ofinterstitials

% V in clusters % V in largeclusters

% I in clusters % I in largeclusters

Ion range (nm)

10 132 73 43 12 24 0 4222 225 166 47 16 27 0 30

M.J. Aliaga et al. / Nuclear Instruments and Methods in Physics Research B 352 (2015) 217–220 219

the energetic Fe atom injected has a strong collision near the top orthe back surface.

The histograms of the ion ranges for both angles are presentedin Fig. 3. Comparing both angles it is clear, on one hand, that in themajority of cases the PKA has enough energy to escape the samplethrough the bottom surface, and, on the other hand, that anincrease in the impact angle from 10 to 22 leads to a reductionin the ion range, as expected. As with the number of defects, thereis a wide spread in the results for the different cases. The caseswhere the ion range is very short coincide usually with the caseswhere the Fe atom has collided very strongly with or near thetop surface. Indeed, in one of the two events for 22 with manyvacancies, 1441 in this case, the PKA is backscattered after collidingwith an atom near the top surface, creating a large h100i loop with

521 vacancies. After this, the bottom surface is also badly damagedby secondary subcascades. On average, the energy deposited for10 impacts is 67%, and 86% for 22.

Table 1 summarizes the mean values obtained from fitting theabove histograms for the number of vacancies, self-interstitialsand the ion range to either lognormal or Gaussian distributions.The percentage of vacancies and self-interstitials in clusters andin large clusters (more than 55 defects) are also shown. It can beseen on the table, as mentioned above, that the ion range for 22with a value of 30 nm is shorter than the value for 10 which is42 nm. The SRIM values [22] are 50 nm for the ion range at animpact angle of 10 and 47 nm for the ion range at an impact angleof 22. This is reasonable because SRIM assumes on one hand a ran-dom target, and on the other hand these events where the ion

Page 71: Simulación atomística de la producción y evolución de ...

Fig. 4. Histograms of number of vacancy (a) and interstitial (b) clusters of different sizes normalized to the number of cascades.

220 M.J. Aliaga et al. / Nuclear Instruments and Methods in Physics Research B 352 (2015) 217–220

interacts very energetically with the surfaces shortening its rangedo not happen.

The increase in the angle also has an effect, as expected, on themean value of vacancies and interstitials. The angle of 10 is stillslightly below the Lindhard critical angle for channeling [23] which,for 150 keV is 12 and this results in a lower number of defects as amean value. Fig. 4 represents the distribution of clusters normal-ized by the number of cascades for both angles. It can be seen a ten-dency to larger clusters of vacancies and interstitials from 10 to22, but this increase is not remarkable because at 10 there areevents where the back surface is profoundly damaged.

4. Conclusions

The primary damage produced in thin films of pure bcc Fe byirradiation with150 keV Fe ions using two different implantationangles has been studied. Results show that, differently from lowerenergies where only the top surface is quite affected, at this energyregime front and back surfaces can be damaged. When the ionimpacts strongly with or near one of the surfaces the creation oflarge vacancy clusters is observed. This is produced by the attrac-tion the self-interstitials suffers when they are close enough tothe surface. These results have important implications for higherenergies, because they indicate that at high irradiation energies,like the range of energies used mainly in the in situ TEM irradiationexperiments, the damage will be produced mostly at the back sur-face of the film. Moreover, these effects should be taken intoaccount in models that predict the latter evolution of damageand damage accumulation, such as kinetic Monte Carlo or rate the-ory calculations.

These calculations provide new insights on the study of radia-tion damage using TEM in situ irradiation experiments, providingthe fundamental background needed to use the data from TEM insitu experiments to understand damage in bulk specimens.

Acknowledgements

We would like to thank Drs. A. Prokodtseva, M. Hernández-Mayoral, Z. Yao and S. Dudarev for fruitful discussions. Simulations

were carried out in the computer cluster of the Dept. of AppliedPhysics at the UA, the HPC-FF supercomputer of the Jülich Super-computer Center, Germany, and the HELIOS supercomputer inJapan. MJA thanks the UA for support through an institutionalfellowship. This work was supported by the European FusionDevelopment Agreement (EFDA), the VII EC framework throughthe GETMAT and MATISSE projects, and the Generalitat ValencianaPROMETEO2012/011.

References

[1] J.L. Boutard, A. Alamo, R. Lindau, M. Rieth, C.R. Phys. 9 (2008) 287.[2] Hernández-Mayoral Mercedes, Estudio por Microscopía Electrónica de

Transmisión del efecto de la irradiación iónica y neutrónica en hierro puro yaleaciones modelo de los aceros de vasija de los reactors nucleares. Madrid,Centro de Investigaciones energéticas, Medioambientales y Tecnologicas(CIEMAT). Departamento de Tecnología, División de Materiales, 2007.

[3] Y. Matsukawa, S.J. Zinkle, Science 318 (2007) 959.[4] <http://jannus.in2p3.fr/spip.php?rubrique15>.[5] C.W. Allen, L.L. Funk, E.A. Ryan, Mater. Res. Soc. Symp. Proc. 396 (1996) 641.[6] M.J. Aliaga, A. Prokhodtseva, R. Schaeublin, M.J. Caturla, J. Nucl. Mater. 452

(2014) 453–456.[7] M. Ghaly, R.S. Averback, Phys. Rev. Lett. 72 (1994) 364.[8] K. Nordlund, J. Keinomen, M. Ghaly, R.S. Averback, Nature 398 (1999) 48.[9] K. Nordlund, J. Keinonen, M. Ghaly, R.S. Averback, Nucl. Instr. Meth. Phys. Res.

B 148 (1999) 74–82.[10] S.V. Starikov, Z. Insepov, J. Rest, Phys. Rev. B 84 (2011) 104109.[11] P.D. Lane, G.J. Galloway, R.J. Cole, M. Caffio, R. Schaub, G, J. Ackland Phys. Rev. B

85 (2012) 094111.[12] R.E. Stoller, J. Nucl. Mater. 307–311 (2002) 935.[13] R.E. Stoller, S.G. Guiriec, J. Nucl. Mater. 329 (2004) 1238.[14] B. Masters, Nature 200 (1963) 254.[15] Z. Yao, M. Hernández Mayoral, M.L. Jenkins, M.A. Kirk, Phil. Mag. 88 (2008)

2851.[16] M.H. Mayoral, Z. Yao, M.L. Jenkins, M.A. Kirk, Phil. Mag. 88 (2008) 2881.[17] A. Prokhodtseva, B. Décamps, R. Schäublin, J. Nucl. Methods 442 (2013) S786–

S789.[18] A. Prokhodtseva, B. Décamps, A. Ramar, R. Schäublin, Acta Mater. 61 (2013)

(2013) 6958–6971.[19] S. Dudarev, P. Derlet, J. Phys.: Condens. Matter 17 (2005) 1–22.[20] C. Björkas, K. Nordlund, Nucl. Instr. Meth. Phys. Res. B 259 (2007) 853–860.[21] J. Lindhard, M. Sharff, Phys. Rev. 124 (1961) 128.[22] J.F. Ziegler, J.P. Biersack, The Stopping and Range of Ions in Matter, SRIM-2003,

Ó1998, 1999 by IB77M co.[23] L.P. Zheng, Z.-Y. Zhu, Y. Li, F.O. Goodman, Nucl. Instr. Meth. Phys. Res. B 268

(2010) 120–122.

Page 72: Simulación atomística de la producción y evolución de ...

III

Page 73: Simulación atomística de la producción y evolución de ...
Page 74: Simulación atomística de la producción y evolución de ...

Acta Materialia 101 (2015) 22–30

Contents lists available at ScienceDirect

Acta Materialia

journal homepage: www.elsevier .com/locate /actamat

Surface-induced vacancy loops and damage dispersionin irradiated Fe thin films

http://dx.doi.org/10.1016/j.actamat.2015.08.0631359-6454/ 2015 Published by Elsevier Ltd. on behalf of Acta Materialia Inc.

⇑ Corresponding author.E-mail address: [email protected] (M.J. Caturla).

M.J. Aliaga a, R. Schäublin b, J.F. Löffler b, M.J. Caturla a,⇑a Dept. Física Aplicada, Facultad de Ciencias, Fase II, Universidad de Alicante, Alicante E-03690, Spainb Laboratory of Metal Physics and Technology, Department of Materials, ETH Zurich, 8093 Zurich, Switzerland

a r t i c l e i n f o a b s t r a c t

Article history:Received 20 May 2015Revised 21 July 2015Accepted 28 August 2015

Keywords:Molecular dynamics simulationsIon irradiationIn situ transmission electron microscopyDefectsMicrostructure

Transmission electron microscopy (TEM) in situ ion implantation is a convenient way to study radiationdamage, but it is biased by the proximity of the free surfaces of the electron transparent thin sample. Inthis work this bias was investigated by performing irradiation of Fe in thin foil and bulk form with ions ofenergies between 50 keV and 100 keV using molecular dynamics simulations. The damage resulting fromthe subsequent displacement cascades differs significantly between the two sample geometries. Themost remarkable difference is in the resulting h100i vacancy loops. Both their size and frequency aremuch greater in thin films, with loops reaching 4 nm in size. This is due to an imbalance between thenumber of vacancies and self-interstitials produced, since the faster self-interstitials can escape to thesurfaces and remain there as ad-atoms. In addition, the self-interstitial clusters are smaller for thin foilsand there is a larger dispersion of the induced damage in terms of defect number, defect clustering anddefect morphology. The study discusses the impact of these results on the study of radiation effectsduring in situ experiments.

2015 Published by Elsevier Ltd. on behalf of Acta Materialia Inc.

1. Introduction

Ion implantation is a common way to produce defects in mate-rials in a controlled manner, in terms of particle type and energy,irradiation dose, dose rate, and temperature. It is therefore a veryvaluable tool for investigating radiation effects in materials witha high degree of control of all involved variables [1]. The damageproduced by implantation can be characterized using transmissionelectron microscopy (TEM), which provides information about theirradiation-induced defects, their number density, size, and possi-bly their type for those defects that are larger than about 1 nm[2]. In some facilities, such as JANNuS at CEA in France [3] or theIVEM-Tandem Facility at Argonne National Laboratory [4], it is pos-sible to perform ion irradiation experiments in situ in a TEM, thusallowing observation while irradiating [5]. This technique is beingused to validate simulation models of radiation effects to predictthe damage produced by the 14 MeV neutrons expected in futurefusion reactors [1], with a particular focus on ferritic materialsbecause of their good resistance to irradiation relative, for instance,to austenitic stainless steels.

The difficulty of extrapolating the results of such in situ exper-iments to the effects of radiation in bulk materials was recognizedlong ago [6]. This difficulty arises because, to be able to use con-ventional TEM (CTEM) for the characterization of defects viadiffraction contrast, the sample must be thin enough to minimizeelectron absorption and inelastic scattering that blur the image,but thick enough to reduce the effect of the free surfaces, implyinga thickness of at least a few tens of nanometers but less than about80 nm [7]. Free surfaces have indeed an impact on the elastic fieldsof defects because of the so-called image forces, which alter theconfiguration of the defects, and their interaction and migration.Surfaces affect the irradiation-induced microstructure also simplybecause defects produced in their vicinity can annihilate there ifthe defects migrate to the free surfaces. It is known experimentallythat in the case of ferritic steels irradiation generates 1/2 h111i andh100i dislocation loops [8]. Because the activation energy formigration of 1/2 h111i loops along h111i directions is low, asshown both experimentally [9] and in computer simulations[10,11], they migrate rapidly and can thus easily meet a surfacewhere they will disappear. Note that if the foil normal is orientedclose to a h111i or h100i direction all 1/2 h111i loops can disap-pear at surfaces, while only half of them will escape if the orienta-tion of the foil is h110i [12]. The nature, vacancy or self-interstitial,of the loops observed after irradiation at low doses is controversialdue to the difficulty of analyzing them when they are small. For

Page 75: Simulación atomística de la producción y evolución de ...

M.J. Aliaga et al. / Acta Materialia 101 (2015) 22–30 23

larger sizes (beyond a few nanometers) obtained at higher doses,they have been identified as interstitial loops [13–16]. At low dosesbut with heavy ions of low energy, large h100i vacancy loopslocated close to the surface have also been identified [13,16].

The damage produced by energetic recoils in bulk bcc Fe hasbeen extensively studied by different research groups [17–20]using molecular dynamics (MD) simulations; for a review, seeRef. [21]. There is a good apparent understanding of the formationof interstitial clusters. For instance, recent simulations by Calderet al. [22] have shown how large interstitial clusters are producedin the early stages of the collision cascade in bulk Fe due to theinteraction between shock waves. Sand et al. [23] have also shownthat a scaling law can be obtained for the size of self-interstitialclusters as a function of recoil energy in bulk tungsten. However,much less is known regarding vacancy clustering and the effectof surfaces in bcc metals, which is relevant to thin foil irradiation.

The effect of free surfaces in fcc metals has been studied by sev-eral groups. The work performed in the 1990s by Ghaly and Aver-back [24–26] revealed how the presence of the surface can changethe morphology of the damage produced by a collision cascadewith respect to the bulk. Later, Nordlund et al. [25] showed thatin Cu and Ni new mechanisms of defect production occur whendamage is close to a surface. In bcc materials, recent MD simula-tions in bcc Mo revealed the formation of large vacancy loops justbelow the surface [27] and Osetsky et al. have described the evolu-tion of cascades close to the surface in bcc Fe [28].

In this work we focus on the differences between irradiation inFe thin film and bulk material which has not been studied in detailup to now. In particular we focus on the formation of vacancy clus-ters, and their orientation, size and frequency. For this comparisonwe study the early stage of the damage produced by recoils ofenergies between 50 keV and 100 keV in thin films and bulk sam-ples of bcc Fe at 0 K using MD simulations with 2 different empir-ical interatomic potentials. We assess and compare the number ofvacancies and interstitials, and the size and morphology of theirclusters immediately after the cool-down of the displacement cas-cade in the thin foil and bulk irradiation conditions, in order toidentify the impact of the free surfaces. We intend to understand,in particular, whether in Fe thin foils vacancy loops of sizes visiblein TEM could be formed directly in the cascade. We also discuss thevarious implications of extrapolating TEM in situ irradiation analy-sis to model radiation effects in bulk materials.

Table 1Irradiation conditions. The table shows the number of runs for each energy of the Feion in dependence of incidence angle, DD or AM interatomic potential, and thin filmor bulk sample geometry used. The sample thickness is 51.4 nm except for ⁄ and ⁄⁄,where it is 45.7 nm and 70 nm respectively.

Energy (keV) Incidence angle () Thin foilNumber ofcascadesperformed

BulkNumber ofcascadesperformed

DD AM DD AM

50 10 14⁄ 14 14 1422 17 30 14 14

100 10 20 20 – –22 20 30 – 14⁄⁄

2. Methodology

The parallel MD code MDCASK developed at LLNL [29] was usedfor the calculations. Two different types of interatomic potentialswere considered for comparison: one developed by Dudarev andDerlet [30] and one by Ackland et al. [31], cited respectively asDD and AM. Both potentials were modified for the high energy-short range interactions according to the procedure used in [32].These potentials were also used previously to study cascade dam-age in bulk samples [33]. They were selected because they repro-duce fairly well the stability of different point defects as obtainedby density functional theory (DFT) calculations. The cluster sizedistribution obtained with these potentials also seems to repro-duce experimental observations better than other potentials [33].

Simulations were performed in bcc Fe at constant volume. Twotypes of calculations were performed: one in a crystalline thin filmoriented along a h001i direction, and the other inside a crystal,quoted as ‘bulk’, with h001i axes. The crystal thin foil is con-structed with periodic boundary conditions along x and y axes,while both surfaces along the z direction were free. An energeticFe atom was launched from the outside of the crystal towardsthe surface with the selected implantation angle and energy. For

the bulk crystal periodic boundary conditions were applied in alldirections and one atom near the top of the simulation box wasselected as the high-energy atom, or primary knock-on atom(PKA), and was given the selected angle and energy.

In order to minimize defect migration and to focus on the effectof surfaces on defect production, the temperature in all simulationswas kept low, close to 0K. This was achieved by means of a thermalbath located at the border of the simulation box. The thermal bathconsisted of two atomic layers where the velocity of the atoms wasscaled every time step to correspond to the desired temperature of0 K. Calculations were followed for a period sufficient to reach aconstant value for the number of defects produced. In this waysimulated time amounts to about 20–40 ps depending on the case.Between 14 and 30 simulations were performed for each conditionstudied in order to obtain statistically significant results. The inci-dent angle, i.e., the angle of the incoming energetic ion withrespect to a line perpendicular to the surface (which would corre-spond to the polar angle in a spherical coordinate system) is keptconstant, while the azimuthal angle in that same coordinate sys-tem is varied via increasing it by 10 for each different case. Theinitial random distribution of velocities of the atoms in the simula-tion box is also different for each case. Two energies, 50 keV and100 keV, were considered, and two incidence angles, 10 and 22,for each energy. We should mention that the critical angle forchanneling according to the Lindhard expression [34] is 21 for50 keV and 15 for 100 keV ions. The simulation conditions aresummarized in Table 1. The simulation volume for most conditionswas 180 a0 180 a0 180 a0, where a0 = 2.8665 Å is the latticeparameter for the Fe potentials considered here, which corre-sponds to a cube of about 50 nm a side. This size is comparableto the thickness of the thin films that are used experimentally inTEM in situ irradiations [12,13].

The analysis of the resulting damage was first conducted usingthe Wigner–Seitz cell method, which gives the number of vacan-cies and self-interstitials in the crystal. Secondly, once pointdefects were identified, their clusters were established usinganother method, which considers that two defects belong to thesame cluster when the distance between them is between the firstand second nearest neighbor of the bcc Fe lattice. Finally, the radiusR of a loop with N defects was calculated according to the relation-

ship [35] R ¼ a0

ffiffiffiffiffiffiffiffiffiN

21=2p

q.

3. Cascade damage: defect type and size

We first present results for the thin foil geometry. Fig. 1 showsthe damage microstructure of the Fe thin film 25 ps after launchingthe energetic ion for two different runs (with different azimuthalangles) under the same condition, i.e. 0 K with a 50 keV Fe atomat an incidence angle of 10 and with the AM potential. Green

Page 76: Simulación atomística de la producción y evolución de ...

Fig. 1. Snapshot of the Fe thin film microstructure 25 ps after launching a 50 keV Fe ion at 10 incidence angle and at 0 K for two different azimuthal angles, 10 in (a) and 80in (b), and initial velocity distributions (a and b). Green/light spheres: vacancies, red/dark spheres: interstitials, yellow spheres: surface atoms. (For interpretation of thereferences to color in this figure legend, the reader is referred to the web version of this article.)

24 M.J. Aliaga et al. / Acta Materialia 101 (2015) 22–30

spheres mark the location of vacancies, while red spheres repre-sent self-interstitials. The arrow indicates the approximate initiallocation of the energetic atom. Surface atoms are also representedin the figure with yellow spheres. In Fig. 1(a) the damage consistsof isolated interstitials and vacancies, two neighboring relativelylarge clusters of 23 and 27 interstitials and two significantly largerclusters of 108 and 148 vacancies. The total number of vacancies inthis case is 448, while the total number of self-interstitials is 140.79% of the vacancies are in clusters and 57% of them are in clusterswith more than 55 defects, which corresponds to a loop of about1 nm in radius. 58% of the self-interstitials are in clusters but nocluster with more than 55 interstitials was found. As can beobserved in the figure, most of the damage is located within12 nm of the surface, while the displacement cascade reached amaximum depth of 20 nm. Note that there are 308 ad-atoms atthe free surface, as seen in Fig. 1(a). The number of missing self-interstitial atoms corresponds to these ad-atoms.

Fig. 1(b) exhibits a clear difference to Fig. 1(a) even though theinitial condition is the same, except for the azimuthal angle and theinitial random distribution of velocities of the atoms in the simula-tion box. The difference in Fig. 1(b) consists of the absence of largeinterstitial clusters and the larger size of the vacancy loops; there isindeed one loop with 317 vacancies, much larger than any of thosein Fig. 1(a). This is a trend we observed in the rest of the condi-tions: when a large vacancy loop is produced then only small inter-stitial clusters are obtained.

Fig. 2(a) shows the damage in an Fe thin film 20 ps after a dis-placement cascade induced by a 100 keV Fe ion with a 10 incidenceangle, with the AM potential. The total number of resulting vacan-cies is 662, while the total number of self-interstitials is 187. As inFig. 1, a significant number of atoms end up at the surface as ad-atoms (469), while 6 atoms are sputtered away. A large vacancy loopof 480 defects can be observed right below the surface. Fig. 2(b)shows a closer view of this large cluster. Its dimensions are approx-imately 9 a0 5 a0 14 a0, or 3 nm 1 nm 4 nm3. It takes theform of a rectangular parallelepiped with a (010) habit plane. Thisrectangular shape correlates well with the calculations by Gilbertet al. [36,37], showing that rectangular h100i vacancy loops areenergetically more stable than e.g. circular ones. Fig. 2(c) shows across section of the specimen through the vacancy cluster, confirm-ing the presence of a vacancy loop. The use of the RHFS rule gives a

Burgers vector b = a0[0–10] (Fig. 2(c)). Note that the presence ofh100i vacancy loops following displacement cascades in Fe wasalready shown by Soneda et al. [19] and Kapinos [38], but in bulkspecimens. They also indicated that their formation was rare, consti-tuting only 1% of all clusters.

We now present results for the bulk geometry. Fig. 3 shows thedamage in bulk Fe 30 ps after a 50 keV recoil with a 22 angle of inci-dence, with the AM potential. The differences in the damage distri-bution and configuration between the bulk and thin foil conditionsare clearly seen when compared with Fig. 1. Note that as expectedfor the bulk the number of vacancies is the same as the number ofself-interstitials, and equals 151. A cluster with 38 vacancies and arelatively large cluster above it with 37 interstitials are observed.

Fig. 4 shows the damage in bulk Fe 22 ps after a 100 keV recoilwith an angle of 22 and the AM potential for two different runs.There are differences: in the first case (Fig. 4(a)) small vacancyand self-interstitial clusters are observed; in the second case(Fig. 4(b)) slightly larger vacancy clusters are found, together witha few small self-interstitial clusters. We should note that, accord-ing to previous studies by Stoller and others [39], the breakdowninto sub-cascades for Fe occurs at around 20 keV. Both cases hereexhibit sub-cascade formation, visible with the different branchesin the cascade.

3.1. Statistical analysis

As mentioned above, several authors have already reportedsome large vacancy loops following displacement cascades in thebulk [19,33,38]. Our simulations indicate that the frequency of for-mation of these large vacancy loops is greater when damage is pro-duced in thin films, as seen in Figs. 1 and 2. In order to quantify thiseffect, we performed a statistical analysis of the data, focusing onthe impact of the two different interatomic potentials. The numberof vacancies and self-interstitials as well as their percentage inclusters were calculated for all conditions. Table 2 provides themean values of the number of point defects, Table 3 lists their clus-tering fraction, and Table 4 provides the ion range projected alongthe [001] direction (i.e. the one perpendicular to the surface) foreach condition. These values were obtained by fitting the differenthistograms to either a lognormal or a Gaussian distribution.

Page 77: Simulación atomística de la producción y evolución de ...

Fig. 2. (a) Snapshot of the Fe thin film microstructure 20 ps after a 100 keV cascade at 0 K and at 10 incidence angle after 20 ps. (b) Closeup of the h100i large vacancycluster. Green/light spheres: vacancies, red/dark spheres: interstitials, yellow spheres: surface atoms. (c) Cross section of the specimen through the h100i cluster showing allatoms with the RHFS circuit (in blue) used to identify its Burgers vector. (For interpretation of the references to color in this figure legend, the reader is referred to the webversion of this article.)

Fig. 3. Snapshot of the Fe bulk microstructure 30 ps after a 50 keV cascade at 0 Kand at 22 incidence angle. Green/light spheres: vacancies, red/dark spheres:interstitials. (For interpretation of the references to color in this figure legend, thereader is referred to the web version of this article.)

M.J. Aliaga et al. / Acta Materialia 101 (2015) 22–30 25

We first present the results for the thin film geometry. It can beseen in Table 2 that the number of self-interstitials is always lowerthan the number of vacancies. As already noted above, the missingself-interstitials correspond to atoms that have moved onto thesurface, remaining as ad-atoms, as seen in Figs. 1 and 2. The num-ber of atoms that are sputtered away is small. The maximumoccurs for the 50 keV Fe ion and 22 incidence with the AM poten-tial and constitutes 6% of the total number of interstitials. Table 2shows that when the energy increases from 50 keV to 100 keV thedamage exhibits an increase of between 50% and 200% for thenumber of vacancies and of between 10% and 110% for interstitials.

Fig. 5 presents the scatter in the data. It shows the histograms ofthe number of vacancies and self-interstitials obtained for the thin

foil. Results for the 50 keV Fe ion with an incidence angle of 10 aregiven in Fig. 5(a), and results for 100 keV and 22 incidence angle inFig. 5(c). These were obtained with the AM potential. Fig. 5(b) and (d) present the results for the DD potential for the 50 keVFe ion and 10 incidence and the 100 keV Fe ion and 22 incidence,respectively. Several common features emerge from all of thesecases.

It appears that the scatter in the number of defects from one runto another under the same irradiation condition is extensive. Someof the cases exhibit a significantly larger number of vacancies rel-ative to the average of 300–400 vacancies. For example, for thecondition of the 50 keV Fe ion, there is one case with close to3000 vacancies (Fig. 5(a)), while for 100 keV and the same poten-tial (Fig. 5(c)) one case has more than 1000 vacancies. Fig. 6 showsthe resulting damage configuration for one of these cases exhibit-ing a large number of vacancies: that with the AM potential, 50 keVcascade and 10 incidence. This damage is significantly differentfrom that shown in Figs. 1–3. Even though the damage appearsas a large vacancy cluster, it actually consists of a dislocated vol-ume of the crystal with a crystalline structure in the center, whichcorresponds to a short dislocation array similar to that obtained byGhaly and Averback [23]. We should point out that these cases arerare when using the AM potential but more frequent with the DDpotential. For 50 keV and 22 incidence, 60% of the vacancy clusterswith more than 20 vacancies correspond to surface damage in thecase of the DD potential, whereas this is only 7% for the AM poten-tial. As seen in Fig. 5(a)–(d), even if the extreme cases with a verylarge number of vacancies are not considered, the spread in thenumber of vacancies is still large, varying from 100 up to 600, i.e.there is still a variation of between 50% and 100%.

We now present the results for the bulk geometry. Note that theslight differences between the mean value of vacancies and self-interstitials (see Table 2) result from an inaccuracy in the methodof calculating defects when large clusters are formed. In bulk cas-cades the number of defects produced does not seem to depend onthe recoil direction. This applies to both interatomic potentials. Inthe case of the DD potential, the numbers of vacancies are 135 and129 for recoil angles of 10 and 22, respectively. The valuesobtained for the AM potential are slightly higher: 159 and 164

Page 78: Simulación atomística de la producción y evolución de ...

Fig. 4. Snapshot of the Fe bulk microstructure following 100 keV cascades at 0 K and at an incidence angle of 22 for two different azimuthal angles and initial velocitydistributions (a and b). Case (a) shows the formation of small clusters only, while larger vacancy type clusters appear in case (b). Green/light spheres: vacancies, red/darkspheres: interstitials. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Table 2Average number of vacancies/interstitials following the displacement cascade in bccFe at 0 K, depending on the irradiation and simulation condition.

Energy(keV)

Incidenceangle ()

Thin foil(number of vacancies/interstitials)

Bulk(Number of vacancies/interstitials)

DDpotential

AMpotential

DDpotential

AMpotential

50 10 173/135 190/172 135/125 159/16422 130/125 237/160 129/122 164/168

100 10 257/148 450/350 – –22 380/267 364/285 – 333/334

26 M.J. Aliaga et al. / Acta Materialia 101 (2015) 22–30

respectively for the same angles. These values are consistent withthose obtained by Björkas [33], with 130 Frenkel pairs with the AMpotential and 131 with the DD potential, with calculations per-formed for random recoil directions and 300 K. The larger numberof defects obtained in our work compared to the values of Björkaset al. could be due to the higher temperature used by the latter,because in this case more recombinations would be expected. For100 keV recoils the number of vacancies that we obtain is 333.Stoller [38] reported 330 Frenkel pairs for the same recoil energybut calculated using a different interatomic potential. It shouldbe noted that the scatter in the data, or the difference in the totalnumber of point defects between runs for the same condition, is atmost 5%, which is less than that observed in the thin film (Fig. 5).

As seen in Table 3, for the bulk material the percentage ofvacancies and interstitials in clusters of any size is very similardespite the different angles of incidence, and this is also true for

Table 3Percentage of vacancies/interstitials in clusters following the displacement cascade inbcc Fe at 0 K, depending on the irradiation and simulation condition.

Energy(keV)

Incidenceangle ()

Thin foil(% in clustersvacancies/interstitials)

Bulk(% in clustersvacancies/interstitials)

DDpotential

AMpotential

DDpotential

AMpotential

50 10 52/34 55/38 40/53 39/3222 68/42 62/38 38/52 45/43

100 10 67/44 42/37 – –22 53/35 54/38 – 46/99

the two potentials studied, with values between 38% and 45% ofvacancies in clusters and between 32% and 53% for self-interstitials in clusters. Note, however, that the fraction of intersti-tials in clusters for the DD potential is slightly larger than the frac-tion of vacancies in clusters. In the case of thin films, the fraction ofvacancies in clusters is much larger than the fraction of self-interstitials in clusters in all cases. In addition, the increase in inci-dent angle from 10 to 22 increases the clustering of vacancies by13–30%.

The examples of the damage produced in thin foils (Figs. 1 and2) and in the bulk (Figs. 3 and 4) visually indicate that in thin films,compared to the bulk, larger vacancy clusters and smaller self-interstitial clusters are produced. Fig. 7 shows the cluster size dis-tribution obtained for the 50 keV cascades and 22 incidence anglewith the AM potential for thin films and bulk samples. Fig. 7(a) and (b) give the size distributions of clusters of vacancies andinterstitials, respectively. They show clearly that thin film irradia-tion produces much larger vacancy clusters and slightly smallerinterstitial ones than bulk material irradiation, with an averagesize of 140 vacancies and 30 interstitials for thin foil and 44 vacan-cies and 40 interstitials for the bulk. The clusters in the thin foilreach sizes that are visible in TEM.

Table 4 gives the mean value of the range of the initial energeticatom for the different conditions, excluding channeling cases. Forthe 50 keV ion in the Fe thin film with an incidence angle of 10the mean value is 20 nm for the AM potential and 21 nm for theDD potential. Both have a standard deviation of 12 nm. The valueof the range obtained from the usual SRIM code [40] for this energyand angle is 18 nm. Note that SRIM does not account for the crys-talline structure of the target. For the 22 impact angle and the DDpotential the ion range increases by as much as 200% when their

Table 4Ion range, in nm, depending on the irradiation and simulation condition. Highchanneling cases are not included in the average.

Energy(keV)

Incidenceangle ()

Thin foilIon range in nm

BulkIon range in nm

DDpotential

AMpotential

DDpotential

AMpotential

50 10 21 20 10 1122 7 11 12 8

100 10 22 26 – –22 21 27 – 32

Page 79: Simulación atomística de la producción y evolución de ...

Fig. 5. Distribution of the number of point defects (vacancies and self-interstitials) in Fe thin film following (a) a 50 keV Fe ion, 10 incidence angle and the AM potential; (b)the same condition as in (a) but with the DD potential; (c) a 100 keV Fe ion in thin films, 22 incidence angle and the AM potential; and (d) the same condition as in (c) butwith the DD potential.

Fig. 6. (a) Microstructure resulting from a 50 keV cascade in Fe thin film at 0 K and at a 10 incidence angle, with the AM potential illustrating the extension of the surfacedamage into the thin film; (b) image of the same cascade, slightly tilted to show the damage directly at the surface. Green/light spheres: vacancies, red/dark spheres:interstitials, yellow spheres: surface atoms. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

M.J. Aliaga et al. / Acta Materialia 101 (2015) 22–30 27

energy is increased from 50 keV to 100 keV. As expected, the meanvalue of the ion range decreases when the incidence angleincreases. Fig. 8 shows the distribution of the ion range obtainedfor the 50 keV ion in the Fe thin film (Fig. 8(a) and (b)) and in

the bulk (Fig. 8(c) and (d)), for all cases. Fig. 8(a) corresponds toan incidence angle of 10 with both the AM potential and DDpotential. As for the number of point defects, there is a wide spreadin the values obtained, with some ions reaching only 5 nm while

Page 80: Simulación atomística de la producción y evolución de ...

Fig. 7. Cluster size distribution of (a) vacancies and (b) self-interstitials comparing bulk cascades with those of thin films for 50 keV and 22 incidence angle with the AMpotential.

Fig. 8. Distribution of ion ranges for the two interatomic potentials studied (AM and DD) and for (a) 50 keV cascades in thin films and 10 incidence angle, (b) 50 keV cascadesin thin films and 22 incidence angle, (c) 50 keV cascades in bulk and 10 incidence angle, and (d) 50 keV cascades in bulk and 22 incidence angle.

28 M.J. Aliaga et al. / Acta Materialia 101 (2015) 22–30

others cross the whole sample thickness. In some cases theimplanted ion is backscattered after colliding with the surface.Fig. 8(b), presenting the results for 50 keV ions at 22 incidenceangle, again shows extensive scatter in the ion range, in particularfor the AM potential. The ion range is on average 11 nm for the AMpotential, with a standard deviation of 8 nm, and 7 nm for the DDpotential, with a standard deviation of 2 nm.

Fig. 8(c) and (d) show the range distribution of the 50 keV Feion in bulk Fe for incidence angles of 10 and 22, respectively.The ion range distribution is narrower than for the thin film.

However, some channeling can be observed (particularly in thecase of 10 implantation), as expected from the Lindhard relation-ship [34], which leads to a greater ion range. As seen in Table 4,the mean values for the range obtained with the DD potentialare similar for the two incidences studied: 10 nm with astandard deviation of 5 nm at 10 incidence and 12 nm with astandard deviation of 8 nm at 22 incidence. Similar values areobtained with the AM potential: 11 nm with a standard deviationof 9 nm at 10 incidence and 8 nm with a standard deviation of3 nm at 22 incidence.

Page 81: Simulación atomística de la producción y evolución de ...

M.J. Aliaga et al. / Acta Materialia 101 (2015) 22–30 29

4. Comparison to experimental observations

The first observations of dislocation loops formed in iron byself-irradiation were performed by Masters [6] and further workwas done in the 1980s by Robertson et al. [41,42]. In ultrahigh-purity (UHP) Fe these loops are mostly of h100i type [12–16].For high doses (>2 1018 ions m2) they are considered to be ofinterstitial type [14,15]. However, the nature of the loops is diffi-cult to assess when they are smaller than 5 nm. Jenkins et al.[14] showed the formation of a0h100i vacancy loops in Fe closeto the surface after irradiation performed with Ni+, Ge+, Kr+, Xe+

and W+ and energies between 40 keV and 240 keV. More recently,Yao et al. [13] studied dislocation loops induced by 30 keV Ga+ ionsin Fe–11%Cr, which produced damage within 10 nm from the sur-face. They were able to determine that the loops (at least thoseclosest to the surface) are of vacancy type.

For comparison with this experimental evidence we simulated a30 keV Ga+ ion implanted in an a-Fe matrix. The AM interatomicpotential was used for Fe–Fe interactions, while the interactionbetween the Ga+ ion and the Fe atoms was calculated using a purerepulsive potential, the so-called Universal potential described in[43]. In this way the damage produced by this ion as it travelsthrough the lattice is described well. Fig. 9 shows the defects pro-duced by the 30 keV Ga ion after 17 ps. The total number of vacan-cies in this case is 799 and the total number of interstitials is 107. Alarge ad-atom island at the surface with 627 ad-atoms is seen. Theformation of a large h100i vacancy loop close to the surface is alsoclearly observed. It has 692 vacancies and is approximately 15a0 9 a0 21 a0, or 4 nm 3 nm 6 nm. This cluster is compara-ble to those observed experimentally by Yao et al. [13], who per-formed these experiments with Fe–11%Cr. As shown, however, inMD calculations by Malerba et al. [44] the damage produced inthe cascade in a-Fe and FeCr alloys is not significantly different.The only difference is in the self-interstitial loops, which in FeCralloys can be a mixture of Fe and Cr atoms.

We should point out that experiments by Robertson et al.[41,42] using low-energy Fe ions (50 keV and 100 keV) in Fe showyields for the formation of loops much lower than those found inthe simulations presented here. A probable cause for this discrep-ancy is the difference in time scales between simulations andexperiments. The simulations have been performed for tens ofpicoseconds. For longer time scales, it is, however, conceivable thatvacancy loops close to the surface are able to climb and disappearby recombination, while self-interstitial clusters may coalesce andform larger loops, resulting in a yield lower than the one obtained

Fig. 9. Microstructure resulting from a cascade in an Fe thin film

in simulations. On the other hand, one should also keep in mindthat for loops smaller than 1 nm the contrast in CTEM is reducedand the image size saturates (because of the diffraction-limitedresolution), making the observation of loops more difficult, orimpossible [45], which may also explain the lower yields observedin the experiment.

5. Conclusions

The calculations presented here show that the damage pro-duced by ion implantation in Fe thin film using 50 keV and100 keV ions is significantly different from that produced in bulkFe by recoils of the same energy. In thin films, results show the for-mation of h100i vacancy loops with sizes visible in the TEM. Thisresults from the imbalance in the number of vacancies with respectto self-interstitials, due to the trapping of the latter at the surfacewhere they remain as ad-atoms.

Statistical analysis reveals a large dispersion in the defects pro-duced: while bulk results present a narrow dispersion in terms ofthe total number of defects or the percentage of defects in clusters,in thin foil the total number of defects varies significantly from onecascade to another, as does the morphology of the damage pro-duced. Two types of structures were identified in the thin foil. Onthe one hand, there are those exhibiting small self-interstitial clus-ters and large vacancy clusters right below the surface in the formof large h100i. On the other hand, there are structures presenting anarray of dislocations and ad-atoms. The latter are, however, rare.Both the AM and DD potentials induce equivalent results in termsof damage, although morphologically the DD potential producesmore frequent surface damage of the type shown in Fig. 6.

A larger fraction of vacancies in clusters are found in thin filmsthan in the bulk, independently of the energy, angle of incidence orinteratomic potential used. These vacancy clusters are also larger.The inverse behavior is observed for self-interstitials: in mostcases, the fraction of self-interstitials in clusters is higher in thebulk and their sizes are larger than in the thin film. Increasingthe energy from 50 keV to 100 keV shows the formation of sub-cascades in the case of bulk irradiation, resulting in smaller self-interstitial clusters for the higher energy. Sub-cascade formationat 100 keV in the thin film can be seen only in some cases.

The formation of large h100i vacancy loops directly in the cas-cade revealed here for Fe thin film agrees well with experiments.However, the nature of these loops has been experimentally iden-tified only in the case of irradiation of Fe–11%Cr with Ga+ ions oflower energy (30 keV) [13], or irradiation of Fe with heavy ions

at 0 K induced by a 30 keV Ga ion at 22 incidence angle.

Page 82: Simulación atomística de la producción y evolución de ...

30 M.J. Aliaga et al. / Acta Materialia 101 (2015) 22–30

(40–240 keV) [16]. Our simulations show that these loops alsoform when irradiating with Fe ions of 50 keV and 100 keV. Theseloops are, however, smaller than those produced in the Ga irradia-tion experiment at lower energy, and are therefore presumablymore difficult to observe and analyze experimentally.

In previous works [33,35] it is shown that the initial damage inthe cascade together with defect mobilities define how damagewill grow with dose. This has consequences for the modeling ofirradiation effects. In the quest to develop models that are ableto describe neutron damage in the bulk, ion implantation experi-ments using thin films are often used for validation. Our resultsshow that one should carefully account for the effect of free sur-faces in these models.

Acknowledgments

We would like to thank Drs. A. Prokhodtseva, M. Hernández-Mayoral, Z. Yao and S. Dudarev for fruitful discussions. Simulationswere carried out using the computer cluster of the Dept. of AppliedPhysics at the UA, the HPC-FF supercomputer of the Jülich Super-computer Center (Germany) and the Helios supercomputer at Rok-kasho (Japan). MJA thanks the UA for support through aninstitutional fellowship. The research leading to these results ispartly funded by the European Atomic Energy Community’s (Eura-tom) Seventh Framework Programme FP7/2007–2013 under Grantagreement No. 604862 (MatISSE project) and in the framework ofthe EERA (European Energy Research Alliance) Joint Programmeon Nuclear Materials and the Generalitat Valenciana PROME-TEO2012/011. This work has been carried out within the frame-work of the EUROfusion Consortium and has received fundingfrom the Euratom research and training programme 2014–2018under Grant agreement No. 633053. The views and opinionsexpressed herein do not necessarily reflect those of the EuropeanCommission.

References

[1] J.L. Boutard, A. Alamo, R. Lindau, M. Rieth, Fissile core and tritium-breedingblanket: structural materials and their requirements, C. R. Phys. 9 (2008) 287.

[2] S.J. Zinkle, G.E. Ice, M.K. Miller, S.J. Pennycook, Z.-L. Wang, Advances inmicrostructural characterization, J. Nucl. Mater. 386–388 (2009) 8–14.

[3] <http://jannus.in2p3.fr/>.[4] C.W. Allen, L.L. Funk, E.A. Ryan, New instrumentation in Argonne’s HVEM-

tandem facility: expanded capability for in situ ion beam studies, in: Mat. Res.Soc. Symp. Proc., vol. 396, 1996, p. 641.

[5] Y. Matsukawa, S.J. Zinkle, One-dimensional fast migration of vacancy clustersin metals, Science 318 (2007) 959.

[6] B.C. Masters, Dislocation loops in irradiated iron, Nature 200 (1963) 254.[7] W.M. Stobbs, C.H. Sworn, The weak beam technique as applied to the

determination of the stacking-fault energy of copper, Philos. Mag. 24 (1971)1365.

[8] R. Schaublin, J. Henry, Y. Dai, Helium and point defect accumulation: (i)Microstructure and mechanical behaviour, C. R. Phys. 9 (2008) 389–400.

[9] K. Arakawa, K. Ono, M. Isshiki, K. Mimura, M. Uchikoshi, H. Mori, Observationof the one-dimensional diffusion of nanometer-sized dislocation loops, Science318 (2007) 956.

[10] N. Soneda, T. Diaz de la Rubia, Defect production, annealing kinetics anddamage evolution in alpha-Fe: an atomic-scale computer simulation, Philos.Mag. A 78 (1998) 995.

[11] Yu.N. Osetsky, D.J. Bacon, A. Serra, B.N. Singh, S.I. Golubov, Stability andmobility of defect clusters and dislocation loops in metals, J. Nucl. Mater. 276(2000) 65.

[12] A. Prokhodtseva, B. Décamps, R. Schäublin, Comparison between bulk and thinfoil ion irradiation of ultra high purity Fe, J. Nucl. Mater. 442 (2013) S786–S789.

[13] Z. Yao, M. Hernández Mayoral, M.L. Jenkins, M.A. Kirk, Heavy-ion irradiationsof Fe and Fe–Cr model alloys Part 1: Damage evolution in thin-foils at lowerdoses, Philos. Mag. 88 (2008) 2851.

[14] M.L. Jenkins, Z. Yao, M. Hernández-Mayoral, M.A. Kirk, Dynamic observationsof heavy-ion damage in Fe and Fe–Cr alloys, J. Nucl. Mater. 389 (2009) 197–202.

[15] M. Hernández-Mayoral, Z. Yao, M.L. Jenkins, M.A. Kirk, Heavy-ion irradiationsof Fe and Fe–Cr model alloys Part 2: Damage evolution in thin-foils at higherdoses, Philos. Mag. 88 (2008) 2881.

[16] M.L. Jenkins, C.A. English, B.L. Eyre, Heavy-ion damage in alpha Fe, Nature 263(1976) 400–401.

[17] D.J. Bacon, A.F. Calder, F. Gao, Defect production due to displacement cascadesin metals as revealed by computer, J. Nucl. Mater. 251 (1997) 1.

[18] R.E. Stoller, G.R. Odette, B.D. Wirth, Primary damage formation in bcc iron, J.Nucl. Mater. 251 (1997) 49–60.

[19] N. Soneda, S. Ishino, T.D. de la Rubia, Vacancy loop formation by ‘cascadecollapse’ in alpha-Fe: a molecular dynamics study of 50 keV cascades, Philos.Mag. Lett. 81 (2001) 649.

[20] E. Zarkadoula, S.L. Daraszewicz, D.M. Duffy, M.A. Seaton, I.T. Todorov, K.Nordlund, M.T. Dove, K. Trachenko, The nature of high-energy radiationdamage in iron, J. Phys.: Condens. Matter 25 (2013) 125402.

[21] L. Malerba, Molecular dynamics simulation of displacement cascades in alpha-Fe: a critical review, J. Nucl. Mater. 351 (2006) 28–38.

[22] A.F. Calder, D.J. Bacon, A.V. Barashev, Yu.N. Osetsky, On the origin of largeinterstitial clusters in displacement cascades, Philos. Mag. 90 (2010) 863.

[23] A.E. Sand, S.L. Dudarev, K. Nordlund, High-energy collision cascades intungsten: dislocation loops structure and clustering scaling laws, Eur. Phys.Lett. 103 (2013) 46003.

[24] M. Ghaly, R.S. Averback, Effect of viscous-flow on ion damage near solid-surfaces, Phys. Rev. Lett. 72 (1994) 364.

[25] K. Nordlund, J. Keinonen, M. Ghaly, R.S. Averback, Coherent displacement ofatoms during ion irradiation, Nature 398 (1999) 49.

[26] K. Nordlund, J. Keinonen, M. Ghaly, R.S. Averback, Recoils, flows andexplosions: surface damage mechanisms in metals and semiconductorsduring 50 eV–50 keV ion bombardment, Nucl. Instrum. Methods Phys. Res. B148 (1999) 74–82.

[27] S.V. Starikov, Z. Insepov, J. Rest, A.Y. Kuksin, G.E. Norman, V.V. Stegailov, A.V.Yanikin, Radiation-induced damage and evolution of defects in Mo, Phys. Rev.B 84 (2011) 104109.

[28] Y.N. Osetsky, A.F. Calder, R.E. Stoller, How do energetic ions damage metallicsurfaces? Curr. Opin. Solid State Mater. Sci. 19 (2015) 277–286.

[29] T. Diaz de la Rubia, M.W. Guinan, New mechanism of defect production inmetals: a molecular-dynamics study of interstitial-dislocation-loop formationin high-energy displacement cascades, Phys. Rev. Lett. 66 (1991) 2766.

[30] S. Dudarev, P. Derlet, A ‘magnetic’ interatomic potential for moleculardynamics simulations, J. Phys: Condens. Matter 17 (2005) 1–22.

[31] G.J. Ackland, M.I. Mendelev, D.J. Srolovitz, S. Han, A.V. Barashev, Developmentof an interatomic potential for phosphorus impurities in alpha-iron, J. Phys:Condens. Matter 16 (2004) S2629.

[32] C. Björkas, K. Nordlund, Comparative study of cascade damage in Fe simulatedwith recent potentials, Nucl. Instrum. Methods Phys. Res. B 259 (2007) 853–860.

[33] C. Björkas, K. Nordlund, M.J. Caturla, Influence of the picosecond defectdistribution on damage accumulation in irradiated alpha-Fe, Phys. Rev. B 85(2012) 024105.

[34] L.-P. Zheng, Z.Y. Zhu, Y. Li, F.O. Goodman, An improved critical angle equationfor channeling, Nucl. Instrum. Methods Phys. Res. B 268 (2010) 120–122.

[35] M.J. Caturla, N. Soneda, E. Alonso, B.D. Wirth, T.D. de la Rubia, J.M. Perlado,Comparative study of radiation damage accumulation in Cu and Fe, J. Nucl.Mater. 276 (2000) 13.

[36] M.R. Gilbert, S.L. Dudarev, P.M. Derlet, D.G. Pettifor, Structure andmetastability of mesoscopic vacancy and interstitial loop defects in iron andtungsten, J. Phys.: Condens. Matter 20 (2008) 345214.

[37] M.R. Gilbert, Z. Yao, M.A. Kirk, M.L. Jenkins, S.L. Dudarev, Vacancy defects in Fe:comparison between simulation and experiment, J. Nucl. Mater. 386–388(2009) 36–40.

[38] V.G. Kapinos, Y.N. Osetskii, P.A. Platonov, The mechanism of nucleation ofvacancy loops with Burgers vectors (100) in bcc metals, J. Nucl. Mater. 173(1990) 229.

[39] R.E. Stoller, in: R.J.M. Konings (Ed.), Comprehensive Nuclear Materials, vol. 1,Elsevier, Amsterdam, 2012, p. 293.

[40] J.F. Ziegler, J.P. Biersack, The stopping and range of ions in matter, SRIM-2003,Ó1998, 1999 by IB77M co.

[41] I.M. Robertson, M.A. Kirk, W.E. King, Formation of dislocation loops in iron byself-ion irradiations at 40 K, Scr. Met. 18 (1984) 317.

[42] M.A. Kirk, I.M. Robertson, M.L. Jenkins, C.A. English, T.J. Black, J.S. Vetrano, Thecollapse of defect cascades to dislocation loops, J. Nucl. Mater. 149 (1987) 21;Kirk, et al., ASTM Spec. Tech. Pub. 955 (1987) 48.

[43] J.P. Biersack, J.F. Ziegler, Refined universal potentials in atomic-collisions, Nucl.Instrum. Methods 194 (1982) 93.

[44] L. Malerba, D. Terentyev, P. Olsson, R. Chakarova, J. Wallenius, Moleculardynamics simulation of displacement cascades in Fe–Cr alloys, J. Nucl. Mater.329–333 (2004) 1156.

[45] R. Schäublin, A. Almazouzi, Y. Dai, Yu.N. Osetsky, M. Victoria, Quantitativeanalysis of CTEM images of small dislocation loops in Al and stacking faulttetrahedral in Cu generated by molecular dynamics simulation, J. Nucl. Mater.276 (2000) 251.

Page 83: Simulación atomística de la producción y evolución de ...
Page 84: Simulación atomística de la producción y evolución de ...

IV

Page 85: Simulación atomística de la producción y evolución de ...
Page 86: Simulación atomística de la producción y evolución de ...

Copyright line will be provided by the publisher

pss-Header will be provided by the publisher Review copy – not for distribution

(pss-logo will be inserted here

by the publisher)

Influence of free surfaces on

microstructure evolution of radiation

damage in Fe from molecular dynamics and object kinetic

Monte Carlo calculations

Maria J. Aliaga *,1

I. Dopico2, I. Martin-Bragado

2, Maria J. Caturla

,1

1 Dept. Física Aplicada, Facultad de Ciencias, Fase II, Universidad de Alicante, Alicante E-03690

2 IMDEA Materials Institute, C/Eric Kandel 2, 28906 Getafe, Madrid, Spain

Received ZZZ, revised ZZZ, accepted ZZZ

Published online ZZZ (Dates will be provided by the publisher.)

Keywords Monte Carlo, molecular dynamics, fusion materials, radiation damage

* Corresponding author: e-mail [email protected], Phone: +34 96590 3400, Fax: +34 965909726

The influence of surfaces on the evolution of damage of

irradiated Fe is studied using object kinetic Monte Carlo

with input from molecular dynamics simulations and ab

initio calculations. Two effects are analysed: the influ-

ence of traps and the initial distribution of damage in the

cascade. These simulations show that for a trap concen-

tration of around 100appm, there are no significant dif-

ferences between defect concentrations in bulk and thin

films. However, the initial distribution of defects plays an

important role not only on total defect concentration but

also on defect type, for the model used in this study.

20 n

m

34 nm

100 keVFe ion

22o

Free surface

Interstitials

Vacancies

Damage produced by a 100keV Fe ion impinging a Fe

thin film. Blue (dark) spheres are self-interstitials, red

(light) spheres are vacancies.

Copyright line will be provided by the publisher

Page 87: Simulación atomística de la producción y evolución de ...

R2

Copyright line will be provided by the publisher

1 Introduction

Ion implantation is used to understand the effect of

damage production in materials for applications in fusion

and fission energy production [1]. Unlike neutron irradia-

tion, ion implantation allows for better control of variables

such as irradiation temperature, dose, energy and, to some

extent, dose rate. However, the extrapolation of results

from ion implantation to neutron irradiation regarding de-

fect production and microstructure evolution is not

straightforward. Differences in dose rate or sample thick-

ness could affect significantly the evolution of the damage.

The influence of the surface on defect production, defect

distribution and damage evolution needs to be understood

in order to develop reliable models that can extrapolate the

results from ion implantation to neutron irradiation condi-

tions.

Atomistic models have proven to be an appropriate

technique to predict the evolution of the irradiation damage

in materials. In this context, we present here a study, using

a recently developed object kinetic Monte Carlo (OKMC)

code, of the effect of surfaces on defect evolution in irradi-

ated Fe. We consider, on one hand, the interplay between

trapping sites for defects and surfaces,. and on the other

hand the influence of the initial defect distribution of de-

fects. It is well known from the early 1990s, that damage

produced close to the surface by an energetic ion in f.c.c.

metals gives rise to defect structures that are significantly

different from those produced in the bulk [2]. Recent simu-

lations in Fe [3] and in Mo [4] have shown that this is also

the case in b.c.c. metals, where large vacancy loops close

to the surface have been identified. In order to study how

this initial damage distribution (during the first few picose-

conds) affects the long term defect evolution, we have used

two databases of cascades, one obtained in bulk Fe and an-

other one obtained in thin films with free surfaces [3].

OKMC simulations have been performed to calculate the

defect concentration, type and size as a function of dose for

different conditions: i) bulk cascades with periodic bound-

ary conditions, to consider the case of recoils within a bulk

sample, ii) bulk cascades in the presence of free surfaces,

which would correspond to implantation at high energies

such that the initial damage is not affected by the surfaces,

and iii) surface damage in a thin film, that takes inot ac-

count the low enery irradiation close to the surface. In sec-

tion 2 we describe the model for damage accumulation and

growth used in this work. Section 3 describes the results

for the different conditions studied which are finally dis-

cussed in section 4.

2 Model for damage accumulation

Values obtained from ab initio calculations [5] and molec-

ular dynamics simulations [6] are used for migration and

binding energies of vacancies and self-interstitials, as

shown in table 1. All mobile defects in table 1 are consid-

ered to move in three dimensions except for ½<111> loops

which move in one-dimension.

All clusters can grow by addition of other defects of the

same type. Recombination occurs between vacancy and in-

terstitial type of defects, whether isolated or in clusters. In

this particular model, the formation of <100> loops occurs

through the interaction between ½<111> loops, following

the atomistic simulations of Marian et. al [7] and, more re-

cently, Terentyev et al [8]. This assumption is not the only

one possible for considering the formation of <100> loops.

Recently it was proposed that <100> loops can be formed

from the nucleation of C15 clusters formed in the collision

cascade [9, 10]. However, the aim of this work is not to

discuss about the model for loop formation but about the

effect of surfaces on a particular microstructure evolution

model. For such a study we have selected the first model of

loop formation; that of coalescence of ½<111> loops. This

model considers that a <100> loop can be formed as long

as the two ½<111> loops interacting have similar sizes,

with a maximum difference of 5%. Once the <100> loop is

formed it can grow by adding new ½<111> loops, <100>

loops or small self-interstitial clusters.

Table 1 Migration energies (Em) and binding energies (Eb) con-

sidered for different defects in the OKMC model (n is the number

of defects in a cluster).

Defect type Em (eV) Eb (eV)

V 0.67*

V2 0.62* 0.3*

V3 0.35* 0.37*

V4 0.48* 0.62*

Vn>4 Immobile **

I 0.34*

I2 0.42* 0.8*

I3 0.43* 0.92*

I4 0.3* 1.64*

In>4 ½<111> 0.06+0.11/n1.6 **

In>4 <100> Immobile **

* [5] ** [6]

Self-interstitial clusters larger than 4 defects are considered

to be ½<111> loops and therefore move with a very low

migration energy barrier and in one-dimension. As a result,

unless some traps are considered in the matrix, these loops

quickly migrate to the surfaces and recombine leaving no

residual damage. This, although a desirable situation from

a radiation resistance point of view, it is not a realistic sce-

nario. In a real sample, there is always some amount of

impurities such as carbon. Therefore, it is necessary to in-

clude traps in the simulation, to consider the effect of dis-

Page 88: Simulación atomística de la producción y evolución de ...

pss-Header will be provided by the publisher 3

Copyright line will be provided by the publisher

1

2

3 4

5 6

7

8 9

10 11

12 13

14

15 16

17 18

19

20 21

22 23

24 25

26

27 28

29 30

31

32 33

34 35

36 37

38

39 40

41 42

43

44 45

46 47

48 49

50

51 52

53 54

55

56 57

persed carbon. In this model ½<111> loops can be stopped

when they find a trap, distributed randomly in the sample,

with a binding energy of 1 eV. This trapped ½<111> loop

is then immobile and can grow by addition of other loops

or small self-interstitial clusters or monointerstitials.

The initial damage distribution produced by the energetic

particle is obtained from MD simulations [3]. We use two

databases of cascade damage: one for damage produced by

a recoil in the bulk and a second set of cascades produced

by launching an energetic Fe ion on a Fe thin film (of

about 50 nm thickness). In the first type of simulations pe-

riodic boundary conditions (PBC) were applied in all three

directions while in the second type PBC were applied only

in two directions and free surfaces are considered in the

third one. All simulations were performed with the intera-

tomic potential of Ackland and Mendelev [11].

The object kinetic Monte Carlo simulator MMonCa [13]

was used in this study. Unlike other OKMC codes, where

clusters are defined as an entity with a particular capture

radius and number of defects, where the individual defect

position is not considered.MMonCa tracks the position of

each individual defect within the cluster. The shape of the

cluster is defined as a property and determines how defects

are distributed within the cluster. Shapes can then be

spherical, two dimensional disks or amorphous. These

shapes can be selected according to experiments, when

known, or other simulations. As a consequence, there is

not a single capture radius associated to a cluster, but each

defect has its own capture radius. In order to select the

most appropriate value for this capture radius we have

compared the defect morphology as obtained from the MD

simulations with the one in the OKMC calculations just be-

fore any diffusion event occurs Figure 1(a) shows the dis-

tribution of vacancies and self-interstitials obtained from a

MD simulation of a 50 keV Fe recoil in Fe. Red (light)

dots correspond to the location of vacancies while blue

(dark) dots correspond to self-interstitials. The position of

these defects is calculated using a Wigner-Seitz cell algo-

rithm. Figure 1(b) shows the location of those vacancies

and self-interstitials as given by the OKMC code when a

capture radius of 0.4 nm is used. This capture radius has

been calibrated to agree with the number of monovacan-

cies and monointerstitials from the MD simulation (con-

sidering that two defects that are at first-nearest neighbours

distance belong to the same cluster). If a larger capture ra-

dius is used in the OKMC, immediate recombination be-

tween vacancies and self-interstitials occurs and the total

number of mono-defects is lower than in the MD results. If

a shorter capture radius is used, the number of defects

identified as monovacancies or monointerstitials is larger

than those used to identify clusters from the MD simula-

tions. Nevertheless, one must keep in mind that this criteri-

on for clustering of defects is somehow arbitrary, since

other cut-off distances could be used. An evaluation of the

damage distribution of defects from MD and OKMC, such

as those shown in figure 1, reveals a good agreement for

the 0.4 nm cut-offradius .

Figure 1 Defect distribution as obtained from MD simulations of

a 50keV Fe ion in Fe (a) and as initial conditions for the OKMC

calculations (b). Red (light) spheres represent the location of va-

cancies, blue (dark) spheres are self-interstitials.

3 Results

Using the model described above, we have studied the

evolution of the damage produced by energetic recoils in

Fe. We have analysed both the influence of surfaces on de-

fect concentration and defect size and the influence of the

initial defect distribution. As mentioned above, recent sim-

ulations of cascades in Fe have shown that damage pro-

duced by ion implantation with low energies (~100keV)

results in defect structures significantly different from

those produced by the same energy recoils but in the bulk

of the material [3]. One important difference is the for-

mation of large (> 1nm) vacancy clusters of <100> type

when the damage is very close to the surface, together with

smaller self-interstitial clusters as compared to bulk dam-

age. Those results, however, only consider the first few pi-

coseconds after the energy is transferred from the recoil to

the lattice. Here, we follow the evolution of those defects

produced in the cascade over longer times and under con-

tinuous irradiation with the use of the OKMC model.

Two data bases for cascade damage were used for the-

se calculations, both obtained with the same interactomic

potential [11]: one of 100keV recoils in bulk Fe and anoth-

er one of 100keV Fe ion implantation in Fe thin films [3].

In order to decouple the effect of surfaces from the effect

of the initial damage distribution we have performed three

types of calculations: (1) bulk cascades with periodic

boundary conditions, (2) bulk cascades in a thin-, and (3)

surface cascades in a thin-film. Thin films have a thickness

of 50 nm, similar to those used in in situ transmission elec-

tron microscopy (TEM) studies. The simulation box was

200 nm x 200 nm x 50 nm and cascades were located ran-

domly within this box with a dose rate of 8 x 1014

ions m-2

s-1

. The same calculation is performed introducing two

concentrations of traps in the lattice: 1 appm and 118 appm,

corresponding to a highly pure sample and a high concen-

tration of carbon sample. Self-interstitial defects bind to

Page 89: Simulación atomística de la producción y evolución de ...

4 Author, Author, and Author: Short title

Copyright line will be provided by the publisher

these traps with a binding energy of 1eV, which should

mimic the effect of having carbon in the sample [13].

Figure 2 shows the total concentration of visible clus-

ters as a function of dose for the case of 1 appm traps and

the three different simulations: bulk cascades with PBC,

bulk cascades in thin films and surface cascades in thin

films. Figure 1(a) represents the concentration of ½ <111>

loops while figure 1(b) is the total number of <100> loops.

Clusters are considered visible if they have more than 100

defects (~1nm radius). As expected, the total number of

clusters when PBC are imposed is higher than when we

consider a thin film. Also, visible clusters appear at lower

doses when PBC are used. This is due to the fast migration

of ½<111> loops to surfaces in the thin film, that lower the

total defect concentration. Notice that no sinks were in-

cluded in the simulations with PBC. Comparing the con-

centration of ½<111> loops for the two cases with free sur-

faces, bulk cascades and surface cascades, no significant

differences are observed, the total concentration is quite

similar in these two examples. However, there is an im-

portant difference regarding the formation of <100> loops.

When bulk cascades are considered almost no <100> loops

are formed but, when surface cascades are used, the con-

centration of <100> loops is comparable to that of ½<111>

loops. This is surprising at first since the size of the self-

interstitial clusters in surface cascades is smaller than in

the case of bulk cascades. Considering that the model that

we are using here for <100> loop formation is the recom-

bination of two ½ <111> loops, the reason for this differ-

ence has to be the higher probability of two small self-

interstitial loops finding each other before reaching the sur-

face or getting trapped by a carbon in the case of surface

cascades.

Figure 2 Concentration of visible (> 100 defects) self-interstitial

clusters as a function of irradiation dose for 1 appm concentration

of traps (a) ½<111> loops and (b) <100> loops. Three different

cases are considered: irradiation in bulk (squares), irradiation in a

thin film with bulk cascades (circles) and irradiation in a thin film

with surface cascades (triangles).

This, in fact, can be rationalized in terms of the distri-

bution of defects within the cascade in the case of bulk

damage or surface defects. Figure 3 shows one example of

a cascade in the bulk (figure 3(a)) and a surface cascade

(figure 3(b)). As it can be seen, bulk cascades are spread

over a longer range while surface cascades are more local-

ized, in this particular case confined to a region of only 34

x 20 x 26 nm. Therefore, there is a much higher probability

for two self-interstitial clusters to interact and form a

<100> loop before reaching the surface or a trap in the case

of surface cascades than in the case of bulk damage.

(a)

(b)70

nm

30 nm

20

nm

34 nm

Figure 3 Defect distribution as obtained from MD simulations

for 100 keV and 22o angle (a) Fe recoil in bulk Fe and (b) Fe ion

implanted in an Fe matrix. Red (light) spheres represent the loca-

tion of vacancies, blue (dark) spheres are self-interstitials.

Calculations were also performed for higher concentration

of trapping sites, 118 appm. Figure 4 shows the concentra-

tion of ½<111> loops (figure 4(a)) and <100> loops (fig-

ure 4(b)) for the three cases considered in this study. Now,

the concentration of ½ <111> loops is almost the same for

the three cases. That is, the effect of the surface is negligi-

ble since all loops were trapped before they can reach the

surface. However, the difference in the concentration of

<100> loops is still clear between bulk and surface cas-

cades, since this is an effect of cascade damage distribution

and therefore almost independent of trapping concentration.

Figure 4 Concentration of visible (> 100 defects) self-interstitial

clusters as a function of irradiation dose for 118 appm concentra-

tion of traps (a) ½<111> loops and (b) <100> loops. Three dif-

ferent cases are considered: irradiation in bulk (squares), irradia-

tion in a thin film with bulk cascades (circles) and irradiation in a

thin film with surface cascades (triangles).

Page 90: Simulación atomística de la producción y evolución de ...

pss-Header will be provided by the publisher 5

Copyright line will be provided by the publisher

1

2

3 4

5 6

7

8 9

10 11

12 13

14

15 16

17 18

19

20 21

22 23

24 25

26

27 28

29 30

31

32 33

34 35

36 37

38

39 40

41 42

43

44 45

46 47

48 49

50

51 52

53 54

55

56 57

The different behaviour of <100> loops with respect to ½

<111> loops in this particular model are also observed in

the average size of these loops as a function of dose, pre-

sented in figure 5. Results are shown for the cases of thin

films with bulk cascades and with surface cascades. In the

case of surface cascades, for any given dose, the average

size of the clusters is larger than for the case of bulk cas-

cades. Loops of the ½ <111> type have a constant average

size for low doses but after a certain dose, which is lower

for surface cascades, the average size of these loops in-

creases rapidly with dose. For the case of <100> loops, the

average size remains constant for all doses studied in this

work. The reason for these differences is the nucleation

mechanism considered by the model. <100> loops grow

through the reaction between ½ <111> loops of similar

size, therefore, nucleation sites for these loops are con-

stantly forming as long as the material is being irradiated.

Since ½<111> loops are highly mobile, they can either mi-

grate to the surface, interact to form <100> loops or be-

come trapped. When all trapping sites are saturated, no

more nucleation sites for ½ <111> loops can be created

and the ones that are already trapped can rapidly grow by

the addition of new ½<111> loops. That gives rise to the

rapid increase of the average size of these loops at high

doses.

Figure 5 Average cluster size as a function of dose for the cases

of bulk cascade in thin films and surface cascades in thin films.

Values for ½<111> and <100> loops.

4 Conclusions This work shows the importance of

surfaces on the microstructure evolution of damage pro-

duced by irradiation. Surfaces act as sinks for defects, and,

as expected, lower the total concentration of defects com-

pared to bulk irradiation. This effect, however, will depend

strongly on the purity of the sample and the presence of

traps. More interestingly, if damage is produced very close

to the surface, the distribution of this damage differs from

that of bulk irradiation, resulting in, not only a different

concentration of defects, but also differences in the type of

damage that can be observed. In this particular model the

differences are mostly related to the ratio of ½<111> to

<100> loops. Irradiation close to the surface favours the

formation of <100> loops due to the localization of the

damage within the cascade.

Acknowledgements Simulations were carried out using the

computer cluster of the Dept. of Applied Physics at the UA, the

HPC-FF supercomputer of the Jülich Supercomputer Center

(Germany) and the Helios supercomputer at Rokkasho (Japan).

MJA thanks the UA for support through an institutional

fellowship. The research leading to these results is partly funded

by the European Atomic Energy Community’s (Euratom)

Seventh Framework Programme FP7/2007-2013 under grant

agreement No. 604862 (MatISSE project) ) and in the framework

of the EERA (European Energy Research Alliance) Joint

Programme on Nuclear Materials. This work has been carried out

within the framework of the EUROfusion Consortium and has

received funding from the Euratom research and training

programme 2014-2018 under grant agreement No 633053. The

views and opinions expressed herein do not necessarily reflect

those of the European Commission.

References

[1] J.L. Boutard, A. Alamo, R. Lindau, M. Rieth, C. R. Phys. 9

(2008) 287. [2] M. Ghaly, R.S. Averback, Effect of viscous-flow on ion

damage near solid-surfaces, Phys. Rev. Lett. 72 (1994) 364.

[3] M. J. Aliaga, R. Schäublin, J. F. Löffler, M. J. Caturla, Acta

Materialia 101 (2015) 22-30.

[4] S. V. Starikov, Z. Insepov, J. Rest, A. Y. Kuksin, G. E.

Norman, V. V. Stegailov, A. V. Yanikin, Radiation-induced

damage and evolution of defects in Mo, Physical Review B 84,

104109 (2011). [5]C. C. Fu, J. Dalla Torre, F. Willaime, J.-L.

Bouquet, A. Barbu, Nature Materials (2004)

[6] Soneda, N. & Díaz de la Rubia, T. Phil. Mag. A 78, 995-1019

(1998)

[7] J. Marian and Brian D. Wirth PRL 88, 25 (2002).

[8] Haixuan Xu, Roger E. Stoller, Yuri N. Osetsky, and Dmitry

Terentyev, PRL 110, 265503 (2013)..

[9] M.-C. Marinica, F. Williaime, J.-P. Crocombette, PRL 108,

025501 (2012)..

[10] Yongfeng Zhang, Xian-Ming Bai, Michael R. Tonks and S.

Bulent Biner, Scripta Materialia 98 (2015) 5-8.

[11] G.J. Ackland, M.I. Mendelev, D.J. Srolovitz, S. Han, A.V.

Barashev, Development of an interatomic potential for

phosphorus impurities in alpha-iron, J. Phys.: Condens. Matter 16

(2004) S2629.

[12] I. Martin-Bragado, Antonio Rivera, Gonzalo Valles, Jose

Luis Gomez-Selles, María J. Caturla Computer Physics

Comunications (2013).

[13] V. Jansson, M. Chiapetto, L. Malerba, Journal of Nuclear

Materials 442 (2013) 341-349.

Page 91: Simulación atomística de la producción y evolución de ...
Page 92: Simulación atomística de la producción y evolución de ...

5. TRABAJOS NO PUBLICADOS

Page 93: Simulación atomística de la producción y evolución de ...
Page 94: Simulación atomística de la producción y evolución de ...

V

Page 95: Simulación atomística de la producción y evolución de ...
Page 96: Simulación atomística de la producción y evolución de ...

epl draft

Surface effects and statistical laws of defects in primary radiationdamage: tungsten vs. iron

A. E. Sand1, M. J. Aliaga2, M. J. Caturla2 and K. Nordlund1

1 Department of Physics - P.O. Box 43, FI-00014 University of Helsinki, Finland2 Dept Fisica Aplicada - Facultad de Ciencias, Fase II, Universidad de Alicante, Alicante E-03690, Spain

PACS nn.mm.xx – First pacs descriptionPACS nn.mm.xx – Second pacs descriptionPACS nn.mm.xx – Third pacs description

Abstract – We have investigated the effect of surfaces on the statistics of the primary radiationdamage formed in the bcc metals iron (Fe) and tungsten (W). Through molecular dynamicssimulations of collision cascades we show that both interstitial and vacancy cluster sizes followscaling laws in these materials, in bulk as well as in thin foils. However, the slopes of the power lawdistributions in bulk Fe are markedly different from those in W, and furthermore the slope of thevacancy cluster size distribution in Fe is clearly affected by the surface in thin foil irradiation, whilein W mainly the overall frequency is affected. The distinct behaviour of the statistical distributionsuncovers different defect production mechanisms effective in the two materials, and provides insightinto the underlying reasons for the differing behaviour observed in TEM experiments of low-doseion irradiation in these metals.

Introduction. – One of the main challenges on theroad to commercial fusion power is presented by the needfor materials that can withstand the harsh conditions ina fusion reactor. Energetic fusion neutrons will cause sig-nificant damage to the wall materials of future reactors,leading to swelling, hardening and embrittlement. The de-velopment of materials that can withstand this irradiationand retain the structural integrity of the reactor requiresa thorough understanding of the radiation damage pro-cesses.

Two materials of prime interest in current reactor de-signs are iron (Fe), in steels for structural components, andtungsten (W) for plasma-facing components. These twometals, though both have bcc structure, exhibit markeddifferences in their response to radiation. While self-ionirradiation produces primary defects in W which are im-mediately visible in TEM experiments [1], in Fe nothingvisible is produced in either neutron or ion irradiation ex-periments until significant dose levels are reached [2–4].Nevertheless, indirect observations of low-dose radiationdamage using a combination of electron irradiation andneutron irradiation indicate that sub-microscopic defectclusters are initially formed also in Fe [4].

Molecular dynamics (MD) simulations confirm the for-mation of clusters directly from collision cascades in Fe

(see, e.g., [5–7]). In W, MD simulations have furthershown that the size-frequency distribution of interstitialclusters in bulk material follows a power law [8], a resultsupported by experiments [9]. The formation of clustersdirectly in cascades has a significant impact on the furtherevolution of the damage, and is therefore an importantfactor in microstructural evolution models.

While ion irradiation experiments serve as a usefulproxy for neutron irradiation, the close proximity of ma-terial surfaces in the former must be taken into account.The surface affects the evolution of the damage via imageforces, and by acting as a sink for defects, but also theinitial formation of defects is known to be affected by anearby surface [10]. As a result the accumulated damagein thin foils and bulk samples shows significant differences[11]. Surface effects are particularly important in the caseof in-situ TEM ion implantation experiments, since theirradiated sample must be less than 100 nm thick [12] tobe transparent to the electrons. They also play a majorrole in low energy (a few tens of keV) irradiation experi-ments, due to the shallow penetration depth of the inci-dent ions. In Fe, for example, vacancy loops have beenidentified close to the surface when irradiating with heavyions of low energy and at low doses [2, 13].

In this work, we investigate the effect of surfaces on the

p-1

Page 97: Simulación atomística de la producción y evolución de ...

A. E. Sand et al.

statistics of the defects constituting the primary radiationdamage in Fe and W. We also consider the differences inthe formation of the damage underlying the dissimilarityin observations of defects in the two materials. We use MDsimulations to study the experimentally invisible defectsize range. With a statistical analysis of results, we areable to shed light also on events which may occur toorarely to be directly captured by MD, due to the limitednumber of simulations that can be performed, in contrastto the thousands of impacts which are recorded in typicalTEM experiments.

Simulation methodology. – We have simulated fullcollision cascades in W and α-Fe using molecular dynamicsmethods. Simulations of bulk cascades were performed us-ing periodic boundary conditions in all directions, and bychoosing the primary knock-on atom (PKA) from amongthe lattice atoms. Thin foil irradiation was simulated withperiodic boundaries in two directions, and free surfaces inthe z-direction. An incident ion was placed above the sur-face, and given the desired kinetic energy in a chosen angletowards the surface.

Cascades in W were simulated with the MD code PAR-CAS [14], using the interatomic potential by Derlet et al.[15], with the short range part fitted by Bjorkas et al. [16].Bulk simulations were performed in a cubic cell of 48 nmto a side. The direction of the PKA in the bulk was var-ied randomly, with a uniform distribution over the unitsphere. Foil simulations were performed in a cell with di-mensions 48×48×65 A, where the lattice was oriented togive a (014)-surface. The incident ion trajectory formeda 15 degree angle with the surface normal. This geom-etry corresponds to that used in recent in-situ TEM ex-periments [9]. Simulations in Fe were performed usingthe MD code MDCASK, with the interatomic potentialof Dudarev and Derlet [17], modified for short range in-teractions following the procedure described in [18]. Bulksimulations were performed in a cubic cell of 34 nm oneach side, where the polar and the azimuthal angles of thePKA were varied for the different cases. Thin film simu-lations were performed in a cubic cell of 40 nm to a side,oriented along a 〈001〉 direction. The incident angle in thiscase was 22 degrees, which corresponds to the geometryused in [11].

Electronic stopping Se in the form of a friction termwas included in the simulations in W, since recent resultsindicate an effect of the dynamic treatment of electronicenergy losses on the residual damage [19]. The frictionterm follows the Lindhard model [20], and is independentof position, with the magnitude determined by SRIM cal-culations [21]. In W simulations it was applied to all atomswith a kinetic energy larger than 10 eV [19]. In Fe, tradi-tionally no electronic energy losses have been included incollision cascade simulations [5–7,18], and here we presentresults using that same convention. The effect on the de-fect statistics of including or excluding Se was neverthelessinvestigated for chosen conditions in each material, and is

presented in the last part of the results section. The wayof introducing Se in the Fe simulations is similar to that inW, with the friction term applied to atoms with a kineticenergy larger than 5 eV.

In the ballistic scenario of the binary collision approxi-mation, the number of defects NNRT produced from cas-cades depends on the initial PKA energy EPKA, the elec-tronic energy losses Eel, and the threshold displacementenergy (TDE) Ed, according to the Norgett-Robinson-Torrens (NRT) formula [22]

NNRT =0.8(EPKA − Eel)

2Ed, (1)

where the term in parentheses equals the damage energyEdam, i.e. the energy available to the ionic system. Inorder to compare cascade simulations in different materi-als, and those performed with and without electronic en-ergy losses, it is therefore reasonable to consider them interms of the reduced damage energy Er = Edam/Ed [23].We calculate the reduced damage energy from the TDEpredicted by the interatomic potential, determined as theminimum energy needed to displace an atom in a givendirection Emind (θ, φ), averaged over all directions [24]

Eavd =

∫ 2π

0

∫ π0Emind (θ, φ) sin θ dθ dφ∫ 2π

0

∫ π0

sin θ dθ dφ. (2)

For the potentials used here, Eavd is 84.5 eV for W [16]and 35 eV for Fe [18].

Residual defects were identified using a Wigner-Seitzcell method which determines the location of vacanciesand self-interstitials in a crystal lattice. Defects werethen grouped into clusters: two vacancies were consid-ered to be in the same cluster if the distance betweenthem was within the 2nd nearest-neighbor distance, whilethe 3rd nearest-neighbor distance was assumed for self-interstitials. Size-frequency distributions of defect clus-ters were determined by binning the data on the numberof occurrences of each cluster size into roughly logarith-mic bins, taking care that the bin width was sufficient toinclude at least a few data points in each bin.

Results. –

Scaling laws in Fe and W. Our results show that inboth Fe and W, the frequency f(N) per ion of the occur-rence of defect clusters of size N closely follows power lawsof the form

f(N) =A

NS, (3)

where N is the size of the defect in terms of the number ofpoint defects included in the cluster, and A is a frequencyscaling factor, in agreement with earlier work in W [8].However, when considering defects down to the smallestsizes, including single point defects, we find that two powerlaws emerge in several cases.

The size-frequency distribution of single point defectsand smaller clusters of size N . 10 follow scaling laws with

p-2

Page 98: Simulación atomística de la producción y evolución de ...

Surface effects and statistical laws of primary radiation damage

Fig. 1: (Color online) Comparison of scaling laws for SIA defectsizes in bulk and foil, W and Fe.

the same slope in Fe as in W. In addition, the frequencyof the defects is approximately the same in the two mate-rials, when comparing defects from cascades with similarreduced damage energy Er (see Figs 1 and 2).

The difference between Fe and W becomes apparent inthe distribution of the larger clusters. For bulk cascadesin Fe, larger clusters of both vacancy and interstitial typefollow the same scaling law as small clusters, while in bulkW, both vacancy and interstitial type clusters of size N &10 follow a scaling law with a lower value of S ≈ 1.6.Parameters for the best fit of the power laws are given inTable 1.

Surface effects. The effect of the surface on the dis-tribution of interstitial-type defects in both materials isminimal, but discernible as a slight preference for the for-mation of smaller defects, leading to a steeper slope inthe distributions. This preference arises from the portionof cascades which occur very close to the surface. Whenthe liquid core of the heat spike extends to the surface, itcauses the cascade to erupt, ejecting large amounts of ma-terial in the form of sputtered atoms and atom clusters.Such cascades form only very few and small interstitial-type clusters.

In the case of vacancy-type defects, the difference be-tween the bulk and foil cascade damage is clear, and es-pecially pronounced in Fe. Near-surface cascades readilyform large vacancy clusters, due to the ejection of material,and material flow to the surface causing an underdense re-gion to form in the core of the cascade. The size-frequecydistribution of these surface-induced vacancy clusters alsofollows a power law, with a slope that is roughly the sameas that for vacancy clusters in bulk W. Thus, in W, thesurface has the effect of simply increasing the overall fre-quency of vacancy-type defect clusters, maintaining thesame slope for the power law. In Fe, however, the surface

Fig. 2: (Color online) Vacancy cluster size distributions, fittedto power laws, for Fe and W in bulk and foil.

mechanism gives rise to a new scaling law for the largervacancy clusters, with S ≈ 1.6. In both materials, thesmallest vacancy-type defects still follow the same powerlaw as in bulk cascades, with S ≈ 3.0. Parameters forthe best fit of the power law to the distributions of defectclusters in thin foils are given in Table 2.

Electronic energy losses. We find that simulationswith and without electronic stopping in Fe result in thesame distributions for both vacancy- and interstitial-typedefects, as shown in Fig. 3 and Table 2. Only an over-all scaling of the frequency occurs due to the difference indamage energy with the two methods, from PKAs withthe same initial energy.

In W, however, the treatment of electronic energy lossesaffects the slope of the frequency-size distribution for bothinterstitial and vacancy clusters. Fig. 4 shows the distri-butions from 200 keV bulk cascades including Se in thesimulations, and for bulk cascades without Se with thesame total damage energy, Edam = 140 keV. The effect ofthe dynamic energy losses can be seen in the distributionof the larger clusters, which shows a decrease in the slopewhen Se is excluded (see Table 1). This effect is espe-cially apparent in the vacancy cluster distribution. Thedistributions of small clusters remains roughly the same.

Discussion. – The different scaling laws appearingin the defect distributions in W and Fe, and in bulk andthin foils, indicate the presence of different defect forma-tion mechanisms. The size-frequency distributions of thesmallest defects, in both bulk and foil cascades, follow thesame power laws in Fe and W. Furthermore, the frequencyof occurrence of these defects is similar for both materialsin simulations with similar reduced cascade energy, indi-cating a connection to the ballistic phase of the cascade.

In bulk W, a separate mechanism for the formation of

p-3

Page 99: Simulación atomística de la producción y evolución de ...

A. E. Sand et al.

Material EPKA (keV) Ered (keV) type A S No. cascades noteW bulk 150 1.3 SIA (small) 24.1 ± 1.0 2.3 ± 0.1 38 Se ≈ 43 keVW bulk 150 1.3 SIA (large) 5.7 ± 1.6 1.6 ± 0.1 38 Se ≈ 43 keVW bulk 150 1.3 vac (small) 59.4 ± 5.2 2.9 ± 0.1 38 Se ≈ 43 keVW bulk 150 1.3 vac (large) 1.0 ± 0.2 1.5 ± 0.1 38 Se ≈ 43 keVW bulk 200 1.7 SIA (small) 26.9 ± 1.6 2.2 ± 0.1 10 Se ≈ 60 keVW bulk 200 1.7 SIA (large) 9.3 ± 2.4 1.5 ± 0.1 10 Se ≈ 60 keVW bulk 200 1.7 vac (small) 149.3 ± 3.8 3.0 ± 0.05 10 Se ≈ 60 keVW bulk 200 1.7 vac (large) 23.4 ± 8.3 2.0 ± 0.2 10 Se ≈ 60 keVW bulk 140 1.7 SIA (small) 27.3 ± 2.3 2.5 ± 0.2 5 no SeW bulk 140 1.7 SIA (large) 2.0 ± 0.2 1.1 ± 0.03 5 no SeW bulk 140 1.7 vac (small) 152.0 ± 6.0 3.0 ± 0.1 5 no SeW bulk 140 1.7 vac (large) 0.6 ± 0.1 1.1 ± 0.1 5 no SeFe bulk 50 1.4 SIA (all) 53.1 ± 1.3 2.6 ± 0.2 18 no SeFe bulk 50 1.4 vac (all) 58.2 ± 1.2 2.7 ± 0.1 18 no Se

Table 1: Power law parameters for defects in bulk cascades.

Material EPKA (keV) type A S No. cascades noteW foil 150 SIA (small) 25.7 ± 2.0 2.4 ± 0.1 49W foil 150 SIA (large) 9.9 ± 1.5 1.8 ± 0.1 49W foil 150 vac (small) 135.1 ± 11.4 3.0 ± 0.1 49W foil 150 vac (large) 6.7 ± 1.0 1.7 ± 0.1 49Fe foil 50 SIA (all) 58.1 ± 1.2 2.8 ± 0.1 20Fe foil 50 vac (small) 104.7 ± 1.1 3.1 ± 0.1 20Fe foil 50 vac (large) 8.7 ± 1.3 1.7 ± 0.1 20Fe foil 100 SIA (all) 56.0 ± 1.4 2.3 ± 0.1 20Fe foil 100 vac (small) 153.8 ± 1.3 3.0 ± 0.2 20Fe foil 100 vac (large) 8.3 ± 2.4 1.7 ± 0.2 20Fe foil 100 SIA (all) 85.5 ± 1.2 2.9 ± 0.1 20 Se ≈ 33 keVFe foil 100 vac (small) 122.4 ± 1.3 3.0 ± 0.2 20 Se ≈ 33 keVFe foil 100 vac (large) 4.1 ± 2.8 1.6 ± 0.2 20 Se ≈ 33 keV

Table 2: Power law parameters for defects in foil cascades.

large defect clusters is apparent, which is absent in Fe.This formation mechanism is likely related to the energydensity of cascades, which is higher in W than in Fe, dueto the lower mass and lower subcascade splitting thresholdof the latter. A dependence on energy density is furtherdemonstrated by the sensitivity of the scaling law to themethod of treating electronic energy losses in W. The dif-ference is especially apparent in the vacancy cluster distri-bution, and cannot be ascribed to different cooling rates,since the Se energy losses take place exclusively during theinitial ballistic phase of the cascade [19], and thus do notaffect the rate of cooling of the heat spike. In fact the sizeof the liquid, in terms of the number of atoms with ener-gies exceeding the melting point, evolves similarly in bothcases. Rather, it is likely that the initially higher energy ofthe PKA and subsequent recoils in simulations where Se isincluded results in an increased probability for the energyto be deposited in a more wide spread region, as com-pared to simulations which do not include Se but instead

initiate the PKA with a kinetic energy corresponding toonly the damage energy. In total the energy deposited inthe ionic system is the same with the two methods, butthe higher likelyhood for compact energy deposition whenSe is excluded translates into an increase in large defects,and thus a decrease in the slope of the scaling law. Amechanism of defect formation depending on the cascadeenergy density is in agreement with experimental obser-vations [25] as well as MD simulations [6], showing thatlarger defects are formed from heavier projectiles, whichdeposit their kinetic energy in a more compact region.

A third mechanism for vacancy defect production occursin near-surface cascades, and involves flow of material tothe surface, leaving large underdense regions in the cas-cade core. This mechanism has been reported in previousstudies [10], and is present in both W and Fe. The effect ofthis surface-induced mechanism is especially pronouncedin Fe, since it introduces a different distribution, as com-pared to bulk, for large vacancy-type defects.

p-4

Page 100: Simulación atomística de la producción y evolución de ...

Surface effects and statistical laws of primary radiation damage

Fig. 3: (Color online) Scaling laws for vacancy (left) and SIA (right) defect sizes in Fe with and without electronic stopping.

The slope of the scaling law for large vacancy-type de-fects formed by the latter two mechanisms is the same.The size-frequency distribution is thus likely a result ofthe recrystallization processes taking place in the core ofthe heat spike, once the conditions for a depleted zonehave been met by the removal of material.

The different defect formation mechanisms present in Wand Fe means that cascade simulations respond differentlyto electronic energy losses. On the one hand, the effect ofelectronic energy losses that we observed in W indicatesthe importance of including electronic stopping in thesesimulations. In Fe, on the other hand, the standard prac-tice of excluding electronic energy losses is supported byour results, which indicate that the main factor affectingdamage production in Fe is the total damage energy, withlittle effect of the dynamics of energy removal.

The scaling laws found in this work show that no defectclusters large enough to be seen in TEM are likely to formin Fe directly from collision cascades in bulk. Thus visi-ble defects in bulk samples have likely formed as a resultof the thermal evolution of the invisible primary damage.In thin foil irradiation, however, the flow of material tothe surface in a heat spike causes the in-cascade forma-tion of large vacancy clusters. In MD simulations of cas-cades in bcc metals, such as α-Fe and W, SIA-type defectsgenerally cluster in 2-dimensional configurations, in otherwords as dislocation loops, while vacancies mainly form3-dimensional clusters. Such vacancy clusters are oftennot perfect voids, but rather form depleted zones, whichhave been directly observed in W [26,27] as a result of ionirradiation. The large vacancy clusters in Fe observed inour simulations may nevertheless become visible in TEMmicrographs after collapse due to cascade overlap, as hasbeen speculated in the literature (see, e.g., [25]). Corre-sponding large SIA defects do not form from this process,and thus SIA defects in Fe large enough to be visible musthave formed from coalescense and aggregation of smallerdefects.

On the other hand, in W, the slope of the scaling law in

the limit of large defect sizes gives a fairly large probabilityof in-cascade formation of visible defects also in the bulk,of both SIA and vacancy type. Since the proximity of thesurface in W foil irradiation also gives rise to the sameprocess as that present in Fe, which is reponsible for thecreation of additional large vacancy-type defects, the fre-quency of vacancy-type defects in foil irradiation is higherthan that in bulk, while SIA defects are formed with sim-ilar frequency as that in bulk. However, the formation oflarge SIA defects happens only in cascades which do noterupt through the surface, and thus only from ions thathave penetrated deeply into the sample. Thus small SIAdefect clusters are favoured in foil irradiation, since onlya percentage of ions penetrate past the surface to producelarge clusters, while all cascades produce small defects.

Conclusions. – We have shown that the size-frequency distribution of defects in the primary damageof both Fe and W follows power laws. A mechanism forthe production of large SIA and vacancy clusters depend-ing on the energy density of cascades is effective in self-ionor neutron irradiation in bulk W, but not in Fe. The ef-fect of nearby surfaces is the same in the two materials,but due to the different bulk behaviour, the impact of thesurface on statistics in Fe is more evident. The surfaceaffects the formation of large vacancy clusters, while thedistribution of single vacancies and small clusters remainslargely unaffected. The formation of SIA clusters is onlyslightly affected by the surface, with a preference for smallclusters in foil irradiation of both W and Fe.

∗ ∗ ∗

The authors thank Sergei Dudarev for valuable dis-cussions. This work has been carried out within theframework of the EUROfusion Consortium and has re-ceived funding from the Euratom research and trainingprogramme 2014-2018 under grant agreement No 633053.The views and opinions expressed herein do not necessarilyreflect those of the European Commission. MJA thanks

p-5

Page 101: Simulación atomística de la producción y evolución de ...

A. E. Sand et al.

Fig. 4: (Color online) Scaling laws for SIA (top) and vacancy(bottom) defect sizes in W with different treatment of elec-tronic stopping, from cascades with damage energy Edam =140 keV.

the UA for support through an institutional fellowship.Simulations were carried out using the computer clusterof the Dept. of Applied Physics at the UA, the HPC-FFsupercomputer of the Jlich Supercomputer Center (Ger-many), the Helios supercomputer at Rokkasho (Japan)and the supercomputers at CSC - IT Center for Science(Finland).

REFERENCES

[1] Yi X., Jenkins M. L., no M. B., Roberts S. G., ZhouZ. and Kirk M. A., Phil. Mag. A, 93 (2013) 1715.

[2] English C. A., Eyre B. L. and Jenkins M. L., Nature,263 (1976) 400401 10.1038/263400a0.

[3] Robertson I. M., King W. E. and Kirk M. A., ScriptaMetall., 18 (1984) 317.

[4] Yoshiie T., Satoh Y., Taoka H., Kojima S. and Kir-itani M., Journal of Nuclear Materials, 155157, Part 2(1988) 1098 .

[5] Stoller R. E., Odette G. R. and Wirth B. D., J.Nucl. Mater., 251 (1997) 49.

[6] Calder A., Bacon D., Barashev A. and Osetsky Y.,Phil. Mag., 90 (2010) 863.

[7] Terentyev D., Lagerstedt C., Olsson P., Nord-lund K., Wallenius J., Becquart C. and MalerbaL., J. Nucl. Mater., 351 (2006) 65 .

[8] Sand A. E., Dudarev S. L. and Nordlund K., EPL,103 (2013) 46003.

[9] Yi X., Sand A. E., Mason D. R., Kirk M. A.,Roberts S. G., Nordlund K. and Dudarev S. L.,EPL, 110 (2015) 36001.

[10] Ghaly M., Nordlund K. and Averback R. S., Phil.Mag. A, 79 (1999) 795.

[11] Prokhodtseva A., Dcamps B. and SchublinR., Journal of Nuclear Materials, 442 (2013)S786 FIFTEENTH INTERNATIONALCONFERENCE ON FUSION REACTORMATERIALS.

[12] Stobbs W. and Sworn C., Philos. Mag., 24 (1971) 1365.[13] Yao Z., Mayoral M. H., Jenkins M. and Kirk M.,

Philos. Mag., 88 (2008) 2851.[14] Nordlund K., parcas computer code. The main princi-

ples of the molecular dynamics algorithms are presentedin [10, 28]. The adaptive time step is the same as in [29](2006).

[15] Derlet P. M., Nguyen-Manh D. and Dudarev S. L.,Phys. Rev. B, 76 (2007) 054107.

[16] Bjorkas C., Nordlund K. and Dudarev S. L., Nucl.Instr. Meth. B, 267 (2009) 3204.

[17] L. D. S. and P. D., J. Phys. Condens. Matter., 17 (2005)1.

[18] Bjorkas C. and Nordlund K., Nucl. Instr. and Meth.B, 259 (2007) 853.

[19] Sand A., Nordlund K. and Dudarev S., J. Nucl.Mater., 455 (2014) 207 proceedings of the 16th Interna-tional Conference on Fusion Reactor Materials (ICFRM-16)Proceedings of the 16th International Conference onFusion Reactor Materials (ICFRM-16), Beijing, China,20th - 26th October, 2013.

[20] Lindhard J. and Sharff M., Phys. Rev., 124 (1961)124.

[21] Ziegler J. F., SRIM-2008.04 software package, availableonline at http://www.srim.org. (2008).

[22] Norgett M. J., Robinson M. T. and Torrens I. M.,Nucl. Eng. Des., 33 (1975) 50.

[23] Setyawan W., Selby A. P., Juslin N., Stoller R. E.,Wirth B. D. and Kurtz R. J., Journal of Physics: Con-densed Matter, 27 (2015) 225402.

[24] Nordlund K., Wallenius J. and Malerba L., Nucl.Instr. and Meth. B, 246 (2006) 322332.

[25] C.A.English and Jenkins M., Materials Science Forum,15-18 (1987) 1003.

[26] Wei C.-Y., Current M. I. and Seidman D. N., Phil.Mag. A, 44 (1981) 459491.

[27] Current M. I., Wei C.-Y. and Seidman D. N., Phil.Mag. A, 43 (1981) 103138.

p-6

Page 102: Simulación atomística de la producción y evolución de ...

Surface effects and statistical laws of primary radiation damage

[28] Nordlund K., Ghaly M., Averback R. S., CaturlaM., Diaz de la Rubia T. and Tarus J., Phys. Rev. B,57 (1998) 7556.

[29] Nordlund K., Computational Materials Science, 3(1995) 448 .

p-7

Page 103: Simulación atomística de la producción y evolución de ...
Page 104: Simulación atomística de la producción y evolución de ...

VI

Page 105: Simulación atomística de la producción y evolución de ...
Page 106: Simulación atomística de la producción y evolución de ...

OKMC study of differences between MD and BCA cascades in neutronirradiated Fe simulations

S. Garcıa-Gonzaleza, A. Riverab, M.J. Aliagac, M.J. Caturlac, I. Martin-Bragadoa

aIMDEA Materials Institute, c/ Eric Kandel 2, 28906 Getafe, Madrid, SpainbInstituto de Fusion Nuclear, Universidad Politecnica de Madrid, Spain

cDept. Fısica Aplicada, Universidad de Alicante, Spain

Abstract

In this work we compare the evolution of neutron damage cascades generated by Molecular Dynamics (MD) andBinary Collision Approximation (BCA) techniques. Differences and similarities are discussed.

Keywords: OKMC, damage generation, BCA, MD

1. Introduction

One of the purposes of this work is to compare theevolution of neutron damage cascades generated byMolecular Dynamics (MD) and Binary Collision Ap-proximation (BCA) techniques.

2. Simulation methods

2.1. OKMCIn this study, radiation damage evolution has been

simulated using MMonCa [1, 2]– an Object KineticMonte Carlo code capable of modeling the atomistic-scale evolution of a system by assuming, essentially,that the transition rate between two different states (ri j)is independent of time and determined as following:

ri j = Pi j · e−Ei j/kBT , (1)

where Pi j and Ei j represent, according to the HarmonicTransition State Theory [3], the prefactor and the activa-tion barrier energy of the transition, that is computed asthe difference in formation energies between final andinitial states plus a barrier energy (typically a migrationenergy):

Ei j = E fj − E f

i + Ebi j. (2)

The OKMC algorithm is based on calculate the cu-mulative functions given by

Ri =

i∑j=1

ri j i = 1, ...,N, (3)

Email address: [email protected] (I.Martin-Bragado)

being N the total number of transitions in the system.After that, two random numbers, r and s, are computedin the interval (0,1] and the event that complies thatRi−1 < rRN ≤ Ri is performed. Lastly, the total sim-ulated time is increased by

∆t =ln(1/s)

RN(4)

and the affected transitions rates are recalculated tocompute the new cumulative functions and repeat thatprocess until the total simulated time has been reached.

2.2. Damage cascades: MD and BCA50 keV and 100 keV PKA (primary knock-on atoms)

damage cascades have been obtained as described be-low.

2.2.1. MDMD cascades have been provided by M. J. Caturla

and M. J. Aliaga and were performed using the molec-ular dynamics code MDCASK, developed at LawrenceLivermore National Laboratory [4], with the interatomicpotential developed by Ackland et al. [5] for α-Fe.This potential was modified for short range interactionsto connect to the Universal potential as described inRef. [6]. Cell dimensions for the 50 keV cascades are180 a0×180 a0×180 a0 and 250 a0×250 a0×250 a0 forthe 100 keV cascades, where a0 is the lattice parameterfor Fe (a0 = 2.8665 Å). Periodic boundary conditionswere used in all directions.

For the identification of defects, vacancies and in-terstitials are quantified and located using Wigner-Seitz

Preprint submitted to Journal of Nuclear Materials May 1, 2016

Page 107: Simulación atomística de la producción y evolución de ...

cells centered in each (perfect) lattice position so that anempty cell corresponds to a vacant and a double occu-pied cell corresponds to an interstitial defect.

2.2.2. BCABCA cascades have been simulated with

SRIM/TRIM [7, 8] considering a displacementenergy, different from the default value (20 eV), of 40eV [9]. This code only computes vacancy coordinates ingenerated cascades, thus corresponding self-interstitialdefects must be added considering that the probabilityof its final position is determined by a Gaussian distri-bution whose range and struggling depend on the recoilenergy. Hence other TRIM simulations are needed todetermine the values that characterize this distribution.Finally, each self-interstitial position is obtained con-sidering the aforesaid distance and a direction randomlygenerated from the vacancy position.

3. Model used in Iron

Depending on the material, different types of defectsare defined in MMonCa. Moreover, allowed migration,annihilation, clustering and emission of each type of de-fect and how they occur are included in the code.

In the particular case of Fe, the current model consid-ers the following defects /table 1): vacancies (V) andself-interstitial atoms (I), irregular interstitial vacancyand atom clusters and <100> and <111> interstitialloop clusters. Because neither ions from radiation norimpurities are dealt with, no other kind of defects is in-cluded. The energy values considered in the model arereferenced in table 2.

Defect Max. size Migration Transform toI & V – Yes –Iclust 9 Yes <111>clust (size>5)Vclust 500 size<5 –

<111>clust 500 Yes Iclust (size<5)<100>clust 500 No –

Table 1: Considered defects and its allowed maximum sizes, migra-tions and transformations.

The reactions between defects are defined as follows:if two same-type clusters (or point defects) react, the re-sult is a same-type cluster except when two <111> clus-ters with similar size interact, those produce a <100>cluster. If two different-type clusters react, generally,the cluster final type is the same as the largest one, butthere are some exceptions: if a vacancy cluster and anatom cluster with the same size react, they annihilate

independent of the interstitial cluster type; if an irreg-ular interstitial cluster reacts with a <111> cluster ora <100> cluster, the final cluster type is each one ofthese respectively independent of the cluster sizes; ifa <100> cluster interacts with a <111> cluster with asimilar size, a <100> cluster is produced.

Concerning interfaces, they are considered as idealsinks where every mobile defect will disappear if anyreach it.

I defects V defectsEI

m = 0.34 eV EVm = 0.67 eV

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .EI2

b = 0.80 eV EV2b = 0.30 eV

EI3b = 0.92 eV EV3

b = 0.37 eV

EI4b = 1.64 eV EV4

b = 0.62 eV. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

n > 4 : E(I|V)nb = EI|V

f +n

23 − (n − 1)

23

223 − 1

(E(I|V)2

b − EI|Vf

)Table 2: Iron model parametrization: migration and binding energyvalues (Em and Eb) and formula utilized to generate binding energiesfor clusters formed by more than four I or V.

4. Simulations set-up

In order to compare MD and and BCA cascades,OKMC simulations have been carried out introducingneutron cascades into a 300 × 300 × 300 nm3 box. Thevalue of the neutron main free path in Fe is ∼12 cm[REFERENCE], in consequence, neutron-caused dam-age takes place in the bulk. To simulate that effect, pe-riodic boundary conditions have been imposed to everyface of the simulation box and the cascades have beenrandomly introduced into this volume. Due to that mainfree path value, it is necessary to overestimate the neu-tron flux coming from a hypothetical fusion reactor. Ifnot, damage cascades would be introduced into the sim-ulation volume every too long time so nothing wouldoccurs during the main part of the simulation and itscomputational time would increase excessively. Thesesimulations have been done considering an average flu-ence rate of 2.5 neutron-cascades/(300 × 300 nm2· s).Depending on the used method to generate the damagecascades and its energies, different amount of defects isintroduced (Table 3). Distinct type of damage simula-tions have been done at temperatures in the range of 0 Kto 800 K considering a dose of 50 cascades introducedduring an average time of 20 s i. e. a dose of 5.556·1010

cm−2.The simulation characteristics of the second part of

this work partially differ from the previous ones. To

2

Page 108: Simulación atomística de la producción y evolución de ...

MD BCA50 keV 100 keV 50 keV 100 keV∼170 ∼340 ∼660 ∼1260

Table 3: Aproximated number of Frenkel pairs in damage cascadesdepending on method and PKA energy.

compare neutron with emulated proton damage, inter-faces have been added to all boundaries so defects canbe annihilated there modeling an ideal material grain. Itmakes possible to reach higher simulated volumes sim-ilar to iron grain sizes. In particular, a 500-nm-side cu-bic simulation box is considered. The fluence rate is thesame as the aforesaid but, in this case, the physical sim-ulated time is 50 s. Temperatures in the range of 0 K to800 K are also considered.

5. Results and discussion

5.1. BCA versus MD cascadesThe distribution of the defects depends on the method

used to calculated cascades. As expected, the numberof Frenkel pairs generated increases with PKA energy –besides, for each method, the damage seems to be pro-portional with energy (Table 3 ) – but, more important,the BCA method generates cascades in the same rangeof length as the MD method but with a higher number ofdefects and a different spatial distribution. In addition,self-interstitial defects in MD cascades are further apartfrom vacancies than in BCA cascades. In consequence,there is not a good agreement between interstitial distri-bution in both methods (Figure 1). The impact of thesefacts is explained as follows.

Some of the main consequences generated by thedifferences between both generating techniques are in-ferred observing the V and V clusters evolution whentemperature increases. In the case of vacancy point de-fects (figure 2 a)), despite the fact that the introduceddamage is ∼4 times higher in BCA than in MD cas-cades (table 3), a good agreement in the number of gen-erated vacancy defects exists between both methods foreach energy value. This point might be interpreted asthat vacancies and self-interstitial defects further anni-hilate in BCA cascades due to its spatial distribution.In connection with vacancy clusters (figure 2 b)), thereis a wide variation in the number of vacancies in clus-ters depending on the method but not in its average size.Furthermore, it is worth mentioning that the fraction ofthe total number of particles corresponding to vacanciesand clusters does not depend on the cascade energy, butit strongly does on the method used to compute them(figure 2 c)).

MD BCA

Figure 1: Example of 100 keV MD (left) and BCA (right) damagecascades appearance. (Colour code: purple – self-interstitial defects,grey – vacancy.)

Concerning self-interstitial point defects and clusters,the two methods reproduce the same evolution withtemperature of the first ones independent of energy al-though it is not very relevant because they vanish dueto recombination at ∼200 K and, this value is out of thetemperature range of a fusion-reactor structural mate-rial. The two methods generate a distinct evolution ofthe I clusters proportion with temperature until ∼400 Kwhen irregular interstitial clusters disappear. Further-more, this evolution also depends on the energy in caseof MD cascades but not in case of BCA ones. Despiteof that, the two methods reproduce in good agreementthe cluster average size growing with temperature, thatis independent with cascade energy (Figure 3).

Figure 3: I particles in spherical clusters proportion (left axis) andcluster average size (right axis) evolution vs. temperature.

Similar to the vacancy clusters, the fact that the BCA

3

Page 109: Simulación atomística de la producción y evolución de ...

Figure 2: V defects and clusters evolution vs. temperature for the different damage cascades : a) Number of V defects, b) Number of V clustereddefects (right axis) and V clust. average size (left axis), c) V defects and defects in V clusters proportions compared to total particles.

method generates a higher number of Frenkel pairs hasas a consequence that <111> clusters grow bigger whenthe temperature increases. In addition, as it is logical,the higher the energy of the cascade, the higher the num-ber of particles (figure 4 a)). The growing of the clustersize is described similarly by the two methods (figure 4b)). Despite there are differences in this plot, they aresmaller than the standard deviation values respect to theaverage, being open for interpretation that each methodreproduces a distinct evolution. A completely differentbehaviour is observed concerning the fraction of parti-cles in <111> clusters respect to the total number: thetwo generation techniques reproduce the same results inall temperature range except at 200 K, when this typeof clusters starts forming (figure 4 c)). The existenceof <100> clusters is also reflected in this graph as thedecrease in the total particles fraction at high tempera-tures but it is not very significant due to the low statisticassociated with this type of defects.

References

[1] I. Martin-Bragado, A. Rivera, G. Valles, J. L. Gomez-Selles, M. J.Caturla, Computer Physics Communications 184 (2013) 2703 –2710.

[2] MMonCA webpage, www.materiales.imdea.org/MMonCa, 2014.[3] G. H. Vineyard, Journal of Physics and Chemistry of Solids 3

(1957) 121 – 127.[4] T. Diaz de la Rubia, M. W. Guinan, Phys. Rev. Lett. 66 (1991)

2766–2769.[5] G. J. Ackland, M. I. Mendelev, D. J. Srolovitz, S. Han, A. V. Bara-

shev, Journal of Physics: Condensed Matter 16 (2004) S2629.[6] C. Bjorkas, K. Nordlund, Nuclear Instruments and Methods in

Physics Research Section B: Beam Interactions with Materialsand Atoms 259 (2007) 853 – 860.

[7] J. F. Ziegler, M. Ziegler, J. Biersack, Nuclear Instruments andMethods in Physics Research Section B: Beam Interactions withMaterials and Atoms 268 (2010) 1818 – 1823. 19th InternationalConference on Ion Beam Analysis.

[8] SRIM/TRIM webpage, www.srim.org/, 2014.[9] G. S. Was, Fundamentals of Radiation Materials Science: Metals

and Allowys, Springer, 2007.

4

Page 110: Simulación atomística de la producción y evolución de ...

Figure 4: <111> clusters evolution vs. temperature for the different damage cascades : a) Number of particles in <111> clusters, b) <111> clustersaverage size, c) Clusters proportion compared to total particles.

5

Page 111: Simulación atomística de la producción y evolución de ...
Page 112: Simulación atomística de la producción y evolución de ...

VII

Page 113: Simulación atomística de la producción y evolución de ...
Page 114: Simulación atomística de la producción y evolución de ...

Insights on loop nucleation and growth in α-Fe thin films under ion implantation from atomistic models

M. J. Aliaga1, I. Martin-Bragado2, I. Dopico2, M. Hernández-Mayoral3,

L. Malerba4, M. J. Caturla1* 1Dep. Física Aplicada, Universidad de Alicante, Spain 2IMDEA Materials Institute, Getafe, Madrid, Spain 3CIEMAT, Madrid, Spain 4SCK-CEN, Belgium

Abstract

The outstanding question of loop growth in a-Fe under irradiation is

addressed using object kinetic Monte Carlo with parameters from

molecular dynamics and density functional theory calculations. Two models are considered for the formation of <100> loops, both based

on recent atomistic simulations. In one model <100> loops are formed by the interaction between ½ <111> loops. In a second

model small interstitial clusters can grow as <100> or ½ <111> loops. Comparing results from the two models to experimental

measurement of loop densities, ratios and sizes produced by Fe

irradiation of Fe thin films, the validity of the models is addressed.

Introduction

An outstanding question in the field of radiation damage effects of

materials is the nucleation and growth of loops in a-Fe under irradiation. Experimentally it is well known since the 1950s that two

types of loops are formed: <100> and 1/2<111> loops [Jenkins, papers from CompMatSci]. However, the type, concentration and

ratio of one loop type vs the other type differs considerably

depending on the experimental conditions. Zhang et al. [Zhang2015] have reported the transformation of C15 clusters to both <100> and

<111>/2 loops in bcc iron, being the <100> loops more probable

(70%).

Model parametrization

We have used our database of 100 keV cascades of Fe irradiation of

Fe thin films as input for the Object Kinetic Monte Carlo code

MMonCa, developed by I. Martin-Bragado [Bragado2013]. This code

is open-source and available from [Bragado]. The simulation box we

have used is also a thin film of 50nm, reproducing a typical TEM

sample. In our code, small self-interstitial atom (SIA) clusters up to

Page 115: Simulación atomística de la producción y evolución de ...

size 4 have irregular shape and are considered mobile, with the

migration energies given in table I obtained from density functional theory calculations [Fu2004]. These self-interstitial clusters are

considered to move in three dimensions. From size 5, they can grow

according to one of this two models:

Model A: In this model all interstitial clusters above size 4 transform

in <111> loops with mobilities given also in table I and obtained from classical molecular dynamics simulations [Soneda01]. These loops

move one-dimensionally, unlike vacancies or smaller SIA clusters.

The interaction between <111> loops results in the formation of

<100> loops, <111> + <111> = <100> when the size of the two <111> loops is similar, with a maximum difference of 5%. Once the

<100> loops are formed, they can grow by addition of other <100> or <111> loops, and small interstitial clusters ( 4). The same occurs

for <111> loops. In these interactions between <100>, <111> loops and small interstitial clusters, the larger species absorbs the

smaller one. And when a <100> loop reacts with a <111> loop of

approximately the same size (5% maximum difference) the resulting

loop is also considered to have a <100> Burgers vector.

Model B: In this model <111> and <100> loops form and grow independently. SIA clusters from size 5 can either transform into

<100> loops with an initial ratio of 5%, or into <111> loops with a

ratio of 95%. These ratios are based on the work by Marinica et al [Marinica2012] according to the proportion of C15 clusters found in

cascades. Once formed, <111> loops grow by addition of other <111> loops or SIA clusters < 5, and <100> loops grow by addition of other <100> loops or SIA clusters < 5 but do not grow by addition

of ½<111> loops of any size .

In both models <111> loops can be stopped by interaction with carbon atoms, with a binding energy of 1.3 eV. These immobile C-

<111> loops can then grow by addition of <111> loops or SIA

clusters < 5. Also, <100> vacancy loops have been included in the

models. The Gilbert equation in [Gilbert2008] has been used for the binding energy of the vacancies in the loop. In this equation the

radius of the loop is calculated using the size and the density of the

loop. The density of the loop has been calculated fitting the equation to figure 4 in [Gilbert2008]. For the binding energies of Vn > 4 and

In>4 clusters, we have used the usual extrapolation law [Soneda98]:

Eb(n)=Ef+[Eb(2)-Ef][n2/3-(n-1)2/3]/(22/3-1). For the smaller species

up to 4 DFT values have been used [Fu2004]. These small vacancy

clusters are considered mobile, with a 3D mobility, while larger

vacancy clusters are immobile. Table 1 summarizes the most

important parameters of the species involved.

Page 116: Simulación atomística de la producción y evolución de ...

Table 1

Type of defect, migration and binding energies of the objects defined

in our OKMC model. Last column corresponds to the dimensionality of

migration. For the mono-defects, V and I, the formation energy is

taken from ab initio calculations [Fu204], Ef(V)=2.07 eV and Ef(I)=3.77 eV.

Defect Migration

Barrier (eV) Binding energies (eV) Migration

type

V 0,67 3D

V2 0,62 0,3 3D

V3 0,35 0,37 3D

V4 0,48 0,62 3D

Vn > 4 immobile As in ref. [Soneda98]

V 100 loops Immobile As in ref. [Gilbert08]

I 0,34 3D

I2 0,42 0,8 3D

I3 0,43 0,92 3D

I4 0,3 1,64 3D

In>4, I111 loops 0.06+0.11/n1.6 As in ref. [Soneda98] 1D

In > 4, I100 loops Immobile As in ref. [Soneda98]

C-I111 loops Immobile 1.3

In MMonCa, it is possible to specify, optionally, the capture radius for

a particular interaction. The interaction will happen when the distance between two particles belonging to each defect is smaller or equal

than the specified capture radius. If not specified, the default value of

lambda is used. In our work, and after a detailed study, we have used the capture radius for individual defects of 0.4 nm, selected to

reproduce the isolated number of defects obtained in MD.

Results

Main results:

- If we consider that all <100> vacancy loops formed in the MD simulations do not migrate or recombine with the surface, the

concentration of <100> loops is extremely high and in complete

disagreement with the experimental observations (Figure 1). Therefore, <100> vacancy loops formed close to the surface must

Page 117: Simulación atomística de la producción y evolución de ...

migrate or recombine with the surface. We have used in the OKMC

simulations a migration energy of 0.5 eV in both models.

Figure 1: Total concentration of defects as a function of dose (a)

<100> vacancy loops are immobile, compared to experiments and

(b) concentrations for different migration energy values for <100> vacancy loops.

- MD simulations of vacancy loops close to the surface are being performed to study this issue.

- Without carbon: all loops are <100> type in both models since <111> loops escape to the surface or recombine to form <100>

(Model A). In both cases, concentrations at any given dose are much higher than in the experiments (see Figure 2). For the case of model

A, the dose dependence follows the same trend as the experiment,

however, for model B the dose dependence does not look like the experiments. The concentration is very high from very low doses and

it remains almost constant.

Page 118: Simulación atomística de la producción y evolución de ...

Figure 2: Models A and B compared to experimental data.

Figure 3: Model B compared to experiments for different distribution

of <100> to <111> loops.

- Model B: changing the distribution of <100> and <111> can reduce

the total concentration of defects. If we consider that only 0.1% of all

self-interstitial clusters are of type <100> and 99.9% are of type <111> the concentrations are close to the only value that we have

experimentally for (001) orientation. Still much higher than for the

(111) orientation (see Figure 3). Things to check: orientation of the

sample. I have already tried this but I do not see any difference in

the total concentration of defects. Calculate the slope of the curves to

see if it fits the experimental power law.

Page 119: Simulación atomística de la producción y evolución de ...

- Model B: including carbon, that is, including traps for <111> self-

interstitial clusters with a trapping energy of 1eV (CHECK) results in both <100> and <111> clusters, with <100> being predominant at a

dose of ~ 4e17 ions/cm2 (Figure 4). Similar experimental trend in

terms of dose. High concentrations. Carbon concentration is 0.1%. Check other concentrations.

Figure 4: Model B including trapping sites for <111> loops.

- Model B: effect of <111> mobility. If the mobility of <111> loops is

reduced with a migration energy of 1 eV again we have two populations of loops, with more <111> loops than <100> loops at

any given dose, and higher total concentrations that for a faster

mobility of loops and similar to those concentrations obtained

experimentally for FeCr alloys (Figure 5). One problem with this

model is that we do not see saturation at high doses. Loop-loop interactions or long range interactions must be reviewed.

Page 120: Simulación atomística de la producción y evolución de ...

Figure 5: Model B including a low migration for <111> loops.

- Model A: effect of the minimum size to produce a <100> loops. Checked cases where loops are formed for any size of <111>

clusters, for clusters with size > 15 and > 30. The total concentration

is reduced slightly as the cut-off is increased. However, the trend does not change: the dose dependence is wrong, with large clusters from very low concentrations and almost no dose dependence (Figure

6).

Figure 6: Model A, effect of minimum size to produce a <100> loop.

Page 121: Simulación atomística de la producción y evolución de ...

- Model A: including Carbon as a trapping site with a binding energy

of 1 eV. Like in Model B, now we have two populations of loops,

<100> and <111> with a higher concentration of <100> than <111> at any given dose. Still wrong dose dependence (Figure 7).

Carbon concentration 0.1% like before.

Figure 7: Model A with Carbon.

- Model A: effect of <111> mobility. Now we also have two population of loops <100> and <111> with <111> at higher

concentration. And the dose dependence follows the experimental

observations (Figure 8). So, it seems the problem with Model A in terms of the dose dependence is that <111> loops migrate so fast (if

we consider the MD migration energies) that the immediately form

large <100> loops, even at very low doses, which is not observed experimentally. Can we modify this model to fit the experiments with

some reasonable assumptions?. Same problem as with Model B with

the saturation of loops.

Page 122: Simulación atomística de la producción y evolución de ...

Figure 8: Model A, effect of <111> mobility.

Conclusions

Both models reproduce the right trend for the effect of carbon (<100> and <111> loops with higher concentration of <100> loops)

and the effect of Cr (<111> and <100> loops with higher concentration of <111> loops, and higher total concentration of

defects). That is, trapping defects or reducing their mobility do not give rise to the same results, at least for a particular concentration of

trapping sites. We should check also as a function of trapping concentration (at higher concentrations there will be also more

<111> than <100> loops). These results are not new, in terms of total concentrations, but have never been shown, to the best of my

knowledge, with the distribution of <100> and <111> loops.

The main conclusion, in view of these simulations, is that Model B

(two independent population of loops) follows closer the experimental

results than Model A (<100> loop formation from <111> loop recombination), since it reproduces the right trend of loops growth

with dose. However, we should check more thoroughly all possible

parameters before making such a conclusion. Also, a model that combines both Model A and Model B should be considered for

completeness.

Bibliography

[Jenkins, papers from CompMatSci] buscar ref

[Zhang2015] Y. Zhang et al. Scripta Materialia 98 (2015) 5-8.

[Bragado] I.M. Bragado, MMonCa. <http://www.materials.imdea.org/MMonCa>.

[Bragado2013] I. Martin-Bragado, et. al. Computer Physics

Comunications (2013).

Page 123: Simulación atomística de la producción y evolución de ...

[Marinica2012] M.-C. Marinica, F. Willaime, J.-P. Crocombette, PRL

108, 025501 (2012). [Gilbert2008] M R Gilbert et al, J. Phys.: Condens. Matter 20 (2008)

345214.

[Soneda01] N. Soneda, T. Diaz de la Rubia, Phil. Mag. A 81 (2001) 331.

[Soneda98] Soneda, N. & Díaz de la Rubia, T. Phil. Mag. A 78, 995-

1019 (1998). [Fu2004] C. C. Fu, J. Dalla Torre, F. Willaime, J.-L. Bouquet, A.

Barbu, Nature Materials (2004).

Page 124: Simulación atomística de la producción y evolución de ...

Conclusiones

En esta tesis se han desarrollado modelos de simulación multiescala para

explicar los efectos de la radiación en materiales para fusión, en particular Fe y

en FeCr, con especial atención al estudio de la utilización de iones como

sustitutos del daño por radiación con neutrones. Se describen a continuación

las conclusiones más significativas de este trabajo.

El estudio estadístico detallado de las cascadas de dinámica molecular muestra

que el daño primario en láminas finas de Fe irradiado con iones de Fe es

totalmente distinto de la irradiación en el interior del material. Esto tiene

importantes implicaciones a la hora de comparar y utilizar la irradiación con

iones para “simular” la irradiación con neutrones en un reactor nuclear de

fusión, así como en el uso de otras técnicas de simulación que utilizan

cascadas de dinámica molecular como datos de entrada. El análisis del daño

primario también ha permitido mostrar que la ley de escala para la producción

de defectos bajo irradiación obtenida para W por Sand y colaboradores [5],

también ocurre en Fe. Este resultado es significativo ya que la ley de escala

obtenida podría emplearse con el fin de generar distribuciones de daño

primario más detalladas para modelos de evolución de la microestructura como

los códigos de Monte Carlo Cinético.

Además, mediante Object Kinetic Monte Carlo hemos simulado el crecimiento

de loops en Fe y en FeCr asumiendo dos modelos diferentes y hemos

mostrado que uno de ellos reproduce la dependencia con la dosis observada

experimentalmente para irradiación de láminas finas.

Por último, las simulaciones de imágenes TEM muestran, por un lado, que los

loops de vacantes obtenidos llegan a tamaños visibles al microscopio

electrónico de transmisión y, por otro lado, que la proximidad de dos loops

cercanos produce un solapamiento en la imagen que hace que se vean como

un único cluster de vacantes.

La conclusión global y más importante de esta tesis es que, para reproducir y

comprender los experimentos de irradiación con iones que se llevan a cabo en

láminas finas de material para ser analizadas por TEM, es necesaria una

Page 125: Simulación atomística de la producción y evolución de ...

descripción detallada del daño inicial producido. Esto implica la inclusión de las

superficies en las simulaciones, ya que, como se ha puesto de manifiesto en

este trabajo, el daño producido es completamente diferente al que se produce

en el interior del material, y los resultados no pueden extrapolarse de un caso

al otro.

Los resultados y conclusiones aquí obtenidos sirven de base para el desarrollo

y mejora de modelos en el programa de fusión, además de para la selección de

experimentos a realizar, con el fin último de hacer viable la fusión como energía

alternativa.

Page 126: Simulación atomística de la producción y evolución de ...

Conclusions

In this thesis we have developed multiscale simulation methods to explain the

effects of irradiation in fusion materials, in particular Fe and FeCr, focusing in

the use of ions as substitutes of neutron damage.

Detailed statistical study of the cascades produced by molecular dynamics

shows that the primary damage in thin films of Fe irradiated with Fe ions is

completely different to irradiation in bulk. This has important consequences

when comparing and using irradiation with ions to ‘simulate’ neutron irradiation

in a fusion reactor, and also in the use of other simulation techniques that use

molecular dynamics cascades as input. By analysis of the primary damage we

have as well demonstrated that the scaling law obtained for W by Sand and

coworkers [5], also occurs in Fe. This result is significant because the scaling

law obtained can be used to generate detailed primary damage distributions as

input for microstructure evolution models as Kinetic Monte Carlo.

In addition, using Object Kinetic Monte Carlo we have simulated the growth of

loops in Fe and FeCr assuming two different models, and we have

demonstrated that one of them reproduces the dose dependence observed

experimentally in thin film irradiations.

Finally, TEM image simulations show, on one hand, that vacancy loops can

grow to visible clusters under the transmission electron microscope and, on the

other hand, that the proximity of two close vacancy loops produces an overlap

in the image that causes the two clusters to be seen as only one.

The global and most important conclusion of this thesis is that, to be able to

reproduce and understand ion irradiation experiments in thin films for in-situ

TEM, it is necessary a detailed description of the primary damage produced.

This implies the inclusion of surfaces in the simulations, because, as it has been

shown in this work, the damage produced is totally different than the damage

produced in the bulk of the material, therefore, results cannot be extrapolated

from one case to the other.

The results and conclusions obtained here can be used as basis for the

development and improvement of the models in the fusion project, as well as for

Page 127: Simulación atomística de la producción y evolución de ...

the selection of new experiments, with the ultimate goal of making fusion a real

alternative energy.

Page 128: Simulación atomística de la producción y evolución de ...

Bibliografía

1. Bloom EE, Busby J.T., Duty C.E., Maziasz P.J., McGreevy T.E., Nelson B.E.,

Pint B.A., Tortorelli P.F., and Zinkle S.J., J. of Nucl. Mat. 367 (2007) 1.

2. R. Phillips, Crystals, Defects and Microstructures: modeling across scales,

Cambridge University Press 2001.

3. B. van der Schaaf, E. Diegele, R. Laesser, and A. Moeslang, Fus. Eng.

Design., 81 (2006) 893.

4. L. K. Mansur, J. of Nucl. Mat. 216 (1994) 97.

5. A. E. Sand, et al., EPL 103 (2013) 46003.

6. ITER Physics Basis Editors and ITER Physics Expert Group Chairs and Co-

Chairs and ITER Joint Central Team and Physics Integration Unit, ITER Physics

basis, Nuclear Fusion 39, 2137 (1999).

7. N. Baluc, Materials for fusion power reactors, Plasma Phys. Control. Fusion

48, B165 (2006).

8. W. J. Nuttall, Fusion as an Energy Source: Challenges and Opportunities,

Institute of Physics: Report (2008).

9. V. Barabash, G. Federici, J. Linke, and C. H. Wu, J. Nucl. Mater. 313-316, 42

(2003).

10. G. R. Odette and G.E. Lucas, Embrittlement of Nuclear Reactor Pressure

Vessels, JOM 53, 18 (2001).

11. R. S. Averback and T. D. de la Rubia, Displacement Damage in Irradiated

Metals and Semiconductors. Solid State Physics, Vol 51.

12. J. R. Beeler, Jr. in Radiation Effects Computer Experiments, Elsevier, New

York, 1983.

13. Douglas S. Billington, James H. Crawford, Radiation Damage in Solids,

Princeton University Press, 1961.

Page 129: Simulación atomística de la producción y evolución de ...

14. R. Bullough and M. H. Wood, “Theory of Microstructural Evolution”, Physics

of Radiation Effects in Crystals, edited by R.A. Johnson and A.N. Orlev.

Elsevier Science Publishers B. V., !986.

15. Radiation Damage in Metals, edited by N. L. Peterson and S. D. Harkness.

Papers presented at a seminar of the American Society for Metals, November 9

and 10, 1975.

16. W. J. Weber, L. K. Mansur, F. W. Clinard Jr., and D.M. Parkin, Journal of

Nuclear Materials 184 (1990) 97.

17. R. Stoller, Radiation Damage in Structural Materials: A Multiscale

Challenge, ICL-GT-ORNL Triad Multiscale Simulation Workshop, Oak Ridge, 3-

5 April 2005.

18. https://www.vasp.at/

19. Jose M Soler, Emilio Artacho, Julian D Gale, Alberto García, Javier

Junquera, Pablo Ordejon and Daniel Sanchez-Portal, Journal of Physics:

Condensed Matter 14 (2002) 2745–2779

20. G. R. Odette and G.E. Lucas, Radiat. Effect. Defects Solids, 144 (1998) p.

189.

21. L. Malerba, J. Nucl. Mater. 351 (2006) 28-38.

22. T. Diaz de la Rubia, K. Smalinskas, R. S. Averback, I. M. Robertson, H.

Hsei, J. Nucl. Mater. 276 (1988) 13.

23. T. Diaz de la Rubia, K. Smalinskas, R. S. Averback, I. M. Robertson, H.

Hsei, Mechanisms of Cascade Collapse (Springer, Berlin, 1988).

24. Z. Yao, M. Hernández Mayoral, M. L. Jenkins, M. A. Kirk, Phil. Mag. 88

(2008) 2851.

25. https://lasers.llnl.gov/

26. http://www-lmj.cea.fr/

27. M. S. Daw, M. I. Baskes, Phys. Rev. B 29 (1984) 6443.

Page 130: Simulación atomística de la producción y evolución de ...

28. M. W. Finnis, J.E. Sinclair, Phil. Mag. A, 50 (1984) 45.

29. S.M. Foiles, M.I. Baskes, M.S. Daw, Phys. Rev. B, 33 (1986) 7983.

30. S. Dudarev, P. Derlet, J. Phys: Condens. Matter. 17 (2005) 1-22.

31. N. W. Ashcroft and N. D. Mermin, Solid State Physics. Philadelphia:

Saunders College, 1976.

32. C. Björkas, K. Nordlund, M. J. Caturla, Influence of the picosecond defect

distribution on damage accumulation in irradiated alpha-Fe, Phys. Rev. B 85

(2012) 024105.

33. G.J. Ackland, M.I. Mendelev, D.J. Srolovitz, S. Han, A.V. Barashev,

Development of an interatomic potential for phosphorus impurities in alpha-iron,

J. Phys.: Condens. Matter 16 (2004) S2629.

34. L.A. Marqués, Tesis Doctoral, Universidad de Valladolid, 1996.

35. M.W. Finnis, J.E. Sinclair, Phil Mag. A, 50 (1984) 45.

36. M.P. Allen, D.J. Tildesley, Computer Simulation of liquids, Oxford Science

Publications (1986).

37. J. A. Brinkman, J. Appl. Phys. 23 (8) (1954) 961.

38. F.Seitz, J. S. Koehler, in: F. Seitz, D. Turnbull (Eds.), Solid State Physics,

vol. 2. Academic Press, NY, 1956, p. 305.

39. A. Seeger, in: Proc. 2nd U.N. Conf. on Peaceful Uses of Atomic Energy,

IAEA, 1958, p. 250.

40. J.B. Gibson, A.N. Goland, M. Milgram, G.H. Vineyard, Phys. Rev. 120 (4)

(1960) 1229.

41. C. Erginsoy, G.H. Vineyard, A. Englert. Phys. Rev. 133-2A (1964) A595;

42. J. R. Beeler Jr., Phys. Rev. 150 (2) (1966) 470.

43. R.S. Averback, R. Benedek, K.L. Merkle, Phys. Rev. B 18 (1978) 4156.

44. M. Ghaly, R.S. Averback, Phys. Rev. Lett. 72 (1994) 364.

Page 131: Simulación atomística de la producción y evolución de ...

45. K. Nordlund, J. Keinomen, M. Ghaly, R.S. Averback, Nature 398 (1999) 48.

46. M. Kiritani, Mater. Sci. Forum 15-18 (1987) 1023.

47. T. Muroga, N. Yoshida, N. Tsukuda, K. Kitajima, M. Eguchi, Mater. Sci.

Forum 15-18 (1987) 1092.

48. R. E. Stoller, J. Nucl. Mater. 307-311 (2002) 935.

49. R. E. Stoller, S. G. Guiriec, J. Nucl. Mater. 329 (2004) 1238.

50. K. Nordlund et al., Nucl. Instr. And Meth. In Phys. Res. B 148 (1999) 74-82.

51. Nicholas Metropolis and S. Ulam. The Monte Carlo Method. J. Am. Stat.

Assoc., vol. 44, no. 247, pages 335–341, 1949.

52. Nicholas Metropolis, Arianna W. Rosenbluth, Marshall N. Rosenbluth,

Augusta H. Teller and Edward Teller. Equation of State Calculations by Fast

Computing Machines. The Journal of Chemical Physics, vol. 21, no. 6, pages

1087–1092, 1953.

53. W M Young and E W Elcock. Monte Carlo studies of vacancy migration in

binary ordered alloys: I. Proceedings of the Physical Society, vol. 89, no. 3,

page 735, 1966.

54. A. F. Voter, Radiation Effects in Solids, NATO Publishing Unit, Dordrecht,

Netherlands, 2007, pp. 1–23

55. Crispin Gardiner. Stochastic methods: A handbook for the natural and social

sciences. Springer, 4th édition, 2009.

56. Kristen A. Fichthorn and W. H. Weinberg. Theoretical foundations of

dynamical Monte Carlo simulations. The Journal of Chemical Physics, vol. 95,

no. 2, pages 1090–1096, 1991.

57. A.B. Bortz, M.H. Kalos and J.L. Lebowitz. A new algorithm for Monte Carlo

simulation of Ising spin systems. J. Comput. Phys., vol. 17, no. 1, pages 10 –18,

1975.

Page 132: Simulación atomística de la producción y evolución de ...

58. Arthur F. Voter, Francesco Montalenti and Timothy C. Germann. Extending

the TIme Scale in Atomistic Simulation of Materials. Annu. Rev. Mater. Res.,

vol. 32, page 321, 2002.

59. George H. Vineyard. Frequency factors and isotope effects in solid state

rate processes. Journal of Physics and Chemistry of Solids, vol. 3, no. 1 – 2,

pages 121 – 127, 1957.

60. I. Martin-Bragado, A. Rivera, G. Valles, J. L. Gomez-Selles and M. J.

Caturla. MMonCa: An Object Kinetic Monte Carlo simulator for damage

irradiation evolution and defect diffusion. Comput. Phys. Commun., vol. 184, no.

12, pages 2703 – 2710, 2013.

61. Cowley, J.M., Moodie, A.M., Proc. Phys. Soc. B70 (1957) 486.

62. Williams, D.B., Carter, C.B., Transmission Electron Microscopy, Springer,

2nd edition, 2009.

63. Stadelmann, P.A., Ultramicroscopy, 21 (1987) 131.

64. Schäublin, R., Microscopy Research and Technique 69:305-316 (2006).

65. C. Björkas, (Private communication).