in copyright - non-commercial use permitted rights ... · unprecedented resolution, but also...

179
Research Collection Doctoral Thesis Mixing Phenomena Relevant to Thermal Fatigue in T-junctions Author(s): Kickhofel, John Publication Date: 2015 Permanent Link: https://doi.org/10.3929/ethz-a-010598898 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library

Upload: others

Post on 21-Apr-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

Research Collection

Doctoral Thesis

Mixing Phenomena Relevant to Thermal Fatigue in T-junctions

Author(s): Kickhofel, John

Publication Date: 2015

Permanent Link: https://doi.org/10.3929/ethz-a-010598898

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

Page 2: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

Diss. ETH No. 22904

MIXING PHENOMENA RELEVANT TO

THERMAL FATIGUE IN T-JUNCTIONS

A thesis submitted to attain the degree of

DOCTOR OF SCIENCES of ETH ZURICH

(Dr.sc. ETH Zurich)

presented by

JOHN LOUIS KICKHOFEL

M.Sc. Nuclear Engineering

ETH Zurich – EPF Lausanne

born on 20.10.1986

citizen of the United States of America

Examiners:

Prof. Dr. Horst-Michael Prasser (ETH Zurich), examiner

Prof. Dr. Eckart Laurien (University of Stuttgart), co-examiner

Prof. Dr. Yassin Hassan (Texas A&M University), co-examiner

Page 3: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new
Page 4: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

dedicated to my parents, Carolyn and Louis

"When it came night, the white waves paced to and fro in the moonlight, and the wind brought the

sound of the great sea's voice to the men on shore, and they felt that they could then be interpreters."

- Stephen Crane (1871-1900), The Open Boat

Page 5: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new
Page 6: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

I

Acknowledgements

This work and I owe a tremendous amount to the guidance and mentorship of my

Doktorvater Prof. Dr. Horst-Michael Prasser. I am forever grateful to have had the chance to

be educated and advised by such a knowledgeable, experienced and passionate

experimentalist over the course of such formative years.

While pursuing my doctoral degree I have had the opportunity to work alongside colleagues

who ensured an environment which was at once welcoming, humorous, supportive and

intellectually rewarding; one which I can only hope to experience again. Namely, those

colleagues are Robert Adams, Paolo D‘Aleo, Hassan Badreddine, Christian Bolesch, Ralph

Eismann, Nathan Lafferty and Abhishek Saxena, among others. I am especially thankful for

all of my colleagues at the Paul Scherrer Institute who have in countless discussions over the

years so adeptly fielded my questions.

I am, furthermore, very thankful for the intelligent and hardworking students who have

passed through our lab, without whom this work would not have been possible. Among them

are those who performed their Master‘s thesis work in the context of my dissertation topic:

Valentina Valori, Lukas Schmidt and Cosimo Trinca. I also had the great pleasure of

supervising Roman Ayer and Christophe Mortier while we navigated the challenges of

developing the new mesh sensor described herein.

I am indebted to the entire IKE/MPA team at the University of Stuttgart for their assistance in

ensuring our fruitful collaboration. Specifically, I would like to thank Prof. Dr. Eckhart

Laurien, Dr. Rudi Kulenovic and Karthick Selvam for their friendly, ready assistance,

patience and trust.

There are, naturally, many other individuals not mentioned here to whom I owe a great deal

with respect to this body of work. Funding has been generously provided by the Swiss

National Science Foundation project #200021_132659 and swissnuclear project #1320.

Finally, I would like recognize Sabina Näf for her steadfast love and companionship for

which I am so incredibly thankful, and also her entire family who have so seamlessly and

generously taken me in.

Zurich, 14th

September 2015

John Kickhofel

Page 7: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

II

Abstract

The focus of this thesis is single-phase mixing in the vicinity of T-junctions in nuclear power

plants (NPP) in the context of the thermal fatigue phenomenon driven by large amplitude

water temperature fluctuations. Specific zones of safety-relevant piping in NPPs are known to

be at risk of developing surface and even through-wall crack networks, often in the vicinity of

T-junctions. At certain locations, turbulent mixing and/or thermal stratification are capable of

generating damage-inducing thermal fluctuations of appropriate frequency and amplitude.

Turbulent penetration in T-junctions with weak in-flow, as in the case of a leaking valve

scenario, while representing a mixing scenario from which commercial NPPs have in the past

suffered damage, has yet to be investigated in detail from a fluid-dynamic perspective. This

work has endeavored to provide a comprehensive, foundational exploration of this mixing

scenario experimentally with a high-density of data in space and time domains and in theory

through the utilization of time-resolved CFD simulations. To this end, an adiabatic T-junction

facility for the study of turbulent penetration was constructed and utilized. A large number of

experiments at the facility with custom built instrumentation in the form of wire-mesh

sensors have illuminated important phenomena related to thermal fatigue inducing scenarios,

such as the Farley-Tihange phenomena.

Another underdeveloped, yet highly practical domain related to thermal fatigue in NPP

piping is that of cross-flow mixing in T-junctions at high ΔT, approaching reactor operating

conditions. For the purpose of studying coolant mixing at these conditions, where thermal

stratification begins to dominate, a novel mesh sensor package has been developed and

patented. The new mesh sensor has been successfully trialed in high-temperature cross-flow

mixing experiments (at up to ΔT = 232ºC at 7 MPa) in collaboration with the University of

Stuttgart. The results elucidate not only the downstream temperature profile with

unprecedented resolution, but also countercurrent stratification developing upstream at high

Richardson numbers.

The new mesh sensor technology is potentially applicable to all NPP-relevant pipe flows in

both pressurized and boiling water reactors with a foreseeable temperature resistance of up to

350ºC at 22 MPa. It represents an evolutionary step from prior state-of-the-art in a robust,

scalable package design which could incorporate additional instrumentation in the form of

thermocouples or added mesh sensor layers for recovering velocity or interfacial area

concentration.

Page 8: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

III

Zusammenfassung

Der Schwerpunkt dieser Arbeit liegt in der Mischung einphasiger Ströme in der Umgebung

von T-Verbindungen in Kernkraftwerken (KKW) vor dem Kontext der thermomechanischen

Ermüdung durch Temperaturfluktuationen mit hoher Amplitude. Dass bestimmte Zonen in

sicherheitsrelevanten Teilen des Rohrnetzes KKWs für Ermüdungsrisse anfällig sind,

insbesondere im Bereich von T-Verbindungen, ist bekannt. An bestimmten Stellen kann

turbulente Mischung und/oder thermische Schichtung zu thermischen Fluktuationen führen,

deren Höhe und Frequenz geeignet ist, Schäden zu verursachen.

Turbulentes Eindringen der Hauptströmung in schwach durchströmte Abzweigungen von T-

Verbindungen, wie es beispielsweise bei Ventillecks vorkommt, hat in der Vergangenheit in

KKWs zu Schäden geführt. Dieses Phänomen ist bisher nicht ausreichend untersucht worden.

Ziel dieser Arbeit war eine umfassende, grundlegende Untersuchung der turbulenten

Mischungsphänomene, sowohl experimentell als auch durch CFD-Simulationen. Mit diesem

Ziel wurde eine Versuchseinrichtung für die Untersuchung adiabatischer Mischungsvorgänge

aufgebaut. Eine grosse Zahl von Experimenten unter Verwendung von speziell für diesen

Zweck gebauten Gittersensoren hat einige wichtige Strömungsphänomene beleuchtet, die im

Hinblick auf thermo-mechanische Ermündung von grosser Bedeutung sind, beispielsweise

das Farley-Tihange Phänomen.

Ein anderes, wenig untersuchtes aber praktisch relevantes Gebiet ist die Mischung von

Kreuzströmen in T-Verbindungen bei grossen Temperaturdifferenzen, in denen die

Temperaturschichtung dominant ist. Zur Untersuchung der Kühlwassermischung unter diesen

Bedingungen wurde ein neuartiger Gittersensor entwickelt und patentiert. Der neue

Gittersensor wurde in Kooperation mit der Universität Stuttgart bei Temperaturdifferenzen

bis zu 232°C und Drücken von 7 MPa erfolgreich getestet. Die Resultate verdeutlichen nicht

nur das unterstromige Temperaturprofil mit bisher nicht erreichter Auflösung sondern zeigen

auch die Entwicklung einer Temperaturschichtung durch Gegenstrom im oberstromigen

Hauptast der T-Verbindung.

Die neue Gittersensortechnologie ist potenziell für alle relevanten Rohrströmungen in KKWs

mit Druck- und Siedewasserreaktoren einsetzbar, bis hin zu Temperaturen von 350°C und

Drücken von 22 MPa. Das robuste, skalierbaren Design erlaubt zudem die Integration von

Thermoelementen und zusätzlichen Gittersensorebenen. Letztere würden es ermöglichen, die

Verteilung der Strömungsgeschwindigkeit und der Grenzflächendichte über dem

Strömungsquerschnitt zu messen.

Page 9: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

IV

Sommario

La ricerca descritta in questa tesi ha come obbiettivo lo studio dei fenomeni di miscelazione

termica che avvengono in prossimità delle giunzioni a T. La presenza di questi componenti

termoidraulici è molto comune nelle tubazioni delle centrali nucleari e spesso costituisce

teatro di fenomeni di fatica termica. Tale fenomeno mette a repentaglio l‘integrità meccanica

delle strutture e, pertanto, la sicurezza dello stesso impianto. L‘origine è attribuita alla

miscelazione turbolenta tra fluidi, spesso flussi d‘acqua, con una differenza di temperatura di

svariate decine di gradi. Lungo le tubazioni metalliche, infatti, le indotte stratificazioni

termiche e/o la miscelazione turbolenta sono in grado di generare fluttuazioni termiche di

rilevante ampiezza nonché frequenze che facilitano la propagazione di cricche nei materiali

costituenti le pareti dei condotti.

Il cosiddetto fenomeno di ―penetrazione turbolenta‖ nelle giunzioni a T in condizioni di

flusso debole, come nel caso di valvole che hanno ormai perso la loro tenuta, rappresenta uno

scenario tipico che, ancora oggi, è causa di preoccupazione per le centrali nucleari

commerciali. Ne consegue la necessità di uno studio maggiormente dettagliato dal punto di

visto fluidodinamico. Pertanto, in questo lavoro si è cercato di fornire una completa

esplorazione sperimentale dei fenomeni di miscelazione che ricalcano i casi succitati. Ciò è

stato reso possibile mediante dati con alta risoluzione spaziale e temporale mentre

teoricamente il contributo è stato apportato dall‘utilizzo di simulazioni in regime transitorio.

A tal fine è stata realizzata una sezione di test che ha permesso di emulare gli scenari di

miscelamento nelle giunzioni a T in condizioni adiabatiche. Inoltre, un vasto numero di

esperimenti è stato condotto tramite strumentazione costruita ad hoc, quale il wire-mesh

sensor, che ha permesso di giungere ad importanti osservazioni relative ai fenomeni di fatica

termica, come quelli avvenuti sia nella centrale di Farley che in quella di Tihange.

Il cross-mixing è un altro di quegli effetti ancora poco studiati nell‘ambito della fatica

termica. Esso risulta più evidente quando ci si avvicina alle condizioni operative reali delle

centrali nucleari, specialmente per alte differenze di temperatura. Allo scopo, poi, di

approfondire lo studio della fluidodinamica in presenza di tali condizioni si è reso necessario

lo sviluppo di un nuovo wire-mesh sensor capace di resistere ad alte pressioni e gradienti di

temperatura (7 MPa a ΔT = 232 °C). La nuova strumentazione è stata brevettata ed utilizzata

con successo in una campagna di misure in collaborazione con l‘Università di Stoccarda. I

risultati non solo hanno messo in luce i profili di temperatura a valle della giunzione a T, con

una risoluzione spaziale e temporale senza precedenti, ma hanno anche evidenziato una

stratificazione termica che si sviluppa a monte per alti valori del numero di Richardson.

Il nuovo equipaggiamento è potenzialmente applicabile a tutte quelle condizioni che si

trovano sia nei Pressurized che nei Boiling Water Reactors con un prevedibile limite in

temperatura che si spinge fino a 350 °C e pressioni fino a 22 MPa. Questo rappresenta un

ulteriore passo evolutivo allo stato dell‘arte, in quanto il progetto e la sua robustezza sono

facilmente scalabili. É anche prevista la possibilità di incorporare strumentazione aggiuntiva

Page 10: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

V

come termocoppie, sensori capacitivi o altri wire-mesh sensor, in grado di misurare

grandezze fluidodinamiche rilevanti quali, ad esempio, i campi di velocità e la concentrazione

dell‘area di interfase.

Page 11: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

VI

Résumé

Le thème central de cette thèse concerne le mélange à une phase à proximité de jonctions en

T dans les centrales nucléaires, dans le contexte du phénomène de fatigue thermique induit

par des fluctuations à large amplitude de la température de l‘eau. Des zones spécifiques de la

tuyauterie pertinentes pour la sureté des centrales nucléaires comportent un risque de

développer des réseaux de fissures en surface, et même, à travers la paroi à proximité des

jonctions en T. À certaines locations, un mélange turbulent et/ou une stratification thermique

sont capables de générer des fluctuations thermiques de fréquence et d‘amplitude appropriées

induisant des dommages.

Une pénétration turbulente dans les jonctions en T à faible débit entrant, comme dans le cas

d‘un scénario avec une valve fuyante, alors qu‘elle représente un scénario de mélange qui a

par le passé causé des dommages dans des centrales nucléaires commerciales, doit être à

présent investie en détail du point de vue de la dynamique des fluides. Ce travail s‘efforce à

fournir une exploration compréhensive et fondatrice du scénario de mélange par l‘expérience,

avec des données à haute densité spatiale et temporelle, ainsi que théoriquement, par la

simulation numérique CFD à résolution temporel. À cette fin, un banc expérimental,

constitué d‘une jonction en T adiabatique servant à l‘étude d‘une pénétration turbulente, a été

construit et mis en service. De nombreuses expériences conduites avec le banc expérimental

et une instrumentation sur mesure sous forme de grille sensorielle ont mis en lumière des

phénomènes importants en lien avec des scénarios induisant de la fatigue thermale tel que le

phénomène Farley-Tihange.

Un autre domaine encore sous-développé, néanmoins hautement pratique, lié à la fatigue

thermique dans la tuyauterie des centrales nucléaires, est celui du mélange à courant croisé

dans les jonctions de type T à haut différentiel de température approchant les conditions

d‘opération d‘un réacteur. La nouvelle grille sensorielle fut mise à épreuve avec succès lors

d‘expériences de mélange à courant croisé à haute température (jusqu‘à 232°C et 7MPa)

menées en collaboration avec l‘université de Stuttgart. Les résultats élucident non seulement

le profil de température en aval de l‘écoulement, avec une résolution sans précédents, mais

aussi la stratification à contre-courant qui se développe en amont pour les hauts nombres de

Richardson.

La grille sensorielle de technologie nouvelle peut être potentiellement appliquée sur chaque

écoulement dans les parties de la tuyauterie pertinentes d‘une centrale nucléaire à eau

pressurisée ou à eau bouillante avec une résistance à court terme à des températures

atteignant 350°C pour 22MPa. Elle représente une étape évolutive par rapport à la dernière

technologie de pointe, de par sa robustesse et son design évolutif pouvant incorporer une

instrumentation additionnel sous forme de thermocouples ou de couches de grilles

sensorielles supplémentaires afin de reconstituer la vitesse d‘une zone de concentration

interfaciale.

Page 12: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

VII

Nomenclature

Latin Characters

f Frequency [Hz]

Fr Froude number [-]

I Raw signal [-]

K Number of ensemble averaged segments [-]

Mass flow rate [kg/s]

p Momentum [kg m/s]

Pr Prandlt number [-]

Q Volumetric flow rate typically [liters/min]

Re Reynolds number [-]

Ri Richardson number [-]

S Arc length [m]

St Strouhal number [-]

T Temperature [ºC]

(t,r) (Wire-) Mesh sensor crossing point

coordinates [-]

U Mean/bulk flow velocity [m/s]

u Local velocity [m/s]

UI Uniformity index [-]

#Db Non-dimensional branch line position [-]

#Dm Non-dimensional main/mixing pipe position [-]

Greek Symbols

α Womersley number [-]

ε Density difference [-] η Dynamic viscosity [kg/m s]

θ Mixing scalar [-]

κ Electrical conductivity [S]

ν Kinematic viscosity [m2/s]

ρ Momentum [kg m/s]

ρ Density [kg/m3]

σ Standard deviation

τ Time constant [s]

ω Angular frequency [Hz]

Vorticity thickness [m]

Page 13: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

VIII

Subscripts

b Branch line

i x-component

j y-component

k z-component

m Main or mixing pipe

r Ratio of main pipe to branch line

RMS Root-mean squared

t,r All (Wire-) Mesh Sensor crossing points

Abbreviations

BWR Boiling Water Reactor

CFD Computer Fluid Dynamics

CVCS Chemical & Volume Control System

DAQ Data Acquisition system

EC Electrical Conductivity

EdF Électricité de France

EPRI Electric Power Research Institute

FEM Finite Element Modeling

FSI Fluid-Structure Interaction

HCF High Cycle Fatigue

HCTF High Cycle Thermal Fatigue

IKE Institute of Nuclear Technology and Energy Systems, University of Stuttgart

INSS Institute of Nuclear Safety System

KEPCO Korea Electric Power Corporation

LCF Low Cycle Fatigue

LES Large Eddy Simulation

LKE Laboratory of Nuclear Energy Systems, ETH Zürich

LMFBR Liquid Metal Fast Breeder Reactor

LWR Light Water Reactor

MPA Material Testing Institute, University of Stuttgart

NEA OECD Nuclear Energy Agency

NPP Nuclear power plant

PSD Power Spectral Density

PWR Pressurized Water Reactor

RANS Reynolds-Averaged Navier-Stokes

SGS Sub-grid Scale model

SNR Signal-to-Noise Ratio

TEC Thermal Expansion Coefficient

URANS Unsteady RANS

USNRC United States Nuclear Regulatory Commission

WALE Wall-adapting Local Eddy-viscosity model

WMS Wire Mesh Sensor

Page 14: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

IX

Table of Contents Acknowledgements ........................................................................................................................ I

Abstract ......................................................................................................................................... II

Zusammenfassung....................................................................................................................... III

Sommario ..................................................................................................................................... IV

Résumé ......................................................................................................................................... VI

Nomenclature ............................................................................................................................ VII

1 Background ............................................................................................................................ 1

1.1 Five articles ...................................................................................................................... 1

1.2 Introduction ...................................................................................................................... 3

1.2.1 Fatigue....................................................................................................................... 3

1.2.2 Thermal fatigue ......................................................................................................... 3

1.2.3 Thermal stratification ................................................................................................ 6

1.2.4 In nuclear power plants ............................................................................................. 6

1.3 T-junctions ....................................................................................................................... 9

1.3.1 Cross-flow mixing .................................................................................................... 9

1.3.2 Cross-flow simulations ........................................................................................... 14

1.3.3 Turbulent penetration .............................................................................................. 16

1.3.4 Turbulent penetration CFD simulations.................................................................. 24

1.4 Parallels to other disciplines ........................................................................................... 26

1.4.1 Shear-driven cavity flow as an analogy to a dead leg ............................................. 26

1.4.2 Automotive flows analogous to cross-flow mixing ................................................ 26

1.4.3 Shear layer studies .................................................................................................. 27

1.5 Thesis structure and outline............................................................................................ 28

2 Outline of LKE T-junction facility..................................................................................... 30

2.1 Design............................................................................................................................. 30

2.1.1 Closed-loop configuration ...................................................................................... 31

2.1.2 Once-through configuration .................................................................................... 32

2.1.3 Measurement procedure .......................................................................................... 33

2.2 Instrumentation............................................................................................................... 35

2.2.1 Wire-mesh sensor.................................................................................................... 35

2.2.2 High-speed camera.................................................................................................. 41

2.3 Appendix ........................................................................................................................ 42

2.3.1 Error analysis .......................................................................................................... 42

2.3.2 Influence of the wire-mesh sensor .......................................................................... 44

Page 15: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

X

2.3.3 Gain-magnitude frequency response ....................................................................... 44

3 Outline of experiments and simulation .............................................................................. 46

3.1 Phase 1 Experiments ...................................................................................................... 46

3.1.1 Geometry................................................................................................................. 46

3.1.2 Test matrices ........................................................................................................... 46

3.2 Phase 1 Large Eddy Simulation ..................................................................................... 50

3.2.1 Simulation test case................................................................................................. 50

3.2.2 Simulation methodology ......................................................................................... 50

3.2.3 Governing equations ............................................................................................... 50

3.2.4 Computational grid ................................................................................................. 53

3.2.5 Initialization ............................................................................................................ 54

3.3 Phase 2 Experiments ...................................................................................................... 56

3.3.1 Geometry................................................................................................................. 56

3.3.2 Test matrices ........................................................................................................... 57

3.4 Regarding the orientation of presented results ............................................................... 61

4 Turbulent penetration with in-flow ................................................................................... 62

4.1 Turbulent penetration and how it differs from cross-flow mixing ................................. 62

4.2 Flow characteristics at the entrance to the branch line................................................... 67

4.3 Onset of turbulent penetration (Ur < 100) ...................................................................... 71

4.4 Flow characteristics in the branch line (Ur = 100) ......................................................... 73

4.5 Flow characteristics in the branch line (100 < Ur < 3000) ............................................. 82

4.6 Global instabilities introduced by pump oscillations ..................................................... 85

4.6.1 Influence of pulsation amplitude ............................................................................ 85

4.6.2 Influence of pulsation frequency ............................................................................ 85

4.7 Influence of T-junction edge geometry .......................................................................... 88

4.8 Influence of density stratification ................................................................................... 90

4.9 Conclusion ...................................................................................................................... 95

5 Mesh sensor package ........................................................................................................... 97

5.1 Introduction .................................................................................................................... 97

5.2 The challenge in brief ..................................................................................................... 98

5.3 Prior state of the art ...................................................................................................... 100

5.4 Material testing ............................................................................................................. 102

5.5 Prototype design and in-house testing .......................................................................... 104

5.6 Capabilities in theory ................................................................................................... 109

5.7 Conclusion .................................................................................................................... 111

Page 16: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

XI

5.8 Appendix ...................................................................................................................... 112

5.8.1 Alternate electrode materials ................................................................................ 112

6 Cross-flow mixing at high ΔT ........................................................................................... 114

6.1 Introduction .................................................................................................................. 114

6.2 Facility .......................................................................................................................... 115

6.3 Instrumentation............................................................................................................. 117

6.3.1 Mesh sensor module ............................................................................................. 117

6.4 Test Matrix ................................................................................................................... 119

6.5 Regarding the orientation of presented results ............................................................. 122

6.6 Downstream results ...................................................................................................... 123

6.7 Upstream results ........................................................................................................... 126

6.8 Combined analysis ....................................................................................................... 132

6.9 Conclusion .................................................................................................................... 136

6.10 Appendix ...................................................................................................................... 137

6.10.1 Mesh sensor calibration methodology .................................................................. 137

6.10.2 Noise filtering ....................................................................................................... 140

6.10.3 Gain-magnitude frequency response ..................................................................... 141

6.10.4 Calculation of instantaneous and maximum average thermal gradient ................ 141

6.10.5 Calculation of cold holdup area ............................................................................ 141

7 Outlook ............................................................................................................................... 143

8 Curriculum vitae................................................................................................................ 144

9 Publications ........................................................................................................................ 145

9.1 Patents .......................................................................................................................... 145

9.2 Peer-Reviewed Journal articles .................................................................................... 145

9.3 Peer-Reviewed Conference papers............................................................................... 145

10 Bibliography ....................................................................................................................... 147

Page 17: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new
Page 18: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

1

1 Background

1.1 Five articles

The field of research encompassing the problem of thermal fatigue in the vicinity

T-junctions in nuclear power plant (NPP) piping has been active for a number of decades

now. Therefore, it enjoys a tremendous wealth of published experimental and theoretical

results, many of which are referenced in this thesis and consolidated in the remainder of this

introductory chapter. Prior to delving into a detailed introduction and review of the field, it is

my desire to convey to the reader five valuable publications (excluding large technical reports

and compilations, e.g. IAEA-TECDOC 1361, EPRI MRP-25, or those of the OECD Nuclear

Energy Agency), knowledge and understanding of which can provide a valuable overview of

the history and current state of affairs in the research community.

1. J. H. Kim, R. M. Roidt and A. F. Deardorff, "Thermal stratification and reactor piping

integrity", Nuclear Engineering and Design 139, 83-95 (1993).

M. Robert, J. H. Kim, R. M. Roidt and A. F. Deardorff, "Letter to the Editor and

Authors' Response", Nuclear Engineering and Design 144, 517-419 (1993).

This paper from Westinghouse and Electric Power Research Institute (EPRI), along

with the subsequent ―Letter to the Editor and Authors' Response,‖ must appear on this list for

their historical relevance alone, scientific value notwithstanding. Kim demonstrates some of

the very earliest turbulent penetration studies in the context of LWRs, exploring such

phenomena as in-flow leakage, swirl, and density stratification. Robert, at Électricité de

France (EdF), another pioneer of the subject, takes exception with some of the findings.

2. S. Chapuliot, C. Gourdin, T. Payen, J. P. Magnaud and A. Monavon, "Hydro-thermal-

mechanical analysis of thermal fatigue in a mixing tee", Nuclear Engineering and

Design 235 (5), 575-596 (2005).

The through-wall crack downstream of a T-junction at the Civaux 1 NPP in France in

1998 is without a doubt the most oft referenced accident related to high cycle thermal fatigue

due to turbulent mixing. Chapuliot at CEA Saclay, France tackled the broad problem of

―establish[ing] natural mechanisms (turbulence, pulsing and instability) which might be the

cause of any substantial thermo-mechanical loading in the piping,‖ and did so in the context

of the pipe failure at Civaux [1].

3. H. Kamide, M. Igarashi, S. Kawashima, N. Kimura and K. Hayashi, "Study on mixing

behavior in a tee piping and numerical analyses for evaluation of thermal striping",

Nuclear Engineering and Design 239 (1), 58-67 (2009).

This paper proved to be of a foundational nature in helping to systematize cross-flow

T-junction mixing nomenclature and provided valuable data regarding all three major mixing

Page 19: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

2

types in cross-flow T-junctions where density effects do not play a major role. Select

experiments, performed in the important and scientifically prolific WALTON T-junction

facility, were simulated and mixing frequency characteristics investigated.

4. B. L. Smith, J. H. Mahaffy and K. Angele, "A CFD benchmarking exercise based on

flow mixing in a T-junction", Nuclear Engineering and Design 264, 80-88 (2013).

The ability of simulations to predict, with high accuracy, time-dependent velocity and

temperature distributions in T-junction mixing is recognized worldwide as an important

milestone. This publication is an overview of the OECD Nuclear Energy Agency (NEA) CFD

benchmark exercise based on cross-flow T-junction studies at Vattenfall Research and

Development in Sweden, written by the chairperson. It demonstrated that poor results may

befall the theoretician who does not take special care in meshing the T-junction as well as the

still questionable performance of time-resolved simulations in predicting peaks in velocity

and thermal mixing spectra.

5. J. Kickhofel, K. Selvam, R. Kulenovic, E. Laurien, and H.-M. Prasser, ―T-junction

cross-flow mixing with thermally driven density stratification,‖1 To be submitted.

In an effort to tie the scientific progress made in this thesis with the greater history of

the field, the final article in this list is a planned joint-publication reproduced in-part in

Chapter 6 of this thesis and in a highly condensed form in Kickhofel (2015) [2]. Described in

detail therein is the successful culmination of a multi-year collaboration between the

Laboratory of Nuclear Energy Systems (LKE)2 at the ETH Zurich and the Institute of Nuclear

Technology and Energy Systems (IKE)3 at the University of Stuttgart. A fruitful

measurement campaign was completed which saw valuable cross-flow T-junction mixing

data collected at reactor-like fluid conditions where density effects being to dominate and,

furthermore, the successful proof of concept testing of a novel mesh sensor package for high-

temperature high-pressure flows in pipelines, described in Chapter 5.

1 Working title.

2 Labor für Kernenergiesysteme

3 Institut für Kernenergetik und Energiesysteme

Page 20: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

3

1.2 Introduction

1.2.1 Fatigue

Fatigue is the cyclic stressing of a material which after a number of loading-unloading

cycles may lead to crack initiation, growth and eventually operating component failure.

Failures due to fatigue are stochastic in nature and the damage is cumulative; there exists

approximate fatigue limits based on experimental data. The number of stress cycles (typically

tensile) N, until failure for a given material is referred to as low cycle fatigue (LCF) if N <

104 and high cycle fatigue (HCF) if N > 10

4. High cycle fatigue is an especially challenging

material behavior to predict since in some cases it may appear that the stress cycles are of a

magnitude below the material‘s elastic limit. This case is unlike LCF in which the elastic

limit is clearly exceeded during each cycle. Once a slip band has formed and a crack is

initiated, the tensile and compressive cycling acts to open and close the crack, resulting in

striations along the length of the crack as it grows due to local plastic deformation at the

crack front. For a variety of reasons it is possible that the crack may arrest, and hence the

uncertainty in fatigue life arises. Laboratory testing of fatigue life is an arduous process.

Typically, a stress cycle is applied to a sample at frequencies greater than 10 Hz until failure.

This process is repeated for many stress magnitudes such that an S-N plot (also known as a

Wöhler curve) may be generated. Note, however, that this result is only valid for a given

material under certain ambient conditions; ambient temperature, corrosion and irradiation

may have significant influences on fatigue life.

1.2.2 Thermal fatigue

Thermal fatigue is a special case of fatigue caused not by external mechanical

loadings but by thermal loadings resulting in stress-generating thermal expansion and

contraction within the material. It may be that the material, for example steel piping, is

mechanically constrained in some way in which case the fatigue is then classified as thermo-

mechanical in nature, this is often the case in engineering applications. A typical example of

thermal fatigue in the field of power generation is cyclic stresses in pipe walls driven by

locally fluctuating internal fluid temperatures. Given appropriate fluctuation amplitude and

frequency at the wall surface, such temperature fluctuations alone may generate crack

networks bringing about component failure. Thermal fatigue is often delineated in literature

between low cycle and high cycle thermal fatigue (HCTF). The high cycle variety is a more

common threat in engineering applications with water-steam cycles since the stresses induced

by fluid temperature alone are rarely sufficient to cause a failure after fewer than 104 cycles.

Significant research has been carried out on the subject of HCTF in the past decades in the

context of NPP reactor coolant systems (RCS) where temperature fluctuations of sufficient

amplitudes and appropriate frequencies have been found to exist.

The thermal inertia of a material, a function of thermal conductivity k, heat capacity c,

and material density ρ,

Page 21: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

4

√ , (1-1)

is an important property which dictates, along with component geometry, which temperature

fluctuation frequencies will generate the largest stress amplitudes in a material sample. For

any given thermal inertia value there exists so called critical frequencies which describe the

frequency at which the maximum stresses can be expected under sinusoidal thermal loading,

for example hoop or axial stress in the case of a pipeline. On one hand, in the case of high

fluctuation frequencies, thermal inertia ensures that damaging stresses cannot yet form before

an alternate stress input, tensile or compressive, is encountered. While on the other hand,

fluctuations of sufficiently low frequency are unable to generate significant temperature

gradients in the material, as a result the elastic limit is never locally exceeded. A generalized

method for determining critical frequencies for hollow cylinders indicates that the most

damaging high cycle fluctuation frequencies for typical NPP stainless steels is between 0.1

and 10 Hz [3-5]. The theory behind the critical frequencies begins with a Henkel

transformation of the simplified sinusoidal thermal loading, from which temperature profile

in the thickness of the pipe wall may be determined, followed by the stress field [5-7].

The material science community has generated a large volume of valuable

information on material response to thermal fatigue, especially that of austenitic stainless

steels such as EN 1.4307 (X2CrNi18-9, ANSI 304L), and EN 1.4404 (X2CrNiMo17-12-2,

ANSI 316L). The broad, shallow crack networks that have been observed to form in such

steels due to thermal fatigue are referring to as ‗elephant skin.‘ The crack networks are the

result of thermal striping, also called crazing. It is believed that thermal striping is the result

of crack arrest due to high stress gradients within the material under the condition of HCTF

[8, 9]. As a result of arrested cracks, a component may remain intact as more and more cracks

are initiated. It is known that residual stresses, especially, in the vicinity of welds is often a

location in which crack networks may develop when exposed to thermal-fatigue inducing

temperature fluctuations. See photographs of thermal striping in Figure 1-1 and Figure 1-2.

Figure 1-1. (a) Thermal striping in the vicinity of a weld seam in the residual heat removal

system of a pressurized water reactor (PWR). (b) depth of crack network.[9, 10]

Page 22: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

5

Figure 1-2. A 316L stainless steel sample, 14 mm thick (a) before and (b) after 55,600

quench cycles with ΔT = 275 K, showing a crack network formed due to thermal fatigue [11].

Lifetime predictions of stainless steel piping components is dependent on knowledge

of the loading spectrum delivered by turbulent mixing or stratified flows, complicated by

additional effects from conjugate heat transfer, for example. At the moment, significant

approximations and assumptions are made, while the degree to which these simplifications

are conservative is not well understood. Modern examples of assumptions in models for

thermal fatigue damage accumulation include assuming the temperature history as a sine

wave and then proceeding to calculate stress intensity factors, interpretable using Paris‘ law

[5]. Another method, common in industry, implements Miner‘s Rule via the rainflow

counting algorithm output given the real fluctuation history as input [12]. In contrast to the

sinusoidal or semi-variable loadings assumed in most laboratory material testing and

theoretical modelling of thermal fatigue, loading in industrial processes are characterized by

variable spectra.4 Such loadings, comprised of a wide range of fluctuation frequencies, may

include phenomena such as periodic overloading and so-called sequence effects. Sequence

effects, for example the degree to which high amplitude cycles followed by low amplitude

cycles is more or less damaging than the reverse case is still under investigation in austenitic

stainless steels. Such phenomena are also explored in the laboratory. Recently, Taheri (2013)

has published a non-conservative model for damage accumulation resulting from variable

amplitude loading with emphasis on the sequence effect [13].

The transition from knowledge of the temperature history from measurements at or

near a wall to understanding of how the material will respond to such loadings is crucial, in a

safety sense. Variable loading of piping material due to fluid temperature fluctuations

remains a great challenge to the material science community in both experimental and

theoretical domains. Due to a dearth of such fluid temperature data, there has recently been

an effort made to artificially generate representative surface thermal loads encountered during

turbulent mixing in pipes [14].

The crux of the problem of thermal fatigue due to turbulent mixing in the context of

industrial piping networks is that without knowledge of the location of potentially dangerous

thermal fatigue loadings within the network over-conservatism reigns. This conservatism can

may be manifest in two distinct ways; 1. as a simple limitation on allowable operating

conditions, e.g. allowable ΔT of the mixing fluids and/or 2. by choosing materials and

4 Sometimes referred to as complex loading.

Page 23: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

6

designs which provide a too-large global safety margin for the whole network or sizeable

regions thereof. The following sections are intended to introduce some history and research

related to thermal fatigue in NPPs.

1.2.3 Thermal stratification

The thermally stratified single-phase liquid flows are known to be the culprit of

thermal fatigue by manifesting high amplitude thermal fluctuations by means of two distinct

mechanisms. The first, sometimes referred to as global stratification loading, is the result of a

typically low frequency time-dependent stratified layer height resulting in high amplitude

thermal fluctuations and a variable bending moment in the pipeline. The second mechanism

leading to potentially damaging stress reversals in a pipe are instabilities in the stratified layer

itself [15]. Modelling of these flows often takes account of the influence of the Richardson

[16].

1.2.4 In nuclear power plants

Cyclic fluid temperature fluctuations have been identified as a cause of cracking and

coolant leakage accidents in commercial nuclear power plants. High cycle thermal fatigue

due to coolant mixing has been responsible for thirteen through-wall cracks in the primary

circuit of commercial pressurized water reactor NPPs between 1982 and 1998, shown in

Figure 1-3 [17]. Historical data seems to suggest that thermal fatigue is not merely a lifetime

management issue, with many accidents occurring early in the 30, 40 or 60 years of

envisioned plant lifetime. Dahlberg (2007) in the final report of the European Commission

Network for Evaluating Structural Components thermal fatigue project concludes,

―In several instances thermal fatigue damage has developed within short times, even less than

a year. Hence the common classification of thermal fatigue as an ageing process (implying

that its probability increases progressively with time) is potentially misleading‖ [18].

Up to 2007, the OECD/NEA Piping Failure Data Exchange (OPDE) Project5 had

recorded 64 non-through-wall crack (wall thinning), 65 leak, and 3 structural failure incidents

for which thermal fatigue is believed to be the culprit [19]. Thermal striping and high cycle

thermal fatigue was the focus of the international community in the 1980‘s and 90‘s due to

cracks that had formed in NPP pressurizer surge lines due to thermal stratification. Prior to

that, in the context of liquid metal and molten salt reactors, the issue was recognized as a

potential threat as early as the mid 1970‘s due to the potentially high temperature gradients

and heat transfer coefficients of the moderators [20, 21]. Significant research on thermal

fatigue and thermal striping in fast reactor T-junctions as it related to mechanical codes has

been compiled in the IAEA TECDOC 1381, the result of a three year coordinated research

5 The project covers ―piping components of the main safety systems (e.g. ASME Code Class 1, 2 and 3)… [and

also] non-safety piping systems that, if leaking, could lead to common-cause initiating events such as internal

flooding of vital plant areas.‖

Page 24: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

7

topic and computational benchmark exercise [22]. A motivating factor in that project was the

detection of a crack near a T-junction in the secondary loop of the French LMFBR Phenix, in

1991 [23].

After a thermal fatigue failure resulting in coolant leakage at Farley 26 in December of

1987, believed to be the result of large thermal gradients caused by coolant stratification, the

NRC released the now well-known Bulletin 88-08 to U.S. LWR operators entitled, ―Thermal

Stresses in Piping Connected to Reactor Coolant Systems‖ [17, 24]. The bulletin was focused

specifically on unisolable piping connected to the RCS. Two days after the release of 88-08,

its first supplement was published with the purpose to ―1) provide preliminary information to

addressees about an event at Tihange 1 that appears to be similar to the Farley 2 event and 2)

emphasize the need for sufficient examinations of unisolable piping connected to the reactor

coolant system (RCS) to assure that there are no rejectable crack or flaw indications‖ [25].

Tihange 1 was one of three Belgian NPPs at the time. Two subsequent supplements went on

to highlight the difficulty in detecting such cracks in stainless steel piping with ultrasonic

methods and shedding light on a third, representative through-wall crack accident in the RHR

of what was described only as ―foreign reactor plant‖7 [26, 27]. These bulletins precipitated

sizeable research projects funded by the utilities not only in the United States but also abroad.

The accident at the Farley 2 plant occurred in a branch line with a leaking check valve,

however, most of the initial research focused on thermal fatigue in more simplified scenarios,

namely without a leaking valve (in-flow) situation.

Figure 1-3. Through-wall crack incidents (red bars) and leak rates (black diamonds) vs. NPP

age between 1982 and 1998 as compiled by the IAEA in TECDOC 1361 [17].

Even before the Farley 2 accident, thermal shock and striping were studied in the

context of thermal stratification, especially in the pressurizer surge line and other

stratification-prone horizontal pipes was a source of concern and research. After 88-08

experiments as a part of the HDR (Heißdampfreaktor) Safety Program studying thermally

stratified flows in a horizontal pipe at large ΔT received ―renewed international attention‖

[15, 28-31]. Since the HDR tests, a number of general experimental investigations of

6 In the RCS loop B cold-leg safety injection line (ECCS) between a check valve and the RCS.

7 Later revealed to be Genkai Unit 1 in Japan.

Page 25: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

8

thermally stratified flows related to nuclear engineering have been published including Jud

(1995) using brine and fresh water and Rezende (2009, 2012) with ΔT approaching 200ºC

[32-34]. Other researchers continue to focus on specific components such as the pressurizer

surge line in the experimental work of Qiao (2014) or CFD simulations of Boros (2008) [35,

36]. Also of interest since the late 80‘s has been countercurrent stratified flows in the cold leg

near its connection with the RPV downcomer, for example, Sibamoto (2000) with regards to

experiments at the ROSA/LSTF facility in Japan [37].

It was not until the 1998 accident at the newly constructed Civaux 1 NPP8 in France,

in which a 180 mm long through-wall crack in EN 1.4307 (ANSI 304L) stainless steel

downstream of a T-junction in the residual heat removal system resulted in a significant

coolant leakage rate of 500 l/min, that the issue of HCTF due to turbulent mixing downstream

of T-junctions started to be addressed by researchers en masse and detailed assessment

methods developed [1, 38-42]. In the case of the Civaux accident (a ‗new phenomena‘ at the

time) an incomplete understanding of locations in the pipe network where turbulent mixing

could manifest large amplitude temperature fluctuations at critical frequencies meant that the

piping network was not of a safe design for the particular operating conditions [43]. Radu

(2007) found semi-analytically that the critical frequencies for the Civaux 1 RHR piping

materials were 0.3 and 0.1 Hz for hoop and axial stresses, respectively [5]. The CEA,

Framatome and EdF intensively investigated the accident at the Civaux 1 plant with

publications related to research borne by the accident still appearing more than a decade later.

Part of their investigation was aimed at crack initiation under thermal fatigue, a major finding

was that, ―[t]hermal fatigue appears to be more damaging than uniaxial isothermal fatigue‖

[44, 45]. Impetus to better understand the phenomena, especially in the vicinity of

T-junctions has grown steadily and over the past decade and a half, especially in the field of

time-resolved CFD and finite element modeling (FEM). Despite this long history, the

prediction of thermal fatigue-prone zones remains a challenge.

8 PWR type N4, 1400 MWe

Page 26: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

9

1.3 T-junctions9

High cycle thermal fatigue due to turbulent mixing in NPPs tends to occur in the

vicinity of T-junctions. It is a reality that in the coolant piping networks of commercial PWR

and BWR reactors there exists many 90 degree intersections between pipes of varying sizes.

T-junctions prone to large fluid temperature gradients tend to be located in the high pressure

injection (HPI), feed water makeup, residual heat removal (RHR), and emergency core

cooling systems (ECCS) of LWRs. Some may exist directly at the hot or cold leg of the RCS

in PWRs where small branch lines lead to auxiliary systems. T-junction mixing is of interest

in the context of thermal fatigue because it is a location at which a hot and cold flows are

regularly meeting, in a so-called ―joining‖ or ―confluence‖ T-junction flow where two

streams meet and exit together via a single conduit. It is for this reason that T-junctions have

been the epicenter of HCTF research in the field of nuclear engineering. The most common

manifestation of such mixing in LWRs is the case of a turbulent, typically hotter flow

approaching the T-junction in the main pipe met by a second, typically cooler flow

approaching from the branch line. Given these boundary conditions, a wide variety of mixing

scenarios are possible. For the sake of clarity, only the most common scenarios will be

discussed. Figure 1-4 is meant to systematize the nomenclature in this field of research and

indeed in this thesis. T-junction mixing in the context of thermal fatigue in NPPs can be

broken down into two major groups, cross-flow mixing and turbulent penetration. Section

1.3.1 and 1.3.2 will review experiments and simulations of cross-flow mixing, respectively.

In the same vein, Section 1.3.3 and 1.3.4 will explore past and current research related to

turbulent penetration.

Figure 1-4. Categorized T-junction mixing scenarios occurring in NPPs.

1.3.1 Cross-flow mixing

In cross-flow mixing (shown in an oft published schematic in Figure 1-5) velocity ratios are

such that turbulent mixing occurs almost entirely in the mixing pipe of the T-junction. This

9 The nomenclature is not limited to ―T-junction;‖ in open literature alternatives include ―Mixing Tee,‖ ―T-type

junction‖ and ―T-pipe,‖ among others.

Page 27: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

10

type of mixing has been delineated further in published research based on the momentum

ratio of the incoming flows, see Table 1-1 [46]. The branch may be oriented horizontally or

vertically. An impinging jet results in branch line flow which impacts the far wall of the T-

junction (relative to the branch). The deflecting jet results in mixing more or less in the center

of the mixing pipe, while a wall jet does not contain enough inertia to escape far from the

wall of the mixing pipe. Figure 1-6 presents a sketch of each jet type. Each jet type may

produce characteristic frequencies downstream of the junction in both the velocity and

temperature fields as a result of large scale vortices formed due to vortex shedding of the

main flow obstructed by the branch flow, including a Karman vortex street or hairpin

vortices.

Jet type Momentum ratio

Wall jet pr > 1.35

Deflecting jet 0.35 < pr <1.35

Impinging jet pr < 0.35

Table 1-1. Traditional categorization of jet types in cross-flow T-junction mixing based on

Kamide (2009) [46].

Figure 1-5. Schematic of cross-flow mixing in a T-junction, figure adapted from Smith

(2013) [47].

(a) (b) (c)

Figure 1-6. Flow behavior in the case of an (a) impinging jet, (b) deflecting jet and (c) wall

jet, figure adapted from Sakowitz (2013) [48].

Cross-flow mixing has been intensively investigated in academia and industry,

typically at low temperatures. Naik (2010) references at least 16 previously published cross-

flow mixing studies in T-junctions since 1980, mostly in the field of nuclear engineering with

water used as the working fluid [49]. Widely cited experimental studies in T-junction cross-

flow mixing are introduced in Table 1-2.

Page 28: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

11

Page 29: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

12

Facility

Name/Location

Max

ΔT

[K]

Velocity

ratio

Ur

Main

pipe

diameter

[mm]

Branch

line

diameter

[mm]

Measurement

technique(s) Notes

Walker (2009)

[50]

- / Paul Scherrer

Institute,

Switzerland

0 0.4 –

1.67 51 51 WMS

Braillard (2007,

2009) [51, 52]

FATHER /

CEA, France 160 4 152 152

TC, fluxmeter

sensor, coefh

sensor

Civaux 1

accident

analysis

Braillard (2007,

2009) [51, 52]

FATHERINO /

CEA, France 75 3 54 54

TC, fluxmeter

sensor, coefh

sensor, IR

camera

Civaux 1

accident

analysis

Igaraishi (2002)

[53, 54]

Ogawa (2005)

[55]

Kamide (2009)

[46]

Kimura (2010)

[56]

WATLON/

Nuclear Cycle

Development

Institute, Japan

15 0.22 –

2.9 150 50 TC, PIV

Upstream

elbow, all

jet types

OECD/NEA

(2011)[57]

Smith (2011,

2013) [47, 58]

- / Vattenfall

Utveckling,

Sweden

38.6 1.0 –

2.35 190 123

TC, fluxmeter

sensor, coefh

sensor

OECD CFD

benchmark

Kuschewski

(2011, 2013)

[59, 60]

- / IKE/MPA

Universität

Stuttgart,

Germany

200 1.2 72 39 TC, PIV,

NWLED-IF

NWLED-IF

proof-of-

concept

Chen (2014)

[61, 62]

Zhang(2015)

[63]

EXTREME /

NTHU, Taiwan 90

0.037 –

0.16 208 21 TC

Present Work

[2]

- / IKE/MPA

Universität

Stuttgart,

Germany

240 1.2 72 39

TC, High

temperature

mesh sensor*

High

temperature

mesh sensor

proof-of-

concept

Table 1-2. Seminal experimental cross-flow mixing studies. *discussed in Chapter 5.

Page 30: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

13

Large international consortiums continue to build and study cross-flow T-junction

mixing from both fluid dynamic and material science perspectives to this day, with increasing

emphasis on CFD code validation and the forming of practical thermal fatigue assessment

methods based on scientific research. The FATHER project was a direct result of the 1998

Civaux accident, it was a 1-1 mockup of the Civaux case built at CEA Cadarache in 2001 and

has since been followed up by the FATHERINO project and most recently the new

MOTHER project [51, 52, 64, 65]. Results and analysis from the FATHER facility continues

to be published [66-69]. The European Commission THERFAT project for thermal fatigue

evaluation in T-junctions which was active from December 2001 to November 2004 brought

together 16 organizations in seven countries to study T-junction mixing [70]. It was

recognized that,

―In general, the common thermal fatigue issues are understood and controlled by plant

instrumentation systems. However, incidents in some plants indicate that certain piping

system Tee-connections are susceptible to turbulent temperature mixing effects that cannot be

adequately monitored by common thermocouple instrumentation‖ [70].

After THERFAT, as a part of a roadmap to a European-wide methodology on thermal

fatigue, the Network for Evaluation of Structural Components (NESC) project was initiated

with ―special attention to turbulent mixing phenomena at mixing tees in light water reactor

systems,‖ bringing together 10 organizations from five European countries, concluding in a

2007 final report [18]. The NESC project generated two primary outputs. The first is a

database of cases of thermal fatigue including 45 ‗operational components‘ and 5 from

laboratory experiments. The second was a European Procedure for Assessment of Fatigue

Damage under Turbulent Mixing which, given certain screening criteria based on material

type and ΔT between incoming flows (for example ΔT = 80°C for austenitic stainless steels),

determines fatigue usage factors and the growth of ―postulated crack-like flaw[s]‖ [18].

The OECD NEA has supported CFD benchmarking activities related the T-junction at

the Vattenfall Research and Development Laboratory in Älvkarleby, Sweden which attracted

65 participants from around the world [71]. Experimental results from the Water Experiment

on Fluid Mixing in T-pipe with Long Cycle Fluctuation (WATLON) facility in Japan, which

also incorporated elbowed main pipes, were integral in forming the JSME ―Guideline for

Evaluation of High-Cycle Thermal Fatigue of a Pipe,‖ which is at once a guideline for

designing piping systems as well as a four step assessment procedure [72].

New T-junction test facilities continue to be built and modeled. The Experiments with

T-junction on Rapid and Emergent Mixing Effects (EXTREME) facility at the National Tsing

Hua University can reach a ΔT of 70K at flow rates suitable to encounter some upstream flow

of the cold branch fluid in the main pipe [61-63]. At the University of Stuttgart in Germany,

the Institute of Nuclear Technology and Energy Systems (IKE) and the Materials Testing

Institute (MPA)10

have since 2011 operated a T-junction facility, the Fluid–Structure

Interaction Setup Stuttgart T-junction, for the study of thermal fatigue and turbulent mixing

in the case of large fluid temperature differences [59-61]. The facility incorporates reactor-

10

Materialprüfanstalt

Page 31: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

14

grade materials and can reach reactor-like conditions with ΔT approaching 240 [73]. The

EXTREME facility and the University of Stuttgart T-junction facilities have the benefit of

studying the mixing in a controlled laboratory environment with realistic working fluid

properties.

1.3.2 Cross-flow simulations

It has been recognized by the OECD NEA that thermal fatigue, the result of T-

junction mixing represents a ―nuclear reactor safety problem for which CFD analysis brings

real benefits‖ [71]. Computer fluid dynamics have historically played a significant role in

research and development related to NPPs, especially in the field of reactor safety [74-76].

With the steady decrease in cost of ever greater computing power it has recently become

practical and popular, within the last few years, to simulate single phase T-junction mixing

with time resolved large eddy simulations (LES). Studies on a wide range of parameters have

been simulated including investigations into the effect of T-junction geometry [77-79],

turbulence or subgrid-scale (SGS) model [39], structure of the mesh [80], time step [81], and

conjugate heat transfer [64, 82].

The fatigue failure at Civaux 1 was simulated using the CAST3M code developed by

the CEA [1, 83]. Two geometries were simulated, one which includes two additional

upstream bends in the branch line of the T-junction in question, see Figure 1-7. The thermal-

hydraulic and mechanical analysis indicated a concerning reality; it found that in the

simulation of the T-junction where only one upstream bend in the branch line was resolved

(Figure 1-7 left), ―no significant thermal or mechanical load was found,― at the position of the

experienced failure. When the upstream bends were included in the computation grid (Figure

1-7 right), however, the findings change dramatically: ―evidence was found of large-scale in-

stability of greater amplitude. This large-scale instability gave rise to pulses which were

characterised by the presence of significant thermal fluctuations‖ [1].

Figure 1-7. Mesh of T-junction from Civaux 1 NPP, figure adapted from Chapuliot

(2005) [1].

Page 32: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

15

Recently, the T-junction research community has been focused primarily on CFD

validation of cross flow mixing cases. The OECD NEA CFD benchmarking activity based on

the cross flow T-junction mixing experiments at Vattenfall in Sweden has been completed

and analyzed [47, 58]. In Smith (2013), the organizers of the benchmark describe the exercise

and some of the main conclusions [47]. The results were limited to the fluid dynamic aspects

of the flow and did not include conjugate heat transfer with the wall. In terms of the top

scoring simulations in the benchmark exercise, LES claimed the top 9 of 29 positions and

reproduced the average and RMS of the flow variables with good accuracy some individual

submissions to the benchmark have been independently published [84-87]. It was observed

that the mesh plays an important role; even for highly refined meshes, just having the largest

number of cells does not guarantee strong results. Furthermore, the issue of peaks in the

power spectra of the thermocouple signal and PIV data was seen as an important test of the

codes‘ ability to reproduce the flow for the purpose of a thermal fatigue analysis, in this

regard the results were underwhelming.

Figure 1-8. Side-by-side comparison of cross-flow T-junction mixing at the Vattenfall

experiment and instantaneous temperatures in the mid-plane of the T-junction from LES

simulations, figure adapted from Odemark (2009) [86].

A major international consortium with 11 partners has been working on the Modelling

T-junction Heat transfer (MOTHER) project [65]. Once again cross-flow mixing is the topic

of concern. The project is managed by Vattenfall, Sweden, but carried out at the laboratories

of CEA in France. The primary goal of the project, which encompasses both experimental

and simulation activities, is the validation of CFD for fluctuating heat transfer. The

international community approaches a validated CFD suite for the cross-flow mixing issue.

Still to be shown, however, is whether such simulations are capable of accurately reproducing

turbulent mixing in more complex geometries (such as those including upstream or

downstream bends), at large ΔT, or the wide variety of mixing scenarios possible in T-

junctions based on velocity ratio (such as turbulent penetration). Advanced simulation work

is already being carried out on reactor condition flows and geometries. These include LES

simulations, some with conjugate heat transfer, at the IKE at the University of Stuttgart with

direct comparisons to experimental data from their T-junction facility [59, 60, 88, 89].

Page 33: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

16

Figure 1-9. LES simulation results showing temperature fluctuation distribution for the case

of a wall jet in the WATLON facility geometry (left). Special attention is paid to the

formation of large hairpin vortices (right), figure adapted from Tanaka (2010) [90].

In Japan, CFD benchmarking efforts pursued in both private and public spheres utilize

experimental results from the WATLON facility (see Table 1-2). Results from that facility

were compared with in-house codes of the Japan Nuclear Cycle Development Institute,

DINUS-3 and AQUA, as well as the LES-based code MUGTHES of the JAEA [53, 55, 90-

93]. Although early simulations were not successful in reproducing accurately the

temperature fluctuations or their spectra, more recent efforts have proven to show good

agreement and have thereby advanced the understanding of the flow mechanisms. For

example results from the MUGTHES code have indicated that in the case of a wall jet,

‗mutual interaction‘ between the Karman vortex street and so-called hairpin-vortexes11

are

the dominant temperature-fluctuation-inducing phenomena; however, the results remain

sensitive to the Smagorinsky constant, with values below some threshold failing to reproduce

hairpin vortices [90, 94-96].

Qian and Kasahara, also in Japan, have for the past five years delved into the mixing

downstream of T-junctions as a means of predicting thermal loading. In a series of high

quality LES simulations of the WATLON cases, the most recent appearing in the first quarter

of 2015, the authors have investigated effects of SGS models, an important step prior to

coupling such codes with an FEM simulation. Their work includes detailed studies of the

influence of numerical schemes for convective term of the energy equation in the LES SGS

turbulence model [97-99]. Furthermore, at the Institute of Nuclear Safety System in Japan, a

number of simulation results have been published. These include DES and LES sensitivity

studies of results testing Smagorinsky constant and temperature diffusion schemes, as well as

FEM with boundary conditions based on CFD results, see Figure 1-912

[100-103]. Most

recently, simulations have been performed in search of long-period fluid temperature

fluctuations (St < 0.2) [104].

1.3.3 Turbulent penetration

11

Hibara (2004) and Sau (2008) show evidence of hairpin vortices in low Reynolds numbers experiments and

simulations, respectively. 12

Nakamura (2009) also mentions hairpin vortices as contributing to temperature fluctuations downstream of

the T-junction.

Page 34: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

17

Turbulent penetration in T-junctions is a mixing scenario which has resulted in

component failure due to thermal fatigue in commercial NPPs [17]. The phenomena,

resulting from a T-junction branch line with zero velocity (a ‗dead leg‘) or weak in-flow or

out-flow (e.g. in a leaking valve scenario), is a multifaceted one once described in a

publication by a researcher at EdF as, ―a complex and somewhat amazing hydraulic

behavior‖ [105]. Turbulent flow from the main pipe interacts with a stagnant fluid or laminar

flow in the branch line resulting in a variety of mixing frequencies and large amplitude scalar

fluctuations localized at different axial and circumferential positions depending on a number

of factors. These factors include, but are not limited to, velocity ratio between the main pipe

and branch line, main flow Reynolds number, junction geometry, and fluid densities. Figure

1-10 shows a schematic of turbulent mixing at velocity ratios sufficient to result in shallow

turbulent penetration. From another perspective turbulent penetration may be the result of a

leaking value from a branch line in, for example, a T-junction with unequal branch diameters

as in Figure 1-11. Flow structures which play a major role in cross-flow mixing, such as

Karman vortex streets or hairpin vortices are not present in the case of turbulent penetration.

Figure 1-10. Schematic of the onset of turbulent penetration in a T-junction with equal branch

diameters.

A systematic exploration of the behavior of turbulent penetration under different flow

conditions and geometries is crucial in adding to the limited experimental data and essentially

nonexistent simulation results available with regards to mixing in T-junctions at high velocity

ratios.

Page 35: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

18

Figure 1-11. Turbulent penetration in a T-junction with unequal branch diameters at high

velocity ratios, in the case of a leaking value. Adapted and colorized from IAEA TECDOC

1361 [17].

1.3.3.1 Dead leg

The most common turbulent penetration scenario in NPPs is that of mixing in a

branch line sealed at one end by a check valve. This mixing configuration is typically studied

with emphasis on thermally stratified layers in upward or downward oriented branches. Near

the T-junction the flow in the branch is characterized as being similar to cavity flow, while

deeper in the branch the flow becomes more chaotic and finally may exhibit coherent swirl.13

Thermal stratification may occur in the branch line between the deep branch region (far from

the T-junction), cooled by heat losses, and the near branch region (nearer to the T-junction)

which is cyclically supplied and re-supplied with hot fluid from the main pipe via turbulent

penetration. Cyclic of the stratified layer may result in significant thermal stresses within the

branch line. Earlier research in the context of thermal fatigue in dead legs, however, focused

on adiabatic turbulent penetration and mixing in horizontal branches. Here, the work of EdF

stands out [105-107]. Around the same time, in the U.S., Kim (1991) at EPRI also outlined

the so-called, ―Thermal stratification, cycling, and striping (TASCS)‖ issue in response to

USNRC NRC Bulletin 88-08 [108, 109].

A parameter of interest is turbulent penetration depth, specifically as a function of

main pipe flow velocity. Kim (1993) developed one of the first such correlations for turbulent

penetration depth as a function of main pipe velocity, shown in Figure 1-12 [110]. Tanimoto

(2002) observed that the penetration depth is dependent on the diameter of the branch line

rather than a ratio of the main pipe to branch line, and reconfirmed that higher velocities in

the main pipe result in deeper penetration depths [111]. The same group, a collaboration

between academic and industrial nuclear interests in Japan, went on to develop a heat balance

model and turbulent penetration depth prediction method for downward closed branches with

or without elbow for a limited range of main flow conditions [112-115]. At the time of

writing, the applicability of SST CFD simulations are being tested with regards to developing

models for the penetration depth and stratified layer location [116].

13

Nomenclature related to the cavity flow-like behavior of turbulent penetration includes ‗cellular‘ eddies and

vortices, while swirl in the branch line is sometimes referred to as a ‗spiral vortex‘ or ‗tornado eddy.‘

Confusingly, some authors refer to the entire turbulent penetration plume as cavity flow penetration.

Page 36: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

19

Figure 1-12. Turbulent penetration depth vs. main pipe velocity in a dead leg with elbow,

experiments and theory plotted by Kim (1993) [110].

Fluid disturbances in the vicinity of the stratified layer, or simply the axial translation

of the stratified layer in time due to thermal dissipation, have been shown to be induced by

residual swirl at the furthest extent of the turbulently penetrating flow. The phenomena of a

swirl extending regularly into the branch line can be seen in the schematics of Nakamura

(2013) in Figure 1-13 [117]. Swirl in the branch line was investigated, and tangential velocity

experimental measured as early as 2000 by Shiraishi [118]. A relatively large volume of

information on the subject has been published, with early work focusing on the cavity flow

behavior in the branch line [119]. In the past it hast been noted that ―mixing in the branch

pipe is mainly due to a swirling mean flow,‖ however, the same researchers at EdF also

believed that main pipe flow with a Reynolds number of less than 2.5 million was incapable

of generating swirling turbulent penetration [120]. Twenty years later, these comments have

been disproven, however, the mechanisms which generate swirl and its intermittent nature

remains unclear. Nakamura (2013) admits, ―[r]easons for this spiral flow fluctuation were not

found‖ [117].

Page 37: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

20

Figure 1-13. A schematic representation of turbulent penetration in a straight (top) or

elbowed (bottom) closed branch which results in a regularly penetrating swirl (‗spiral

vortex‘) interacting with a thermally stratified layer. Penetration depth is indicated by Ls and

Lb [117].

An invaluable body of research regarding turbulent penetration, in the case of a

downward oriented dead leg, has come from the Institute of Nuclear Safety System (INSS) in

in Japan.14

For the past decade experimental trials have, and continue to be performed at the

facilities of the INSS; their work often cites the thermal fatigue failure in 1999 of piping in

Kansai Electric Power Co. Inc. (KEPCO) Mihama Unit 2 as inspiration [121]. The inspection

of that particular accident revealed a 7 mm crack on the outer surface of the elbow in a

branch line of the letdown system in the CVCS. A series of publications and reports have

elucidated the phenomena turbulent penetration in a downward pointing branch line and

elbow leading to a horizontal closed end with a main pipe flow temperature of 65 ºC [117,

122-124]. Most recently, in late 2014, in a continuing effort to characterize the swirling

aspect of turbulent penetration, results of visual swirl observations in the branch were

reported [125]. These results indicate three velocity streamline patterns in the branch, shown

in Figure 1-14, where the first pattern tends to occur more frequently near the T-junction and

the third, farther from the junction, between the two regions small vortexes also appear

(Pattern 2). Furthermore, the group has investigated the effect of T-junction edge geometry

on penetration depth and behavior, finding, in part, that rounded edge T-junctions experience

less turbulent penetration than those with sharp edges [123].

14

The INSS was incorporated in March 1992 by KEPCO as a result of a steam generator tube rupture, also at

Mihama Unit 2, in February 1991. Some of their most recent work has yet to be published in English, references

to the Japanese publications are provided.

Page 38: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

21

Figure 1-14. Figure adapted from Miyoshi (2014) showing three typical velocity streamline

profiles in a cross-section of the branch line [125].

Yet another consistent source of research on turbulent penetration in dead legs, in

both upwards, horizontal and downwards orientations, comes from EPRI. Kim (1991, 1993),

from Westinghouse, in collaboration with EPRI published dead leg experiments measuring

velocities in a horizontally oriented branch line with and without induced swirl in the main

pipe [108, 110]. In 1994 EPRI published their final TASCS program report which began in

1989 shortly after USNRC Bulletin 88-08. Unfortunately, the issues of stratification and

swirling remained enigmatic and the findings in the TASCS program could not predict the

accident at Farley 2. As leaks due to thermal fatigue continued to appear in NPP branch lines

towards the end of the last century, EPRI formed a task group to ―pro-actively‖ address the

problem [126]. Starting in the year 2000 EPRI began publicly sharing thermal fatigue

monitoring experience in auxiliary piping of the RCS [127-129]. Keller (2002, 2003), in

collaboration with EPRI, studied swirl in turbulent penetration using compressed air, in order

to achieve Reynolds numbers on the order of 107, in a T-junction with upwards oriented dead

leg branches [130-132]. The work of Keller and others culminated in 2004 in a ―Thermal

Cycling Screening and Evaluation Model‖ for non-isolable RCS branch lines, the details of

which, along with its 2008 addendum, remain proprietary [133-135]. As a part of the EPRI

Material Reliability Program (MPR-146), an ongoing program is designed to ―assist PWR

owners manage thermal fatigue concerns in normally stagnant, non-isolable reactor coolant

system branch lines‖ [126, 136-139].

1.3.3.2 In-flow

Turbulent penetration under the condition of in-flow into the branch line can be

shown schematically as a piping and instrumentation diagram, one example is shown in

Figure 1-15. This mixing scenario relies on two conditions: firstly that the pressure gradient

ensures flow from the high-pressure source to the RCS via the branch line and, secondly, that

the fluid is injected at mass flow rates well below that of the RCS (e.g. as would be the case if

a valve leaked). Such conditions may exist in NPPs, for example in the HPI system of some

Westinghouse plants where the discharge pressure of the injection pump is approximately

200 psi higher than the operating pressure in the RCS [140]. Furthermore, recall that the

accident at Farley 2 was a case of turbulent penetration with in-flow due to a leaking check

valve. The Farley 2 accident represents what is now understood to be one of the more

complex mixing scenarios possible in the vicinity of a T-junction in an NPP. It is for this

Page 39: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

22

reason that this work focuses heavily on the issue of turbulent penetration in the case of a

weak in-flow.

Figure 1-15. Schematic of the thermal hydraulic case of turbulent penetration with in-flow in

an NPP. Inflow is ensured by the larger pressure behind the leaking valve and unisolable

piping present at the RCS [140].

Table 1-3 reviews the experimental works previously published in the field of

turbulent penetration mixing in T-junctions with in-flow. All prior works, aside from the

study of Zboray (2011) at the Paul Scherrer Institute, were performed outside academia at

Westinghouse, Mitsubishi, or EPRI [141]. Important to note, are studies of mixing in T-

junction mockups of a downcomer and intersecting cold leg in PWRs. In these studies a hot

flow travels down a vertically oriented main pipe before interacting with a horizontal branch

line carrying cold flow. Hassan (1982) presented experiments and simulations near the onset

of turbulent penetration in this small break LOCA case, providing a flow pattern map,

delineating zones of penetration and no penetration, as a function of cold and hot water

Froude number [142]. Later, Sibamoto (2000) published a 1D hydraulic model for the

stratification level in density-stratified countercurrent flow in the branch line based on data

from mixing with significant density difference at the ROSA/LSTF integral test facility and

an adiabatic experiment with brine simulating the heavier cold flow [37].

Page 40: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

23

b

m

rU

UU

2

2

4

4

bbb

mmm

r

DU

DU

m

22

2

4bbb

bmmmr

DU

DDUp

ε Rem T-Junction

Geometry

Max

ΔT

[K]

Kim (1993)

[110]

340–

2.7E3 1.0E3–8.2E3 2.2E5–1.4E7 0.25 1.4E6 ? 0

Deutsch

(1997) [107]

100–400 676–2.7E3 3.3E4–5.3E5 0 9.0E5 Sharp 0

Nakamori*

(1998) [143]

2.5E3–

7.4E4 ? ? 0.33 ? ? 306

EPRI (2002)

[127]

2.4E3–

7.1E4 5.0E4–1.5E6 2.7E7–2.4E10 0.33 3.0E7 Rounded 303

Kim (2005)

[144]

33.3–

42 148–186 4.7E3–2.9E5 0.02 1.5E5 ? 50

Zboray

(2011) [141]

33.7–

77.8 33.7–76.9 0.7E3–1.45E3 0 6.2E4 Sharp 0

Present

Work Phase

1 [145, 146]

28.6–

400 28.6–400 1.0E3–2.0E5 0

4.2E4

7.6E4

Sharp 0

Present

Work Phase

2 [147]

1E3-3E3 750–822 3.0E6–3.3E6 0-0.08

6.5E4

1.1E5

Sharp &

Rounded 0

Table 1-3. Review of experimental turbulent penetration studies in T-junctions at high Ur.

*Unknown main pipe diameter, Dm. Mass flow rate ratio, , and momentum ratio, , are

defined as in Kamide (2009) [46].

EPRI and Westinghouse, in the publications of Kim (1991, 1993) mentioned earlier,

detailed their turbulent penetration experiments in a ―standard‖ 6‖ x 3‖ T-junction with

branch oriented perpendicular to gravity [108, 110]. Leakage flow simulations at the facility

were performed with a salt solution in the branch line, representing 100ºF (37.8ºC) leakage

into a 500ºF (260ºC) circuit, and visualized with a vegetable dye. Kim reports turbulent

‗bursts‘ disrupting the stratified layer in the branch, resetting its progress towards the T-

junction while delivering the mixed fluid back to the T-junction. Furthermore, at the velocity

ratios explored the cold tongue was found typically 3 to 4 branch diameters from the T-

junction, and only occasionally reached all the way to the T-junction.

Nakamori (1995, 1998) at Mitsubishi Heavy Industries in collaboration with the

Kyūshū Electric Power Company published monitoring experience at unisolable branch lines

in Japanese PWRs, in direct response to the accidents at Farley and Tihange [143, 148]. The

project, involving five Japanese PWR utilities and Mitsubishi, was tasked with investigating

―thermal stratification and its cycling due to the leak flow through the closed stop valve.‖

Visualization tests were carried out at a low temperature facility followed by small and large

leak high temperature testing at a PWR simulation loop. A test section representing portion of

the Farley 2 ECCS, where the thermal fatigue failure occurred, was exposed to similar fluid

flow conditions present at the time of the accident. Time dependent thermal stress analysis

was performed with the MARC code; conclusions include recognition of damaging thermal

cycling when the leak rate is large, e.g. equal to or larger than the what was encountered at

Farley 2 (159 kg/hr). It is recognized by Lund (1998), however, that the results of Nakamori

do not correspond with the location of the Farley through-wall crack (approximately 5.2

diameters from the T-junction), and therefore left unanswered the question ―what caused the

fatigue failure at Farley?‖ [24]. Kim (2005) reports additional in-flow and dead leg

Page 41: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

24

experiments at reactor conditions in a mockup of Uljin Nuclear Power Plant Units 3 and 4

instrumented with thermocouples in the center of the piping. Stratification in both up-

horizontal and down-horizontal branches were explored, concluding in part that, ―[a]dditional

tests of turbulent penetration depth must be conducted systematically‖ [144].

In collaboration with EdF, EPRI instrumented a number of dead leg branch lines in

French NPPs with thermocouples to investigate temperature fluctuations due to turbulent

penetration for long durations, up to two operating cycles. More pertinent to this work,

however, were the tests performed at the Westinghouse-designed 900 MWe Blayais 1 NPP.

These experiments, also relying on thermocouple instrumentation attached to the outside of

the branch line, involved the injection of cold flow from the CVC to the RCS during

operation, essentially recreating the Farley-Tihange phenomenon [127].

The work of Zboray (2011) at the Paul Scherrer Institute, which could be considered

the precursor to the study of turbulent penetration in this thesis, performed experiments in a

T-junction with equal main pipe and branch diameters of 52 mm, at the facility of Walker

(2009) [50, 141]. The tests were performed in an adiabatic acrylic test section instrumented

with wire mesh sensors at velocity ratios up to 77, capturing the onset of turbulent penetration

at a depth of 1 diameter into the branch.

1.3.4 Turbulent penetration CFD simulations

The entirety of the aforementioned CFD simulations were performed in the context of

cross-flow mixing in T-junctions. Turbulent mixing in the case of turbulent penetration in the

branch line has not been nearly as thoroughly investigated in large part due to the

computational expense and complex meshing requirements. It is characterized by a large

recirculation zone, the result of main flow being slowed as it mixes with the incoming or

stagnant branch flow. That is to say, that a wide range of mixing frequencies and Reynolds

numbers are expected, including a transition between laminar and turbulent regimes in the

regions of interest. This is a very challenging case for single phase CFD, requiring long

simulation times.

A review of CFD results on the subject available in open literature is shown in Table

1-4. Hassan (1982) utilized the 3D steady-state code COMMIX-1A to simulate thermal

mixing and turbulent penetration at velocity ratios between 3 and 13.3 in the case of a vertical

main pipe simulating a downcomer and a branch line simulating the cold leg [142]. First

attempts at simulating turbulent penetration with a horizontal main and mixing pipe utilized

the standard k-ε turbulence model with a dead leg boundary condition. Early k-ε simulations

by Robert (1990) failed to reproduce the non-symmetric velocity field, due to swirling, in the

branch line [106]. Deutsch (1997) also showed k-ε to perform poorly in predicting

penetration depth at a velocity ratio of 400; promising results were shown rather with second

momentum turbulence closure models. Recently Ikeda (2007) performed k-ε simulations in a

downwards oriented branch pipe with buoyancy noting that the sheer force of the main flow

and friction losses at the wall in the branch pipe were important factors [149].

Ur Turbulence Flow scenario ε T-Junction ΔT Comparison

Page 42: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

25

model

Geometry [K] with

experiment

Hassan

(1982)

[142]

3-13.3 COMMIX-1A

code In-flow 0.03 Sharp 73 Yes

Robert

(1990)

[105, 106]

- k-ε Dead leg 0 ? - Yes

Deutsch

(1997)

[107]

400 k-ε and SMC Dead leg 0 Sharp - Yes

INSS

(2014)

[117]

- none Dead leg 0.02 Sharp * Yes

Ikeda

(2007)

[149]

- k-ε Dead leg 0.02 Sharp 45 Yes

Kim (2013)

[144] 16950

URANS SST-

FEM

Dead leg

In-flow 0.25 ? 245 No

Lu (2013,

2015) [78,

79, 150]

- LES Dead leg

Out-flow - Sharp 50 No

Present

work [151] 100 LES In-flow 0 Sharp 0 Yes

Table 1-4. Review of published CFD simulations of turbulent penetration in T-junctions.

*10ºC and 20ºC branch-end wall temperatures.

The INSS is active not only in dead leg experiments but also has published a number

of CFD studies [124, 152, 153]. Their most recent publication describes DNS simulations of

straight and elbowed branch lines with heat loss and no leakage. Some simulation work on

the subject of in-leakage to a T-junction was performed by EdF in the 1990‘s and took

advantage of k-ε and second momentum closure turbulence models to simulate turbulent

penetration in a dead leg at a velocity ratios up to 400. Since then, publications related to the

issue have been largely absent. Recently, researchers at the Korea Institute of Nuclear Safety

performed coupled URANS-FEM simulations of turbulent penetration into a branch line with

elbows, concluding that, ―in-leakage has a high possibility of causing considerable structural

problems in RCS piping‖ [154]. The time step of 0.5 s in those simulations, however,

precluded an assessment of high frequency fluctuations seen in turbulent mixing near the

junction.

As for LES simulations, Lu (2013) has recently published results of out-leakage

simulations in a T-junction branch line, focusing, however, on stratification rather than

turbulent mixing [78, 79]. Utilities continue to focus on the issue of ―Thermal Fatigue in

Normally Stagnant Non-Isolable Reactor Coolant System Branch Lines,‖ as EPRI designates

it. They provide an entire ‗Product Set‘ on the topic including thermal fatigue evaluations and

thermal-cycling models [155]. There also exist models for predicting turbulent penetration in

dead legs based on heat transfer considerations.

Page 43: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

26

1.4 Parallels to other disciplines

It is an easy mistake to limit the scope of one‘s research, knowledge, and overall

awareness to one‘s particular field, in this case, nuclear engineering.15

Therefore, as means of

combating this tendency, I wish to provide some information and references regarding

turbulent penetration mixing in other fields beyond nuclear engineering which typically go

entirely unnoticed (and uncited) within the research sphere directed at nuclear safety issues

related to HCTF in T-junctions. There are potentially significant amounts of knowledge

which may be garnered (and redundancy avoided) by looking beyond nuclear.

1.4.1 Shear-driven cavity flow as an analogy to a dead leg

Turbulent penetration mixing in T-junctions is highly similar to the long-active

research topic in fluid dynamics known as shear-driven cavity flow.16

The study of cavity

flows originates in the study of shear-driven compressible flows, specifically in the context of

flow acoustics. These flows occur typically over airframes with indentations which may take

on geometries such as shallow indentations or deep cavities resembling a T-junction branch

line. Non-compressible cavity flows have also been investigated. Typically, this field is based

in simulations, especially in 2D, as a means of model validation. The study of 3D deep cavity

flow (where the cavity is of depth greater than its characteristic length) begins to mimic very

closely the dead leg branch line situation in an NPP (when an incompressible fluid is

considered) albeit without significant thermal or buoyancy effects. These flows are typically

modelled, using a lattice Boltzmann method, as laminar or transition flows where the interest

lies in precisely simulating steady vortex structures in the branch. This field is potentially

valuable to nuclear engineers studying T-junction mixing in terms of both high quality CFD

and experimental results.

1.4.2 Automotive flows analogous to cross-flow mixing

One such non-nuclear field where turbulent mixing in T-junctions is of interest is in

the automotive industry, again in compressible flows. For example, recent research has

focused on cross-flow mixing in a T-junction with steady and pulsed inlet conditions as may

be found when applying exhaust gas recirculation in internal combustion engines. The work

of Sakowitz (2013, 2014) utilizes steady state RANS as well as time resolved LES

simulations. Flow patterns include all three major types of cross-flow mixing including wall

jet, deflecting jet and impinging jet in a T-junction with a branch diameter ratio of 0.5 [48,

156].

15

Fortunately, nuclear engineering is in itself a highly multidisciplinary endeavor. 16

Not to be confused with lid-driven cavity flows.

Page 44: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

27

1.4.3 Shear layer studies

Finally, since the departure of the main flow into the branch must be recognized as an

important, initiating event in turbulent penetration mixing is it evident that fundamental

studies of shear flows are of value. In their seminal paper on the topic, Brown and Roshko

(1974) helped to elucidate the turbulent structures appearing in turbulent free shear layers

with high-speed footage [157]. More recently, work at the Paul Scherrer Institute has focused

on fundamental studies of shear layer mixing in flows with small and large differences in

density in stable and unstable configurations [158-160]. The facility hosting those

experiments, GEMIX, has also recently hosted measurements with a newly developed wall

heat flux sensor [161, 162].

Page 45: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

28

1.5 Thesis structure and outline

The work covers two particular thermal-fatigue relevant mixing cases in T-junctions.

Chapters 2, 3, and 4 address the first of these scenarios: turbulent penetration mixing with in-

flow from the branch line. The second chapter introduces the LKE T-junction facility and

associated measurement instrumentation including details of the measurement procedure at

the facility, wire-mesh sensor construction and calibration methodology. Test matrices, along

with experimental and CFD simulation methodology are covered in Chapter 3. Chapter 4 is

an extensive analysis of the results as a whole, meant to generate a comprehensive

understanding of the turbulent penetration phenomena in the case of in-flow. Within these

chapters is described,

The design, construction, commissioning and utilization of an adiabatic single phase

T-junction facility for the study of turbulent penetration in the branch line in the

context of thermal fatigue.

An investigation of turbulent penetration in T-junctions with in-flow against the

penetrating main flow direction in the branch line by means of wire mesh sensors and

time-resolved CFD simulations.

Test matrices including T-junction geometry, density stratification, pulsatile flow and

velocity ratios up to 3000.

A detailed description of turbulent penetration behavior in the Farley-Tihange

phenomena, beyond what is available with regards to experiments or simulations in

open literature, including Kim (1993), Kim (2005), and Kim (2013) [110, 144, 154].

Chapters 5 and 6 are dedicated to the second mixing scenario which is cross-flow T-

junction mixing at high temperatures and pressure. Chapter 5 introduces the new, patented

high-temperature high-pressure mesh sensor package design, built for the purpose of cross-

flow mixing studies at the University of Stuttgart T-junction facility. This is followed by

results from mesh sensor prototype testing in cross-flow T-junction mixing experiments at

that facility, in Chapter 6. Major topics to be addressed include,

Design, construction, assembly, and proof of concept testing of a novel mesh sensor

package capable of withstanding high-temperature steam-water environments (up to

350ºC) at high pressures (up to 22 MPa).

Investigation of cross-flow mixing in T-junctions at reactor-like temperatures (256ºC)

and pressures (7 MPa) at the University of Stuttgart T-junction facility by means of

novel mesh sensor instrumentation.

Page 46: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

29

Detailed description of flow patterns and mixing frequencies both downstream and

upstream of the T-junction at high Richardson numbers with strong thermally-driven

density stratification.

Finally, in Chapter 7 a brief outlook is provided.

Page 47: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

30

2 Outline of LKE T-junction facility

2.1 Design

In the context of this work, a new T-junction facility (Figure 2-1) has been design and

constructed for the express purpose of studying turbulent penetration with a high density of

instrumentation, primarily in the form of wire mesh sensors (WMS). Prior to this work,

preliminary turbulent penetration experiments by Zboray (2011) were carried out at the T-

junction facility of Walker (2009) which was optimized for cross-flow mixing tests at

velocity ratios close to unity. The boundary conditions of these tests, in terms of turbulent

penetration, were therefore limited to the onset of turbulent penetration [50, 141]. The new

test facility described in this chapter was designed and instrumented for the case of weak

branch in-flow with a high degree of geometrical flexibility, allowing for various branch

diameters, sensor positioning, and T-junction geometry.

The facility operates under adiabatic conditions at atmospheric pressure. It is

comprised of an acrylic glass 90 degree T-junction and inlet run sections fed by flexible PVC

hose and PVC-U piping. Each branch contains a honeycomb section at the start of the acrylic

inlet sections for flow conditioning. Two high-density polyethylene feed tanks of 750 l

volume each separately supply the junction with tap water (approx. 150 µS/cm) in the main

pipe and deionized water (approx. 3 µS/cm) in the branch line such that electoconductive

measurement techniques, specifically WMSs (see Section 1.1), can distinguish mixing at the

T-junction. A waste tank of 1000 l volume acts as a sump for the working fluid.

Page 48: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

31

Figure 2-1 Schematic of the T-junction facility including main and branch supply tanks,

waste tank, numbered solenoid valves, pumps, and flow meters.

The T-junction is supplied by two frequency-controlled centrifugal multistage pumps.

The main pipe pump (Grundfos CME) is capable of delivering in upwards of 200 liters per

minute (12 m3/h) while the branch pump (Grundfos CRE) is tasked with delivering less than

a single liter per minute (< 0.06 m3/h) against a relatively large hydraulic head. Manual

regulating valves downstream of the pump (PD1 and PD2) help to further reduce the branch

line flow rates, measured in Phase 1 experiments by a vortex flow meter and by an oval gear

positive displacement flow meter in Phase 2. An electromagnetic flow meter in the main

circuit delivers high accuracy flow velocity readings with a frequency of 8 Hz. A National

Instruments CompactRIO chassis interfaces with the LabView software on a PC. Solenoid

valve control is handled within a LabView Virtual Instrument and two tuned PID controllers

help to maintain the appropriate flow rates in the pumps. Information regarding tap water

conductivity is also fed into the program. During filling of the facility, air bubbles are flushed

(via Bubble0) while tap and deionized water are independently recycled through their

respective supply tanks.

2.1.1 Closed-loop configuration

The facility may operate in two configurations; the first, ―closed-loop‖ is used for

tests without density stratification The closed-loop configuration (Figure 2-2) allows for long

duration measurements to be performed which is beneficial from many points of view,

especially in reducing noise in mixing spectra while maintaining low frequency components

when ensemble averaging or in the search for long-period fluctuations. When mixing tap

water and deionized water at high velocity ratios the addition of small amounts of deionized

water, pre-mixed during the return trip from the T-junction to the tap water supply tank,

induces a change of 0.25% or less in the electrical conductivity over the course of a

measurement. To compensate for this already small disturbance the calibration procedure

takes advantage of the linear response of the WMS. The tap water calibration used during

data processing is the postulated sensor signal at the average tap water supply tank

conductivity during the measurement, calculated from a linear interpolation between two tap

water calibrations at known conductivities before and after the measurement series.

Page 49: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

32

Figure 2-2. Closed-loop configuration at the T-junction facility. Mixed fluid exits the T-

junction and returns to the main pipe supply tank via the RV4 valve.

2.1.2 Once-through configuration

Once-through tests (Figure 2-3) are performed in the case of non-equal inlet flow

densities or if the velocity ratio is relatively low, such that considerable contamination would

occur by delivering the mixed fluid back into the main tank. Increased density in the branch

flow is achieved by adding sugar to the deionized water tank. The tank is mixed until

homogeneous and then a sample is taken and the density is measured with a hydrometer.

Measurement durations of 60 seconds were achieved when running the facility in the once-

through configuration; water in the mixing pipe is sent directly to a waste tank which is

drained after the measurement.

Page 50: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

33

Figure 2-3. Once-through configuration at the T-junction facility. Mixed fluid exits the T-

junction into the waste tank via the Dump1 valve, to be drained.

2.1.3 Measurement procedure

Prior to the start of an experiment the facility is filled by the main pump which

recycles tap water from the Main Tank, via the T-junction. The branch pump recycles

deionized water from the Branch Tank. The Bubble0 valve is opened temporarily in order to

flush any air present in the circuit out with tap water via the branch line to the Waste Tank.

Page 51: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

34

Figure 2-4. Flow in the T-junction during calibration recordings and initial condition prior to

a measurement. Gravity points into the page.

The experimental procedure is essentially consistent for all measurements at the

facility. The experiment starts when the SideFlow2 valve is opened (RecycleSide4 is then

closed), bringing deionized water into the branch line.17

The branch line flow rate is kept high

in order to completely flush tap water from the auxiliary piping and the branch itself. The

main flow is kept low during this flushing procedure. At this point the PID controller is

activated and seeks the set point flow rate in the branch. As the branch flow rate slows to the

set point, the main flow PID controller is also activated and the velocity in the main pipe

increases to its respective set point. As the flow rates approach the values for the given

experiment, turbulent penetration begins. The measurement begins at least 10 seconds after

both the branch and main flow velocity readings have stabilized. This procedure is shown

schematically at the T-junction in Figure 2-4.

17

It is at this point that in a once-through experiment the Dump1 valve would be opened and RV4 closed.

Page 52: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

35

2.2 Instrumentation

2.2.1 Wire-mesh sensor

Single phase mixing in the vicinity of the T-junction was measured by an electrode-

mesh device known as a wire-mesh sensor (WMS) by recording the electrical conductivity of

the flow at many points in space between transmitter and receiver electrodes, in this case thin,

tensioned stainless steel wires. The WMS reads analog current detected in a multiplexed-

manner where each transmitter electrode is sequentially delivered a square-wave voltage

pulse (± 3V) during which all receiver electrodes are sampled in parallel. Cross-talk between

neighboring receivers or transmitters is suppressed by means of low-impedance drivers and

inputs for the transmitters and receivers, respectively. The measurement technology is

typically applied in, but not limited to, closed conduits such as flow channels or pipelines

where the conductivity is measured throughout the cross section of the duct with a typical

resolution, determined by the pitch of the electrodes, of 1 to 5 mm. The voltage pulses

through the transmitter electrodes are multiplexed at up to 10 kHz. In this manner the

conductivity of the fluid can be accurately measured by a WMS with high temporal and

spatial resolution. A schematic of the operating principle of the WMS can been see in Figure

2-5. More information can be found in Prasser (1998) [163]. The current signals are

converted to voltage and recorded by 12-bit analog to digital converters in the WMS data

acquisition (DAQ) electronics for the entire measurement duration prior to being transferred

via USB to a desktop computer.

The WMS accurately measures the conductivity of the flow only when calibrated.

Calibration of the WMS itself is not performed as a part of the experiments in this work. The

calibration of the WMS itself could be achieved with a solution of known resistivity in a

procedure analogous to the calibration of traditional conductivity cells. In practice, when we

are interested in conductivity only as a measure of a dimensionless mixing scalar used to

visualize single phase mixing, absolute conductivity values are not of interest. Wire-mesh

sensors have been effectively implemented in measuring single phase mixing of fluids with

different temperature by means of detecting the conductivity change of water with

temperature. Often, however, a salt tracer is introduced as a means of rendering the mixing

visible to the sensor. By recording uniform measurements of each working fluid, in this case

tap and deionized water, a baseline is set from which to measure scalar concentrations.

Signals are converted to a normalized scalar meant in this case to be analogous to fluid

temperature. This analogy is valid in the case of turbulent flows where turbulent diffusivity

dominates and is furthermore believed to introduce only small errors in the case of slower,

laminar flows sometimes discussed in this work since the thermal diffusivity of water and

mass diffusivity of NaCl solutions have similar magnitudes [164]. Salt tracer experiments

represent true adiabaticity since the walls are unable to absorb or dampen concentration

changes in the fluid. It has been shown to be the case, however, that the boundary layer acts

to dampen fluctuations in salt tracer concentration near the wall [160].

Page 53: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

36

Figure 2-5. Schematic of the operating principle of the wire mesh sensor from Prasser (1998)

[163].

2.2.1.1 Construction

Sensors at the T-junction facility were designed in-house as a part of the thesis work

described herein. The components of a WMS include identical transmitting and receiving FR-

4 PCB boards on which 16 or 32 0.05 mm diameter EN 1.4404 (AISI 316L) stainless steel

wires are tensioned and soldered to solder pads on the boards parallel to one another with a

constant pitch of 1.68 mm or 1.56 mm, respectively. The boards are placed perpendicular to

one another forming a Cartesian wire-mesh separated by a thin 1.5 mm sealing plate made of

soft-PVC of hardness 77 Shore A which also acts to seal the conduit at this location. Acrylic

glass flanges with O-rings fix the WMS assembly in position in the branch line or mixing

pipe. Shielded cables with D-SUB 25 connectors connect the DAQ unit to the receiver and

transmitter PCB boards, respectively. Figure 2-6 shows the arrangement of the PCB boards

with a 26 mm diameter opening for the flow to the past through the wire-mesh as well as M6

bolt and centering pin (3 mm, h8) holes for connection and alignment with the flanged

piping. Embedded copper traces carry the signal from the solder pads to the D-SUB 25

connector. The PCB boards with 16-by-16 electrodes measures 160 mm x 80 mm x 3.2 mm

thick, those with 32 electrodes are slightly larger, 160 mm x 100 mm x 3.2 mm thick. In

Figure 2-7 a photo of a completed transmitter board with 32 electrodes is shown.

Page 54: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

37

Figure 2-6. Schematic of transmitter and receiver PCB boards positioned perpendicular to

each other to form a 16-by-16 electrode (left) WMS and 32-by-32 electrode (right) WMS.

Flow passes through the center opening.

Figure 2-7. Photograph of a single PCB board with 32 electrodes, 64 soldering pads, two D-

SUB 25 connectors and sealing ring.

Of the crossing points, i.e. measurement points, 232 lie within the 26 mm diameter

flow area in the case of the 16-by-16 sensor, and 856 of the 32-by-32 electrode sensor lie

within the 50 mm diameter flow area. The orientation of the mesh and it‘s crossing-point

coordinate system (t,r) relative to gravity (pointing in the –z direction) is shown in Figure

2-8.

Page 55: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

38

Figure 2-8. Convention for identifying individual crossing points (t,r). A 16-by-16 grid over a

26 mm flow area and a 32-by-32 grid is drawn over a 50 mm diameter flow area according to

the viewing orientation. Gravity is pointing downwards.

2.2.1.2 Installation

Installation of the sensors in the branch line or mixing pipe of the T-junction is

achieved in one of two ways depending on the sensor position. If the sensor is installed at the

T-junction itself, four screws are fed through the flanged end of the branch line or mixing

pipe along with two centering pins and fixed directly to the T-junction which is fitted with

Heli-Coil® inserts to avoid fracturing the component upon screw-tightening. For positioning

of the sensor within the branch line, flanged acrylic glass spacer pieces have been

manufactured to be connected directly to the junction. Sealing between the flanges and the

PCB boards is ensured by standard rubber O-rings. Figure 2-9 is a technical drawing of the

flange design. Furthermore, a photograph of a WMS installed in the branch line is shown in

Figure 2-10.

Figure 2-9. Flange design with four through-holes for tightening, two centering pin holes, and

O-ring housing. Values in mm.

Page 56: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

39

Figure 2-10. A single 16-by-16 electrode WMS installed in the 26 mm diameter branch line

of the T-junction. Visible on the left is the 50 mm diameter main pipe, on the top right, the 50

mm diameter mixing pipe, and, on the bottom right, the 26 mm branch line.

2.2.1.3 Calibration methodology

The calibration procedure normalizes the signal I(t), the WMS can expect to measure

at every crossing point based on the working fluids. By means of calibration measurements of

each working fluid, the natural, albeit small, discrepancies in the cell constants between the

electrode wires at each crossing point in the sensor is accounted for. The calibration

procedure is carried out prior to ever measurement series, and every time the supply water is

exchanged.

The facility is circulated with tap water at which time a ―high‖ calibration (Ihigh)

signal is recorded for 2 seconds for all crossing points at the planned measurement frequency

(1600 Hz or 2500 Hz). The branch line valve (SideFlow2) is then opened allowing deionized

water to fill the branch line completely at which time a ―low‖ calibration (Ilow) is recorded,

also for 2 seconds. For once-through tests the low calibration measurement would take place

while the circuit was emptying into the waste tank via the T-junction. Also in the case of

once-through measurements, the calibrations are valid for the entire measurement duration

and therefore the mixing scalar is calculated at a given crossing point simply as,

(2-1)

where I(t) is the measured signal at a given crossing point and Ilow and Ihigh are the average of

the calibration values at that crossing point. Figure 2-11 shows the average signal strength at

each crossing point in a 16 transmitter, 16 receiver electrode sensor

In closed-loop operation some deionized water from the branch line is recycled into

the main tank over the course of a measurement, disturbing the conductivity of the main pipe

flow, and therefore also deviating it from the initial tap water calibration signal. A

conductivity meter showed that the change in conductivity between the initially calibrated

value and the tank conductivity at the end of a series of measurements was always less than

Page 57: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

40

0.25%. For greater accuracy, two high calibration measurements were recorded for each

closed-loop measurement series at different conductivities, I‟high and I‟‟high for the purpose of

generating a linear interpolation between the crossing point signals during each calibration.

Knowing that the signal output of the WMS is linear with electrical conductivity, a tap water

calibration function, Ihigh = f(I‟high, I‟‟high, κmain), for each crossing point in the sensor was

computed for an arbitrary main tank conductivity, κmain. Then, given the average main tank

conductivity during the course of a closed-loop measurement, (based on a linear

interpolation of pre-and post-measurement readings), the normalized mixing scalar is,

(2-2)

Data acquisition software provided is by teletronic Rossendorf GmbH [165]. The

software controls the amplifier ‗pre‘ gain and ‗main‘ gain. It is necessary that these values are

set such that the measured signal is never saturated or very near saturation (i.e. always < 85%

of the maximum possible signal strength). Should the signal strength approach saturation not

only would there be a loss of information but there may also be unwanted artifacts introduced

in the form of cross-talk where receivers may detect current travelling from non-neighboring

transmitters.

Figure 2-11. Average normalized crossing point values from a two second tap water, Ihigh,

calibration (left) and a two second deionized water, Ilow, calibration (right) measurement.

The signal-to-noise ratio (SNR) is calculated as the reciprocal of the coefficient of

variation, that is the average of the raw signal (or mixing scalar) divided by the standard

deviation of that same signal,

⁄ (2-3)

In Figure 2-12 the signal strength has been calculated according to Eq. 2-3 for the

same calibration measurements shown in Figure 2-11. The SNR at high signal strengths is

more than two orders of magnitude larger than that at the lowest signal strengths. This owes

mostly to the reduction in average signal approaching the noise rather than a large change in

the noise itself. Note that the lower SNR of the low calibration has little effect on the

uncertainty of the measured dimensionless scalar due to the error propagation though Eq. 2-1

or, respectively, 2-2.

Page 58: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

41

Figure 2-12. Signal to noise ratio for (left) a two second tap water, Ihigh, calibration and (right)

a two second deionized water, Ilow, calibration measurements.

2.2.2 High-speed camera

Some additional understanding has been gleaned from videos recorded by a high-

speed camera installed to observe turbulent penetration in the branch line. To visualize the

mixing two blue LED spotlights induced fluorescence in the fluorescein-doped main flow.

The camera, a HighSpeedStar 3, produced by LaVision, recorded the mixing at a frame rate

of 60 Hz at 1024 × 1024 pixels. The spotlights produce a light intensity of 600 μW/cm2 at a

wavelength of 470 nm, near the 480 nm peak in the excitation wavelength of fluorescein

sodium salt (emission wavelength 525 nm). Approximately 0.85 g of fluorescein was mixed

with 350 liters of tap water in the main tank. Additionally, a Nikon DSLR camera was

positioned above the branch line which recorded still images and video. Visual observations

without the assistance of a camera were also conducted.

Page 59: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

42

2.3 Appendix

2.3.1 Error analysis

The instrumentation installed at the facility, specifically the flow meters and WMS,

introduce sources of error in the boundary conditions assumed at the T-junction. Table 2-1

shows the measurement errors cited by the manufacturers of each flow meter used at the

facility. In addition, an estimated measurement error of the WMS itself is provided. The error

shown in the table related to the WMS is not the error induced by the WMS due to its

disturbance of the flow which is more difficult to quantify (see Appendix Section 2.3.2).

Device Type Use Measurement

error

Measurement

frequency [Hz]

Krohne IFM 1010

K

Krohne IFC 010

Electromagnetic Main flow meter ±0.05% of full scale

±0.5% of reading 8

Gundfos VFM 1-12

QT Vortex

Branch flow meter

Phase 1

Section 3.1

±3% of full scale Response time < 3

seconds

KOBOLD

DOM-S10HR21Z3

Positive-

Displacement

Branch flow meter

Phase 2

Section 3.3

±1% of reading 1

Wire-Mesh Sensor Electrical

conductance

Mixing

measurement

instrumentation

±3.62% Up to 10000

Table 2-1. Measurement error associated with flow meters used at the facility during various

experiments, and the WMS.

A number of errors contribute to the WMS output. These include calibration error,

quantization error and error due to electronic noise. An estimation of each of these errors is

generated based on a representative calibration and measurement (Ur = 100, 180 s, 1600 Hz)

with a 16 transmitter, 16 receiver WMS installed in the branch line of the LKE T-junction

facility.

Fluctuations in the signal due to electronic noise during the tap water calibration near

the center of the mesh at crossing-point (8,8) and near the wall at (16,8) are shown in Figure

2-13 along with the time series of each signal from the 2 s (3200 sample) calibration

measurement. During the low-end calibration of the WMS with deionized water the

fluctuations induced by electronic noise are much larger relative to the mean signal. For

example, near the center of the mesh at crossing-point (8,8), μ = 33.7 and σ = 3.1.

In the case of the representative high and low calibration measurements the systematic

quantization error of the 12-bit WMS signal near the center of the mesh and near the wall is

defined as the change in mixing scalar induced by a change of 0.5 in the raw signal. For both

near-center and near-wall crossing-points (8,8) and (16,8) this error is very small, at 0.024%

based on a maximum signal, in this case, of less than 75% of saturation (in order to avoid

cross-talk, as discussed previously).

Page 60: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

43

Figure 2-13. (top) Probability density function and normal-distribution fit at a central

crossing-point (8,8) and a near-wall crossing-point (16,8) extracted from a 2 s, 1600 Hz tap

water calibration measurement. (bottom) Time series of each signal.

Calibration error varies from trial to trial. The linearity of the WMS response has been

investigated by Rohde (2005) and Bulk (2012).[166, 167] The report of Rohde (2005)

describes linearity errors of up to ±3% within a similar range of conductivities as the working

fluids in the experiments presented herein.

Systematic error is incurred when running closed-loop experiments at the facility due

to the changing EC of the tap water in the main supply tank. At a velocity ratio of 100, for

example, measurements indicate a change in main tank conductivity of 1.4 μS/cm over a long

duration 660 s measurement. Given that the main flow conductivity chosen as the high-

reference for the calibration curve is the average conductivity during a measurements, the

maximum discrepancy in conductivity is ±0.7 μS/cm at the high end. In other words, the error

induced is largest when measuring water of EC equal to 100% tap water and is equal to

0.62%.18

For a large number of samples, statistical errors such as the quantization error and

error due to electronic noise are greatly reduced (as a function of √ ). Left over is the

calibration error due to the linearity of the signal and error incurred due to changing main

flow conductivity in closed-loop tests. These errors are dependent on the fluid EC. The

18

Since the deionized water in the branch line supply tank is unaffected by close-loop experiments, this error

vanishes at conductivities equal to the low calibration.

Page 61: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

44

overall error is estimated as a sum of the maximum possible errors incurred by both

components for this representative measurement.

2.3.2 Influence of the wire-mesh sensor

The WMS is an intrusive measurement technique. The electrode grid enhances

turbulent mixing at sufficient flow velocities and represents a pressure drop. The pressure

drop coefficients of each sensor, based on the tabulated data of Idel'chick (1966) for plane

grids19

is approximately 0.06. Furthermore, the percentage of cross-sectional free area is

approximately 94% for both sensors [168] The recirculation of flow through the WMS in the

branch line is not the most typical flow scenario for the sensor. Traditionally WMSs have

been positioned in pipelines with a dominant unidirectional axial flow (such as the T-junction

mixing pipe), however, more and more frequently three layer and dual-WMS are being

installed in such flows. The influence of the WMS has been reported in cross-flow T-junction

measurements by Walker (2009) [50]. As cylindrical blunt bodies it is likely that the

electrodes induce vortex shedding at a variety of frequencies based on the local velocity.

Figure 2-14 shows the WMS reducing turbulent penetration depth when installed near the T-

junction (2Db) observed using fluorescein dye.

Figure 2-14. Visually observed turbulent penetration depth at velocity ratios up to 900 with a

WMS installed in the branch (WMS) and without (w/o WMS) along with logarithmic fits.

2.3.3 Gain-magnitude frequency response

Mesh sensor technology is not limited in frequency resolution by thermal inertia, but

rather by the spatial averaging of the measured quantity within the space between transmitter

and receiver electrodes, which leads to a low-pass filtering. The cut-off frequency is inverse

proportional to the average residence time of the fluid in this space; it is therefore dependent

on the flow velocity. We assume that the diffusion timescale is significantly small, such that

it can be neglected, and focus entirely on the advection of ions by the flow. The volume-

averaging performed by the sensor results in the failure to resolve the true conductivity

19

Specifically, Diagram 8-9.

Page 62: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

45

fluctuation amplitude in the case of high frequency fluctuations transported through the mesh

grid.

A simple first-order low pass filter with transfer function,

, (2-4)

is sufficient to describe the frequency response of the sensor. The time constant, τ, of the

mesh sensor is defined as the quotient of the distance between electrode grids (1.5E-3 m) and

the fluid velocity. The real part of Eq. 2-4 is,

| |

√ . (2-5)

Assuming a fluid velocity of 0.05 m/s, τ = 3E-2 s. The gain-magnitude frequency response of

the mesh sensor, based on Eq. 2-5 is shown in Figure 2-15 (left) for multiple flow velocities,

while the PSD of the filter (based on τ = 3E-2) and of sinusoidal input is plotted in Figure

2-15 (right). Uncertainty in the frequency response of the mesh sensor is the result of

variability in local fluid velocity. For example, crossing points very near the wall will exhibit

a more damped power spectrum due to a combination of smaller eddies and local fluid

velocities.

Figure 2-15. Gain-magnitude frequency response of the WMS built for the LKE T-junction

facility for a number of fluid velocities.

Page 63: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

46

3 Outline of experiments and simulation

3.1 Phase 1 Experiments

3.1.1 Geometry

In the first phase of experiments at the LKE T-junction facility the main pipe was a

flanged acrylic glass tube with Dm = 50 mm, meeting the branch line in a sharp edged T-

junction. The inlet run of the main pipe was 32Dm. The branch line is a half meter long

acrylic glass tube of diameter Db = 50 mm or 26 mm depending on the experiment. There is

an inlet run of at least 20Db for the 50 mm diameter branch, or at least 38Db in the case of the

smaller 26 mm diameter branch. The inlet run in the branch line changes depending on the

WMS position. Acrylic spacer pieces, comprised of a single thick flange, or a short pipe

capped by two flanges is installed between the T-junction and the WMS to position the

sensor at various non-dimensional locations x/Dm or y/Db within the turbulent penetration

plume, see Figure 3-1. The mixing pipe, defined as the pipe immediately downstream of the

T-junction, is also of diameter 50 mm and is 10Dm (500 mm) in length.

Figure 3-1. Sketch of acrylic glass flanged spacer pieces installed in the branch line for

positioning of the WMS.

3.1.2 Test matrices

As a first clarification of the turbulent penetration phenomena, the test matrix in Table

3-1 was completed using a WMS installed at various branch positions. Main pipe and branch

line Reynolds numbers are indicated in Table 3-1 for each tested velocity ratio. Main flow

rates were held steady at 100 l/min or 180 l/min while the branch line flow rate was varied,

always within the laminar regime. No density stratification was imposed.20

For each flow

combination a single WMS was positioned at y = 1, 1.5, 2 or 4Db (where y is measured from

the center of the T-junction), such that a 2D data set spanning the 3rd dimension, the branch

axis, could be constructed. Figure 3-2 shows a schematic of the WMS positions and the

20

Small density differences may result from a small ΔT of the two working fluids, as delivered from the

building network.

Page 64: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

47

orientation of the coordinate axes. For some measurements a WMS was installed in the

mixing branch at x/Dm = 1, in addition to the branch line sensor, such that the two sensors

recorded the scalar mixing simultaneously.

Figure 3-2. Schematic of the T-junction with locations of the WMS measurement planes and

the orientation of the coordinate axes. Gravity runs through the page.

Ur Um

[m/s]

Ub

[m/s] Rem Reb

[l/min]

[l/min] y/Db

28.61 0.85 0.030 42441 1484 100 3.5 1, 1.5, 2, 4

33.31 0.85 0.025 42441 1274 100 3 1, 1.5, 2, 4

501 0.85 0.017 42441 849 100 2 1, 1.5, 2, 4

1002 1.53 0.0153 76394 764 180 1.8 2, 4

1502 1.53 0.0102 76394 509 180 1.2 2, 4

2002 1.53 7.64E-3 76394 382 180 0.9 2, 4

4002 1.53 3.82E-3 76394 191 180 0.45 4

Table 3-1. Test matrix for T-junction mixing tests with steady inlet conditions. 160 second

open loop measurements at 2500 Hz, 2170 second closed-loop measurements at 1600 Hz.

In addition to equal branch ratio experiments, tests with a 26 mm diameter branch line

were performed at velocity ratios from 50 to 100. Table 3-2 indicates the flow conditions for

which experimental data has been obtained in this geometry, i.e. a branch ratio of 0.52. The

Phase 1 T-junction configuration, with 26 mm branch line and velocity ratio 100, was

reproduced in the LES simulation, discussed in Section 1.1.

Ur Um

[m/s]

Ub

[m/s] Rem Reb

[l/min]

[l/min] y/Db

50 1.53 0.030 76394 792 180 0.97 2

75 1.53 0.020 76394 522 180 0.65 2

100* 1.53 0.015 76394 392 180 0.49 2

Table 3-2. 170 second measurements at 1600 Hz. *A 660 second duration measurement was

performed for this case for comparison with an LES simulation described in Section 1.1.

3.1.2.1 Pulsatile flow

Finally, artificial pulsations in the main pipe were induced as a means of studying

large scale (so-called global) instabilities. Such instabilities may be generated, for example,

by pump fluctuations. A thorough review of the physics of turbulent pulsating flow is outside

Page 65: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

48

of the scope of this work. It should be noted, however, that a tremendous body of work

spanning decades does exist, mostly related to pipe flows.[169, 170]

In the case of a leaking valve scenario, such pulsatile instabilities could result in or

exacerbate the problem of thermal cycling in the branch line. Favored frequencies in the

mixing spectrum, which have been observed in prior experiments, as discussed in the

introduction, may be the result of global instabilities in the flow. A regularly extending and

receding turbulent penetration could manifest a strongly favored frequency at which the

scalar fluctuations are strongest. In order to investigate such a situation, forced oscillations

have been introduced by means of a sinusoidal voltage input to the frequency controller of

the main pipe supply pump, which induces sinusoidal flow rate oscillations with a high

degree of accuracy,

(3-1)

where Qm(t) is the time-varying main pipe volume flow rate, a is the amplitude of the

oscillations, f is the frequency and is the mean, baseline main pipe volume flow rate. A

series of measurements were recorded for different amplitudes and frequencies of the volume

flux oscillations. The transfer from sine wave input via the pump controller to a time-

dependent bulk velocity, which oscillates accordingly, is not entirely linear. Still, the

hydraulic pressure drop is nearly proportional to the square of the flow rate and the pump

head is nearly proportional to the square of the impeller rotation speed. The flow rate under

steady state conditions is thus proportional to the pump rotation speed, which is in turn

proportional to the control voltage supplied to the frequency transformer. The real transfer

function is slightly more complicated, especially due to the effect of flow acceleration and

deceleration, and some distortions from the sine wave profile may take place. It was

nonetheless decided to remain with a sine wave control of the pump and check the result via

the flow meter output. It was confirmed that the flow follows a traditional sine wave within

the time resolution (8 Hz) and accuracy (±0.05% of full scale and ±0.5% of the reading) of

the flow meter.

The velocity profile in pulsatile flows is dependent on the Womersley number21

,

defined as the ratio of transient inertial force to viscous force [171].

(3-2)

where ω is the angular frequency, the characteristic length is Dm, the main pipe diameter, and

ν is the kinematic viscosity. The definition also extends to the Reynolds and Strouhaul

numbers in the following manner,

√ (3-3)

21

Originally introduced in the context of arterial flows.

Page 66: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

49

Where St is defined here, based on the main pipe as . Note that dynamic

similarity is guaranteed given the same Reynolds and Womserley numbers. These numbers

are reported for the pulsatile flow experiments in Table 3-3.

The 18-measurement test matrix chosen for the global instability measurements are of

exaggerated amplitude for the purpose of simplifying the understanding of trends under such

mixing conditions, see Table 3-3. Main pipe velocity oscillation amplitudes range from ±5%

to ±20% of the average velocity with a frequency of 0.1 Hz, while information regarding the

effect of the frequency of the main pipe oscillations is garnered from measurements with an

amplitude of ±10% at 0.1, 0.2, and 0.3 Hz. The WMS data was collected over 170 seconds at

1600 Hz. The T-junction geometry remains in the aforementioned configuration with all

acrylic pipes leading to the junction being 50 mm in inner diameter. The velocity ratio for all

measurements was 34.5.

Hz

Oscillation

amplitude, a

(% of )

Oscillation

amplitude

[m/s]

[-]

St

[-]

α

[-]

[m/s]

[m/s] y/Db

0.1

5

15

20

0.076

0.23

0.31

42441 0.006 39.6 0.85 0.025

1,

1.5,

2

0.1,

0.2,

0.3

10 0.15 42441

0.006

0.012

0.018

39.6

56.0

68.6

0.85 0.025

1,

1.5,

2

Table 3-3. Test matrix for T-junction mixing tests with pulsed flow in the main pipe.

Page 67: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

50

3.2 Phase 1 Large Eddy Simulation

3.2.1 Simulation test case

A representative case of turbulent penetration was chosen from the Phase 1

experimental test matrices at the LKE T-junction facility to be reproduced by time-resolved

CFD simulation. While velocity ratios ranging from around 25 to 3000 have been

investigated experimentally in this thesis, a ratio of 100 was selected for comparison with a

large eddy simulation (LES). The velocity ratio, mass flow rate ratio, and momentum ratio

are, , , and , respectively. The momentum ratio is defined as in

the work of Kamide (2009) [46]. Note that the Reynolds number ratio is also 100 since the

fluid properties are assumed to be nearly identical. The main flow is firmly in the turbulent

regime with a Reynolds number of approximately 76,000 while the branch line in-flow is

laminar, Reb = 400. For comparison, momentum ratios above 1.35 represent a cross flow

mixing case known as a wall jet while Zboray (2011) has shown the onset of turbulent

penetration in earnest at a momentum ratio of approximately 700 [141].

This test case was investigated with a long duration WMS measurement at the LKE

T-junction facility. The measurement is 660 seconds in duration with a recording frequency

of 1600 Hz at a non-dimensional location y = 2Db in the branch line. The WMS data at this

position indicated a broad peak in the scalar mixing power spectrum near the downstream

wall of the branch centered around 6 Hz, indicating the presence of some kind of regular flow

instability.

3.2.2 Simulation methodology

The study of thermal fatigue due to turbulent mixing necessitates resolving time-

dependent temperature fields since the internal stresses induced in steel pipe walls are

dependent on the frequency of thermal fluctuations in the fluid. Far more computationally

expensive than unsteady RANS simulations and considerably cheaper than DNS, LES is the

tool of choice for simulating T-junction mixing in the nuclear engineering domain.

Simulating turbulent penetration in T-junctions requires simulating long flow times which put

the work at the periphery of what is today computationally reasonable. The 13 seconds of

simulation presented herein are the result of over 100,000 CPU-hours. The simulation was

performed with the commercial code Star-CCM+ version 8.04.007-R8 from CD-Adapco.

The simulation was setup as follows. An implicit LES with 2nd

order temporal

discretization and segregated solver with segregated species was utilized. A passive scalar, θ,

was included and solved for with the help of an additional transport equation using the

default value of Turbulent Schmidt number of 0.9, such that tap water was assigned a value

of θ = 1 and deionized water , θ = 0. Five inner iterations were performed for each time step

of 2.5E-3 seconds.

3.2.3 Governing equations

Page 68: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

51

The LES methodology is based on separation of scales into resolved and subgrid

components. Fully resolved are large scale turbulences in time and space while modeling of

smaller, subgrid scale (SGS) eddies is achieved by incorporating an added viscosity. The

following discussion draws from the derivations and discussions found in the textbooks of

Pope (2000) and Sagaut (2007) and lecture notes by Celik (1999) [172-174].

Figure 3-3. A simplified schematic of scale separation in physical space and Fourier space

from Sagaut (2007) [173].

The LES governing equations and those of unsteady RANS follow an identical

formulation. However, rather than averaging the Navier-Stokes equations in time, LES

modifies these equations in the same way but in the spatial domain in an operation known as

filtering. The sub-gird-scale methodology is as follows, given a field u(x,t),

𝑢 𝒙 𝑢 𝒙 𝑢′ 𝒙 (3-4)

where 𝑢 𝒙 is resolved, in this case via a spatial filtering process described below and

𝑢 𝒙 is the unresolved SGS component, see Figure 3-3. In general, a variable 𝑢 𝒙 is

filtered in 3D space utilizing a convolution kernel G, in the following operation,

𝑢 𝒙 ∫ ∫ 𝑢 𝒙 ′

. (3-5)

The conservation equations for momentum and kinetic energy, as well as continuity

are filtered. A decomposition of the non-linear filtered product 𝑢 𝑢 is performed, the

decomposition being named after Leonard (1974) [175], which leads to the residual stress

tensor,

�� �� 𝑢 𝑢 . (3-6)

Noting the residual kinetic energy is,

, (3-7)

Page 69: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

52

it follows that the deviatoric component22

of the residual stress tensor (from the anisotropic

part of the stress tensor) is,

. (3-8)

The hydrostatic, isotropic component is incorporated into to the filtered static pressure term.

It is , often referred to as the subgrid stress tensor, that must be modeled. Modeling of the

tensor is based on the hypothesis proposed by Boussinesq known as Boussinesq‘s turbulent

viscosity hypothesis which assumes a linear relationship between subgrid stress tensor

written above, and strain rate tensor [176]. Utilizing this hypothesis the previous equation

becomes,

, (3-9)

where is the eddy viscosity and is the resolved strain rate tensor,

(

). (3-10)

Essentially, the SGS turbulence is modeled via an eddy viscosity. Two major eddy

viscosity models are in use today for computing , they are the Smagorinsky model and the

Wall-Adapting Local Eddy-Viscosity (WALE) model. The WALE model is the default

model of its kind in the Star-CCM+ software, it has seen successful application for a number

of years. Ma (2008), for example, made an assessment of SGS models with experimental

backing, in simulations of turbulent flow in a water turbine; the WALE model performed the

best [177]. Furthermore, in the Vattenfall T-junction benchmark exercise LES simulations

utilizing the WALE model placed higher than those using the Smagorinsky model ranked by

comparison to thermocouple data [47]. Menter (2012) notes that the WALE model is the

simplest model able to automatically provide a zero eddy viscosity in laminar shear flows

which ―is especially important when computing flows with laminar turbulent transition‖

[178].

The simulation presented in this thesis was performed using the WALE model which

calculates the eddy viscosity as a function of velocity gradient,

(3-11)

where,

(3-12)

22

The other component being hydrostatic stresses.

Page 70: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

53

is based on the square of the velocity gradient tensor,

(3-13)

(3-14)

The model accounts ―for the effects of both the strain and the rotation rate of the smallest

resolved turbulent fluctuations;‖ a model constant Cw, has been left unchanged with a value

of 0.544 [179]. This value of Cw has been shown to work well for a variety of turbulent

conditions, including channel flows [180].

3.2.4 Computational grid

A decision on the cell base size of the mesh was made based on the work of Addad

(2008) on optimal unstructured meshing for LES, later demonstrated in cross-flow mixing

simulations for the Vattenfall benchmark exercise published by Kuczaj (2010) [85, 181].

Essentially, the degree of grid refinement is chosen based on knowledge of turbulent length

scales. Since LES filtering addresses eddies on the order of the Kolmogorov scale by

implementing SGS models, the length scales of importance for meshing is rather the Taylor

microscale. This scale is defined as, √ 𝑢 , where the R subscript indicates the

values are obtained by a stead state RANS simulation while k and ε are the turbulent kinetic

energy and dissipation, respectively. A RANS simulation was performed for the Ur = 100 test

case using the Reynolds Stress Model. A field function was generated to compute the Taylor

microscale over the entire domain. The minimum Taylor microscale was approximately

0.275 mm. This data was converted into a mesh size table such that the polyhedral mesh base

size was made to vary with the local Taylor microscale. All regions with λR > 1 mm, such as

in the laminar portion of the branch line, were set to be meshed with a cell base size of 1 mm.

The result is a computational grid with half of the cells compared to one with a universal base

size determined by the smallest Taylor microscale. The mesh contains 6,451,848 polyhedral

cells.

The issue of wall treatment is critical. The decision on the choice of mesh

characteristics were taken with the findings of Jayaraju (2010) in mind in which LES

simulations were carried out using the Star-CCM+ and the WALE SGS model [182]. In his

work a strong case is made for the necessity of a fully resolved wall for simulating cross-flow

T-junction mixing. Specifically, an under-prediction of fluctuations in the temperature and

velocity fields in the vicinity of the walls and in the wall heat-flux was observed. Since wall

functions are derived from steady state conditions in pipe flow, they cannot necessarily be

trusted in more complex mixing geometries.

The extent of the computational domain is limited in the axis of the main pipe (x) and

extended beyond the region of expected deepest turbulent penetration, based on experimental

observations, along the branch axis (y). The domain resides in the region (-0.02 m < x < 0.03

m) and (-0.025 m < y < 0.13 m) where the origin lies in the center of the T-junction. In non-

Page 71: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

54

dimensional terms, the branch boundary resides at y = 5Db. The decision to limit the extent of

the main pipe was primarily an issue of computational savings. Confidence in the domain

size was gained via experimental observations of this flow case with fluorescein dye which

did not appear to show mixing upstream of the branch in the main pipe. Figure 3-4 shows the

mesh as seen in the z = 0 mid-plane of the T-junction.

Figure 3-4. 6.45 million cell Polyhedral-prismatic mesh viewed in the z = 0 plane.

3.2.5 Initialization

Boundary conditions were not measured experimentally beyond bulk values recorded

by flow meters at the facility. Without detailed information about the velocity profiles at the

boundary of the LES domain, a fully developed profile is adopted and recognized as a

potential source of discrepancy when comparing the simulation and experimental results.

Fully developed profiles are generated for both the branch and main pipe in periodic k-ε

steady state RANS simulations. The periodic domains are of length 2πD as a means to

capture the effects of the largest potential eddies in the flow. The converged cross-sectional

velocity profiles from these precursory simulations are imported into the full T-junction

domain. Artificial turbulence is then induced in the main pipe (but not in the laminar branch

line) by the Synthetic Eddy Method which superimposes isotropic velocity field fluctuations

on the steady, developed velocity field from the periodic domain simulation [180]. This

method requires as input the turbulent intensity I, and turbulent length scale l, both of which

have been calculated based on correlations for pipe flow as follows,

, where

the Reynolds number is defined as a function of the hydraulic diameter, ⁄ ,

and l = 0.038Dm, respectively. The main pipe turbulent intensity at the inlet boundary is

3.92% and the turbulent length scale is 1.9E-3 m.

Initialization proceeded as follows. In order to reduce the time necessary for the LES

simulation to start producing meaningful results, a steady state RANS simulation was

performed using the Reynolds Stress Model. Once converged, the switch was made to LES

and approximately 3 seconds of flow time were simulation in a 12.3 million cell mesh with

uniform polyhedral base size of 0.275 mm, equivalent to the smallest Taylor microscale

encountered within the domain. After confirming the convergence of the simulation, and

Page 72: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

55

observing qualitatively realistic mixing in the branch line as expected from experimental

results, the mesh was transitioned to the 6.45 million cell grid with polyhedral size based on

the local Taylor microscale, with an upper cell base size limit of 1 mm, for the remainder of

the simulation.

Page 73: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

56

3.3 Phase 2 Experiments

3.3.1 Geometry

In Phase 2, the LKE T-junction geometry was modified such that the main pipe was

reshaped into a rectangular duct.23

The modified main pipe geometry, demonstrated by

Robert (1992), Deutsch (1997) and, among others, researchers at the INSS in Japan, was

implemented at the LKE T-junction by shrinking the main pipe cross-section in order to

achieve higher flow velocities (sometimes referred to as ‗blocking‘) [105-107, 124]. The new

main pipe conduit was manufactured from two acrylic halves bonded to form a rectangular

flow area 10 mm in width and 50 mm in height, with hydraulic diameter 16.67 mm

(henceforth, Dm). The hot or cold leg in a PWR has an inner diameter of, for example,

0.74 m, therefore the curvature seen at the junction by a standard 2‖ (50 mm) or 4‖ (102 mm)

diameter branch line is small. For this reason we find acceptable the use of a rectangular

cross section of the main pipe for the purpose of turbulent penetration experiments.

Furthermore, Tanimoto (2002) concluded in-part that, ―[t]he penetration length is

dependent not on the diameter ratio of the branch pipe to the main loop but on the

diameter of the branch pipe‖ [111]. Intersecting the main pipe on its tall side is the same

Db = 26 mm branch line from Phase 1. The inlet run in the main pipe is now 34.1Dm. A

sketch of the Phase 2 geometry is shown in Figure 3-5.

Figure 3-5. Sketch of Phase 2 LKE T-junction geometry with 10 mm rounded edge, 10 mm

wide (by 50 mm tall) main and mixing pipe and 26 mm diameter branch line. Shown are the x

and y-axes located at the origin from which dimensionless branch position is defined. Gravity

is directed opposite the z-axis. Red arrows indicate main flow, blue arrows indicate branch

flow. View (a, top) shows the x,y plane, while (b, bottom) shows the x,z plane.

23

For the sake of continuity, the nomenclature will remain unchanged; ‗main pipe‘ still refers to the conduit in

which turbulent main flow is traveling prior to its arrival at the T-junction.

Page 74: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

57

The second design phase also saw the introduction of modifications to the junction

geometry. Recognizing the potentially large importance of the edge geometry at the

intersection of the main pipe and branch line, an additional T-junction with a curved edge

with constant radius of curvature at the intersection with the branch line was manufactured

(also visible in Figure 3-5). The branch line is connected to the junction, as usual, with

acrylic spacer pipes for the adjustment of the WMS position. Only a single WMS is installed

in the branch at a given time and no WMS is present downstream in the mixing pipe.

3.3.2 Test matrices

The introduction of a reduced area main pipe in Phase 2 of the LKE T-junction

facility design enabled higher main flow velocities and higher velocity ratios, up to 3000, to

be explored. High speed camera footage was recorded for velocity ratios ranging from 50 to

900 while WMS measurements were performed between Ur = 400 and 3000. Experiments at

these higher velocity ratios look to describe the turbulent mixing behavior deeper into the

branch line. Additionally, a two-axis test matrix explores, with WMSs, T-junction geometry

and density effects by introducing a deionized water-sucrose solution as the branch line

working fluid.

The following text matrix is intended to explore higher velocity ratios, up to 3000. In

addition, the effect of main flow Reynolds number is investigated at a fixed velocity ratio of

2000. All tests were carried out at two branch positions, y = 3.73Db and 5.73Db, where y is

the distance from the precipice of the branch line (the new origin of the coordinate system).

The large amount of turbulent penetration at these velocity ratios means that measurements at

positions close to the junction (e.g. 1, 1.5, and 2Db etc.) are trivial; in the branch line at these

positions the scalar is essentially 100% tap water. The full test matrix was carried out in both

sharp edged and rounded edge T-junctions for a total of 20 measurements, see Table 3-4.

Each measurement was recorded at 1600 Hz for a duration of 180 s.

Ur Qm Um Rem Qb Ub Reb

1000 173.49 5.78 96000 0.184 0.0058 149.76

2000 117.47 3.92 65000 0.062 0.0019 50.7

2000 144.58 4.82 80000 0.077 0.0024 62.4

2000 173.49 5.78 96000 0.092 0.0028 74.88

3000 173.49 5.78 96000 0.061 0.0019 49.92

Table 3-4. Test matrix with 26 mm diameter branch line.

Most academic research in the field of T-junction mixing utilizes sharp edged T-

junctions, while forged stainless steel components in NPPs tend have a large radius of

curvature at the junction. Two T-junctions have been tested, one with sharp edges and one

with a curved edge (r = 10 mm). The two junctions share the same hydraulic diameter, 16.67

mm, in the main and mixing pipe. Henceforth, ‗T-junction geometry‘ will refer to the degree

to which the edges of the intersecting channels are rounded or curved, sometimes called the

welding chamfer although it is equivalent to a fillet.

Page 75: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

58

3.3.2.1 Density and edge geometry

Another important parameter in T-junction mixing, as it relates to the reality in NPPs

is the density of the coolant and its temperature dependence. In the case of a leaking valve

scenario, the water entering the branch line is of a much lower temperature than the main

flow, especially in the case when the hot flow is directly from the hot or cold leg of the RCS.

In the normally stagnant fluid behind valves in branch lines heat loses have a strong negative

impact on fluid temperatures above ambient. Density difference generated by addition of

sucrose to the deionized water supply of the branch line at the LKE T-junction has allowed

for a number of tests at the facility with density differences , of 0.024 to

0.088 (density ratios ρr, of 0.98 to 0.91).

Adiabatic single phase mixing experiments have been carried out in the past with

artificial density difference in a variety of conduit geometries including T-junctions [110,

158]. While the mixing scenario of interest calls for a colder branch flow to meet a hotter

main flow at the T-junction, in adiabatic tests it is therefore proposed to increase the density

of the branch flow artificially using a solute, typically sucrose or calcium chloride. In the

presented experiments, sucrose is chosen to increase the density of the branch flow. Large

density differences between two incoming flows were achieved which would otherwise

require a pressurized high-temperature test facility. Figure 3-6 shows the increase in the

relative density of water with increased sucrose mass concentration; at 20ºC at around 30%

concentration by mass the solution becomes saturated.

ε

[-]

Edge radius, r

[mm] Measurement duration (s)

Iso-dense 0 0 & 10 170 (closed-loop)

Sucrose 1 0.024 0 & 10 60 (once-through)

Sucrose 2 0.051 0 & 10 60 (once-through)

Sucrose 3 0.070 0 & 10 60 (once-through)

Sucrose 4 0.088 0 & 10 60 (once-through)

Table 3-5. Test matrix including density stratification and T-junction geometry.

Table 3-5 shows some details of the density-enhanced experiments performed in

Phase 2. All tests were carried out at a velocity ratio of 2000 where Um = 3.92 m/s,

Rem = 6.5E4 and Ub = 2.0E-3 m/s. Wire-mesh sensor measurements were performed, as

above, at two branch positions, y = 3.73Db and 5.73Db for every flow case. Data was

recorded at a frequency of 1600 Hz.

Page 76: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

59

Figure 3-6. Relative density of deionized water vs. weight concentration of sucrose. Beyond

30% sucrose by weight, precipitation of sucrose in the solution occurs.

The use of sucrose rather than calcium chloride (which would enable greater densities

to be reached) was required due to the signal saturation experienced by the WMS in the

presence of highly conductive fluids. While realistic density differences may be approached,

viscosities found at high temperatures in LWR piping cannot be appropriately matched. Fluid

properties for all tests are shown in Table 3-6 along with two representative reactor cases

from a PWR which have not been investigated at the LKE T-junction facility. Viscosity

values are gathered from literature based on the density of the solution. The densities and

viscosities are averages between the values during the individual WMS experiments at

3.73Db and 5.73Db.

Reb

[-]

Rer

[-]

ρm

[kg/m3]

ρb

[kg/m3]

ρr

[-]

ε¹

[-]

ηb

[cP]

ηr

[-]

Frb

[-]

Ri

[-]

ΔT

[K]

Iso-dense 51 1.3E3 997 997-998 1 0 1.00 1 - 0

Sucrose 1 43 1.5E3 997 1021 0.98 0.024 1.17 0.85 3.3 0.092 0

Sucrose 2 34 1.9E3 997 1050.5 0.95 0.051 1.48 0.67 1.5 0.44 0

Sucrose 3 28 2.3E3 997 1072.5 0.93 0.070 1.78 0.55 1.1 0.83 0

Sucrose 4 23 2.8E3 997 1093.5 0.91 0.088 2.21 0.45 0.9 1.2 0

Reactor

1* 34 1.2E6 950 1005* 0.945 0.055 0.99 0.25 - - 100

Reactor

2* - - 851 1005* 0.847 0.15 0.99 0.15 - - 200

Table 3-6. Fluid properties related to the test matrix shown in Table 3-5. *Examples based on

a branch line fluid temperature of 20°C at 15 MPa pressure. Estimated Reynolds numbers.

Density difference ε, is an important parameter which has a significant impact on

turbulent mixing in T-junctions. Furthermore, the Froude number is commonly used to

characterize flows with large density differences between species. The dimensionless number

is a function of velocity, reduced gravity, and a characteristic dimension. We use the

following definition of the Froude number based on branch line values,

Page 77: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

60

√ , (3-15)

where ′ is the reduced gravity defined as gravity multiplied by the density difference,

where the density difference is defined as,

, (3-16)

See Table 3-6 for Fr numbers. Also referenced is the Richardson number, defined here as

. (3-17)

This definition of the Richardson number is identical to that found in Chapuliot (2005) for

cross-flow mixing. The values found here, however, are three orders of magnitude higher

owing to the slow branch line velocity [1].

Page 78: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

61

3.4 Regarding the orientation of presented results

The orientation of the WMS data contour plots, as well as cross-sectional simulation

results from the branch line, shown in the following chapter are always oriented as sketched

in Figure 3-7, unless explicitly stated. Regarding the definition of the coordinate axes

orientation, the orientation is defined as the center of junction (i.e. intersection of the center

of the main pipe and center of the branch line) in Phase 1, and in the middle of the branch

line at the precipice between the main pipe and branch line in Phase 2.24

The x-axis runs from

the origin down the mixing branch, the y-axis from the origin into the branch line, and the z-

axis is oriented, as usual, with +z pointing in the direction opposite to gravity. Figure 3-7

shows two viewing perspectives, one describing the viewing orientation of WMS data for the

case of installation in the mixing pipe, and the other for results in the branch line.

Figure 3-7. Representative position of the WMS and data viewing orientation for installation

in the mixing pipe (left) looking away from the T-junction and branch line (right) looking

towards the T-junction.

When WMS results are plotted, they are plotted on a normalized scale based on the

tap water (θ = 1) and deionized water (θ = 0) signal values at every crossing point. The color

bar in contour plots is scaled such that tap water is red and deionized water is blue.

Furthermore, when specific crossing points are described, for example when showing a time

sequence of the signal at a particular location, or a power spectra, they are referenced as a

Cartesian coordinate (Transmitter t, receiver r), recall Figure 2-8. Often, positions in the

branch line are referred to as being near the ―upstream wall‖ or ―downstream wall‖ based on

the orientation seen in Figure 3-7.

The RMS of the scalar is always calculated as,

, (3-18)

where is the average in time of the scalar during the measurement period. The scalar RMS

is the square root of the scalar variance which is equivalent to the integral over the PSD of the

scalar signal.

24

Regarding the definition of the origin, in the case of an artificially reduced main flow area as in the case of the

rectangular cross section of the main pipe in Phase 2, the definition chosen here is common practice.

Page 79: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

62

4 Turbulent penetration with in-flow

The following chapter draws on findings from three sources described in detail in the

previous chapter: 1. wire-mesh sensor measurements performed at the LKE T-junction, 2.

Star-CCM+ LES simulation performed at a velocity ratio of 100 with a 50 mm main pipe and

26 mm branch line, 3. high-speed camera footage. Note that all discussion is related to iso-

dense mixing, until Section 4.8 where experiments with a heavier branch flow are analyzed.

4.1 Turbulent penetration and how it differs from cross-flow mixing

Turbulent penetration in T-junctions is a flow configuration categorically different

than cross-flow mixing. Although, in the case of turbulent penetration, mixing between

branch and main flows continues in the mixing pipe downstream of the T-junction, its

character is unlike mixing at the same location at velocity ratios near unity (i.e. cross-flow

mixing). To varying degrees on a spectrum between the dead-leg case ( ) and that of

velocity ratios near the onset of turbulent penetration ( ), turbulent penetration in the

branch line renders flow at and downstream of the T-junction ‗pre-mixed.‘ The consequence

being that in the mixing pipe large amplitude scalar fluctuations are unlikely, if not

impossible, and thereby the mixing pipe is no longer a region of interest with regards to

thermal fatigue.

Simultaneous measurements by WMSs in the branch line and mixing pipe at the same

non-dimensional positions were analyzed. Figure 4-1 (left) shows comparison between the

scalar fluctuation profile in the mixing pipe and branch line, both one diameter from the T-

junction at a velocity ratio of 33.3 where the main tap water flow is normalized to 1 and the

deionized branch flow to 0 (sharp-edged T-junction, Dm = 50 mm Db = 50 mm). Maximum

scalar fluctuations present in the mixing pipe at this position are approximately half of that

found at an equal distance into the branch line. Further mitigating the relevance of these

fluctuations are their high frequency nature. A normalized PSD from two near-wall positions

in both the branch line and mixing pipe are plotted in Figure 4-1 (right). Scalar fluctuations in

the mixing pipe at x =1.0Dm are characterized by high frequencies and low amplitudes; while

in the branch line, damped high frequencies result in low frequencies (< 10 Hz) contributing

far more to the scalar RMS, relative to the same location in the mixing branch.

Page 80: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

63

Figure 4-1. (left) RMS of the scalar in the mixing pipe at x = 1.0Dm (top) and the branch line

at y = 1.0Db (bottom). (right) PSD of scalar fluctuations near the wall (mm values represent

distance from the wall) indicated by red dots in contours on left, Ur = 33.3.

Turbulent penetration results in the flow exiting the T-junction through the mixing

pipe in what could be described, in the nomenclature of cross-flow T-junction mixing, as a

weak wall jet. Upon exiting the branch line the fluid is accelerated down the mixing pipe in a

concentrated, crescent shaped near-wall region, seen in Figure 4-1 (left, top), likely dictated

by the turbulent velocity profile in the main flow. A time series near the downstream wall25

in

the branch line and mixing pipe are show in Figure 4-2 for a velocity ratio of 50. In the

mixing pipe at x = 1.0Dm fluctuations are notable for their high frequency and low amplitude

nature visible already before a more detailed spectral analysis. Meanwhile, in the branch,

pulsations of main flow (i.e. turbulent penetration), some with amplitudes of 50% of the

scalar, are visible on a much larger characteristic time scale.

25

According to the measurement orientation, see Figure 3-7.

Page 81: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

64

Figure 4-2. Signal from crossing points near the wall (16,31) in the mixing pipe and in the

branch line.

The profile of the mixing spectra downstream of the T-junction during turbulent

penetration is found to be highly similar, in slope and shape, to the spectra at similar locations

in cross-flow mixing cases. Figure 4-3 shows a comparison of the PSD at a crossing point

near the wall in the mixing branch for the case of cross-flow mixing at Ur = 1 in the facility

of Walker (2009) versus in the branch line of the LKE T-junction facility for the case

turbulent penetration at Ur = 50 26

[50]. Both flow configurations result in a nearly flat

spectrum across low frequencies in the mixing pipe followed by a steepening around 10 Hz.

The only major difference between the spectra is that the amplitude across all frequencies is

significantly higher in the cross-flow mixing scenario, indicating a higher scalar RMS, as

anticipated. The PSD near the downstream wall in the branch line at Ur = 50 is also plotted

and, as suspected, shows weighting heavily on the low frequency side with a rapidly damped

high frequency contribution akin to a low-pass filtering. This is expected in the sense that the

average Reynolds number in the branch line, even in the case of violent turbulent penetration

at high Ur, is far less than in the mixing pipe. A reduction in the turbulent kinetic energy of

the flow and the presence of a recirculation region (i.e. where thermal cycling occurs) leads

naturally to less high frequency high amplitude fluctuations and more power in the low

frequencies of the spectrum.

Figure 4-3 A comparison of PSD at locations of high RMS in the branch line (y/Db) and

mixing pipe (x/Dm) at Ur = 50 (Dm = Db = 50 mm) compared with the results of Zboray

(2011) at Ur = 1 (Um = 0.5 m/s, Dm = 52 mm) in the mixing pipe [141].

Time resolved CFD simulations (Dm = 50 mm, Db = 26 mm) corroborate these

experimental measurements (Dm = 50 mm, Db = 50 mm) with regards to mixing downstream

of the T-junction for a non-unity branch ratio geometry. Having outlined the flow behavior in

the mixing pipe in the case of turbulent penetration and its apparent lack of threatening

characteristics (in a HCTF-sense), the remainder of this chapter will focus on the turbulent

mixing occurring in the branch line itself.

26

The Walker (2009) facility having main, branch, and mixing pipe inner diameters of 52 mm, marginally larger

than the 50 mm pipes of the LKE T-junction pipes, as configured in this measurement.

Page 82: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

65

As discussed initially in Section 1.3.3.1, turbulent penetration depth is a value of

interest in T-junction mixing studies. Visual observations in a 26 mm branch line without

accompanying WMS, by employing a fluorescein-doped main flow, indicate a logarithmic

dependence of turbulent penetration depth on the velocity ratio, see Figure 4-4. A fit function

has been generated for velocity ratios up to 900 and main flow Reynolds numbers up to

1.3E5,

, (4-1)

with an R2 of 0.9319. An increase in main flow Reynolds number at a fixed velocity ratio is

also shown to enhance the depth of main flow ingress into the branch line.

Figure 4-4. Turbulent penetration depth in an empty (no WMS) 26 mm branch line as a

function of velocity ratio with logarithmic fit.

There are some typical distributions of the time-averaged scalar and fluctuating scalar

(RMS) profiles in the branch line at ‗shallow‘ depths during the onset of turbulent

penetration, however, prior to a description of the flow behavior in a mostly time-averaged

sense, it is valuable to understand the highly transient nature of turbulent penetration. Figure

4-5 and Figure 4-6 show two pairs of such scalar distributions, which, while dissimilar, were

both manifest within a short period of time. Nonetheless, the flow is not totally chaotic and

on a large enough time scale an average profile with a distinct, interpretable scalar

distribution is measured. It is evident from the measurements that the main flow appears in

the branch (at any significant depth) initially along the downstream wall of the branch. This

has been shown to be a valid description for all velocity ratios, branch ratios, and branch

geometries explored. Since the WMS was never installed closer than 1.0Db in the branch line,

the LES is relied on for a more complete understanding of the flow behavior at the mouth of

the branch line.

Page 83: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

66

Figure 4-5. Two frames from a measurement in a 26 mm branch at 5.73Db exhibiting vastly

different scalar distributions for a given boundary condition, in this case Ur = 2000.

Figure 4-6. Two photographs separated in time by 100 ms showing a sudden change in

turbulent penetration depth in a 26 mm branch line. Main flow is visible in green.

Page 84: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

67

4.2 Flow characteristics at the entrance to the branch line

The flow at the T-junction is reminiscent of turbulent cavity flows. A great deal of

research has focused on compressible cavity flows in the context of aeronautical engineering

acoustics. Analogously, these flows exhibit wavy instabilities developing after the main flow

separates at the upstream wall of the cavity, which then are sheared by the downstream wall

of the cavity. At some velocity ratios, such as the simulated case of Ur = 100, the shear layer

also contains large instabilities in the scalar field, rather than the velocity field alone. Figure

4-7 shows a snapshot of the uj velocity field (contour rings) and scalar field (filled contour) in

the z = 0 mid-plane of the T-junction at the branch mouth where the instabilities are clearly

seen.

Figure 4-7. Free shear layer instabilities in velocity (contour lines) and scalar field (filled

contour) forming after main flow boundary layer separation (flow right to left). Instantaneous

snapshot with three large waves each with positive and negative uj velocity areas.

The limited distance (≤ 26 mm) and time (approximately ≤ 20 ms) in which the

instabilities have to form when crossing the branch mouth results in waves which, while

breaking, never fully roll over to form the large billows often associated with such shear

layers. They are reminiscent of the famous turbulent mixing shadowgraphs downstream of a

splitter plate by Brown (1974), or of certain atmospheric flows, see Figure 4-8 [157]. The

area of low pressure behind each wave entrains branch fluid into the turbulent main flow, a

phenomena described variously as ‗gulping,‘ ‗enfolding,‘ or ‗entanglement;‘ some of this

fluid exits the T-junction through the mixing pipe. While on the other hand, main flow wave

crests, upon being sheared by the downstream wall of the branch may catapult into the branch

line. The result of this countercurrent flow in the branch line is a second shear layer growing

into the center of the branch in the broad recirculation zone. While the region of shear itself is

unbounded and therefore could be described as a free shear layer, the proximity of the wall

boundary layer plays an important role. This second shear layer defines the central axis of the

recirculation zone.

Page 85: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

68

Figure 4-8. Instabilities forming at the branch mouth in the mid-plane of the T-junction (left),

similar, slower instabilities may exist in air masses (right, photo credit: Brooks Martner,

NOAA/ETL).

Figure 4-9. Top view of the T-junction from LES. Well-defined free shear layer growing

with progression across the mouth of the branch shown in a contour plot of average ui (left)

and, similarly, a second shear layer within the branch visible in the average uj field (right).

Both shear layers are visible in average velocity contour plots in Figure 4-9, recall

that by definition a shear layer is a region of significant velocity gradient. The shear layers

can also be isolated by plotting the magnitude of the strain rate tensor S, calculated from the

velocity field. Where the tensor is defined as,

(

), (4-2)

with magnitude based on the inner product,

‖ ‖ √ (4-3)

The strain rate magnitude of the shear layer due to turbulent penetration in the branch is

significantly less than the fast moving free shear layer over the mouth of the branch. It is for

this reason that in Figure 4-10 two different color bar scales are used to discern both zones of

elevated strain rate magnitude.

Page 86: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

69

Figure 4-10. Top view of the T-junction from LES. Magnitude of the strain rate tensor is

plotted on different scales such that both shear layers are visible (circled in red).

Time-averaged analysis of the two shear layers occurring in turbulent penetration

mixing is performed with simulations results from the z = 0 plane. The two shear layers are

visible in the average ui and uj velocity fields, respectively. As the main flow boundary layer

separates from the wall a free shear layer is born which precipitates the formation of coherent

large scale instabilities. The strong velocity shear between the fast main flow and branch line

flow, which at the precipice of the branch exhibits near-zero mean ui velocity, causes the

shear layer to grow. The velocity difference is ΔU = Umi - Ubi. The main flow free stream

velocity is defined at the center of the velocity profile in the main branch where, Umi = 1.83

m/s. The mean velocity is equal to ⁄ m/s and the constant

⁄ . The branch line free stream velocity is more difficult to calculate because

of the recirculation zone therein, for this reason some, in the study of cavity flow, have opted

against the calculation of the momentum thickness, favoring rather vorticity thickness as a

means of quantifying the shear layer growth [183]. The vorticity thickness is defined in this

case as | 𝑢 ⁄ | ⁄ [157].

The instabilities are disturbed by the oncoming wall, some of the main flow is

catapulted into the branch line with sufficient –uj velocity as to generate a shear layer in the

branch line, perpendicular to the layer previously discussed. Without traditional boundary

conditions, this shear layer is very difficult to characterize. Nonetheless, a pseudo vorticity

thickness may be calculated. The Ubj is the free stream velocity of the in-flow to the branch

line while the main flow ingression is given a variable free stream velocity based on the

maximum –uj velocity in each line profile through the shear layer, | | such that

| | | 𝑢 ⁄ |

⁄ .

The free shear layer vorticity thickness growth across the mouth of the branch can be

seen in Figure 4-11. The vorticity thickness grows linearly, as is typical of such flows, with a

slope of 0.3. Upon reaching a maximum value of approximately 6.5 mm the thickness

collapses rapidly just prior to the downstream edge of the branch at its intersection with the

main pipe. The second shear layer growing in the branch is much broader than the first, and

although ΔU is smaller than in the first layer, the vorticity thickness, also shown in Figure

Page 87: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

70

4-11, is an order of magnitude smaller, being always less than 0.8 mm,. Furthermore, the

increase in vorticity thickness increases into the branch with a mostly linear trend, the slope

being 0.02.

Figure 4-11. Development of the vorticity thickness of the free shear layer at the mouth of the

branch (left) and within the branch (right) with moving average fits.

At a higher velocity ratios (e.g. Ur > 2000), or in the case of a dead leg, it has been shown the

penetration depth of the main flow extends much farther into the branch resulting in the

absence of scalar mixing in the vicinity of the branch mouth. At the simulated velocity ratio

of 100, however, the shear layer is stable both in the velocity field and scalar field such that

periodic scalar fluctuations are present and may be followed far into the branch. The spectra

of scalar mixing along with the uj velocity field is plotted in Figure 4-12, at least two

dominant frequencies are present in the mixing layer in both fields.

Figure 4-12. Power spectral densities at the center of the mouth of the branch line (see black

dot in insert) showing peaks centered around 51 and 86 Hz (vertical black dotted lines) in

both the scalar and uj velocity fields, computed from an LES simulation.

Page 88: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

71

4.3 Onset of turbulent penetration (Ur < 100)

The onset of turbulent penetration in the branch line is characterized by a single,

coherent zone of mixing in an area surrounding the downstream wall of the branch. In scalar

average and fluctuation profiles from WMS measurements. For the estimation of the

penetration depth, only few axial measuring positions are available with the WMS

instrumentation. At Ur = 28, the penetration depth starts exceeding one pipe diameter, at Ur =

33, two diameters, and at Ur = 100, the tap water flow penetrates beyond four branch line

diameters.

The mixing region in the branch line grows with increasing velocity ratio, 28 to 50,

while the highest main flow concentrations of up to 50% remains fixed to the wall. Near the

depth of farthest penetration, however, this region appears to separate slightly from the

downstream wall (see Figure 4-13, 2Db at Ur = 50). Scalar fluctuations measured in these

experiments, which are of the utmost interest when discussing thermal fatigue, are seen to

vary tremendously in terms of location (e.g. proximity to the conduit walls) and magnitude

throughout the turbulent penetration region, see Figure 4-13. Initially, at the lowest velocity

ratios for which turbulent penetration extends to the neighborhood of 1Db (i.e. to the first

available measuring position of the WMS), the region of highest scalar average and RMS

overlap near the right wall. As the velocity ratio increases the penetration along the right wall

becomes steadier, as seen in the higher average scalar, therefore causing the region of highest

scalar RMS to transition to the area in which the average scalar gradient is highest. A

crescent shaped region of high RMS is formed, extending from the top wall to the bottom

wall. Far from the junction, near the extent of the turbulent penetration at the end of the

thermal cycling region, which is extended by increasing velocity ratio, the maximum of the

scalar average and fluctuation profiles are again overlapping and concentrated at the right

side wall. It is in this region, at the deepest WMS measurement cross section, that the RMS

has its highest global values (per velocity ratio).

The behavior, in general, is the result (as verified by LES) of a free shear layer in the

main flow as the flow passes the leading (upstream) edge of the branch line, followed by the

impact of these turbulent eddies, now deviated slightly into the branch line, into the trailing

(downstream) edge of the branch line. A fraction of the main flow is then sheared off and

propelled by its own momentum into the branch line along the wall until all momentum is

lost against the encroaching branch line flow.

Page 89: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

72

Figure 4-13. Contour plots of the normalized average scalar (left) and scalar RMS (right) in

the branch line for three velocity ratio cases up to 50.

Page 90: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

73

4.4 Flow characteristics in the branch line (Ur = 100)

The present section is primarily dedicated to reviewing a time-resolved CFD

simulation27

of turbulent penetration mixing at a velocity ratio of 100 (Rem = 76,394) in the

LKE T-junction, for which there is experimental data from WMSs. The average velocity

field, shown in Figure 4-14 (right) demonstrates the coherence of the recirculation zone

within the branch line, appearing as an oval shape in the branch extending to 3Db. At the

downstream wall of the branch the average fluid velocity is its fastest anywhere in the branch

at up to approximately 0.45 m/s or nearly 30% of the mean main flow velocity. The velocity

field vectors aim directly into the branch and slightly towards the center of the channel,

reducing in magnitude until the 180 degree turn back towards the junction is made at very

low speed. In the upstream half (x < 0) of the branch in the mid-plane the average velocity

vectors are directed in the –y direction with little deviation until near the branch where a

growing +i component exists, completing the oval-shaped recirculation zone. The mean

velocity field and instantaneous velocity field at an arbitrary time is seen in Figure 4-14 (left).

In a snapshot of the velocity field in time it is evident that fluid behavior at the downstream

wall of the junction, in which the mean velocity magnitude is large and its direction is

dominated by the +j component, is not steady. Rather, clear evidence of pulsations in the

form of local high or low +j velocity vectors is present along the wall.

Figure 4-14. Snapshot of an instantaneous velocity field (left) and the mean velocity field

(right) with a cutoff of 0.6 m/s in the z = 0 mid-plane of the T-junction. The highest velocities

are found in the shear layer at the branch entrance. The snapshot indicates pulsations along

the downstream wall of the branch slowing and forming a coherent recirculation zone in the

average map.

Past research by EdF in a larger geometry with an order of magnitude higher main

flow Reynolds number saw significant swirling in the branch [105-107]. By analyzing the

velocity data, for example mean of uz velocity in the mid-plane (Figure 4-14 left), it is

27

LES setup and methodology described in Section 1.12.

Page 91: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

74

possible to see the first signs of weak swirling in the branch line at Ur = 100. Additionally,

the mean velocity along the axis of the branch line, uj, also shown in Figure 4-15, indicates

that the main flow moves from the downstream wall of the branch to the upstream wall not

through the mid-plane of the branch (i.e. swirling transports it from one wall to the other).

Figure 4-15. Mean uk velocity (left), mean uj velocity toward the T-junction (uj < 0, middle)

and away from the T-junction (uj > 0, right) in the branch line indicating weak swirling in the

turbulent penetration.

The average scalar as viewed in the mid-plane (z = 0) of the T-junction indicates a

broad region of mixing with the highest average scalar in the branch located against the

downstream wall. The scalar maintains a value above 0.6 along the wall a number of

diameters into the branch. Mixing reduces the average scalar deeper the branch and

eventually the main flow reaches its farthest average depth in the branch on the upstream

wall, see Figure 4-16 (middle). A snapshot of the scalar indicates coherent instabilities

forming after the separation of the main flow at the upstream edge of the branch. These

instabilities are sheared by the downstream wall of the branch and appear to enter the branch

very close to the wall. At a greater depth the partially mixed main flow appears in the branch

as plumes of varying scalar concentration. For example, in Figure 4-16 (left) a large plume of

scalar approximately 0.7 is seen, and at a location farther into the branch another plume is

seen with a scalar value closer to 0.2. Finally, the scalar RMS in the mid-plane, shown in

Figure 4-16 (right), indicates a concentrated region of high RMS near the downstream wall of

the branch centered at 3Db. Another area of elevated RMS is observed in the shear layer

between the main flow and mouth of the branch.

Page 92: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

75

Figure 4-16. The instantaneous scalar (left), time-averaged scalar (middle) and the RMS of

the scalar (right) in the mid-plane of the T-junction.

A cross-sectional view, as viewed looking towards the junction, of the mixing in the

branch is presented in Figure 4-17 for a series of non-dimensional positions from 1Db to 4Db

steps of 1Db, in. At 1Db the center of the branch cross section contains an elevated average

scalar and RMS, also visible in Figure 4-16 (a) as the growing shear layer. A high scalar

average and low RMS is present in a thin layer along nearly the entire circumference of the

downstream side of the branch. At 2Db the scalar remains elevated near the wall, decreasing

steadily from downstream wall across the pipe to the upstream wall. By 3Db the average

scalar is almost uniform through the cross section while the scalar RMS indicates a region of

high RMS on the downstream half of the pipe in the region of thermal cycling. Finally, the

average scalar is nearing zero throughout most of the cross-section at 4Db and some diluted

main flow resides in the upper corner of the upstream wall. Although the scalar has a

relatively low average value the RMS is still elevated at this depth, near the farthest extent of

the penetration.

Page 93: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

76

Figure 4-17. Average scalar (top row) and scalar RMS (bottom rom) for various non-

dimensional positions in the branch, from LES.

Thermal cycling in the branch line of a T-junction is the result of a non-steady

turbulent penetration depth along the branch axis. Thermal cycling occurs near the extent of

the average turbulent penetration where the highest scalar RMS values are observed. The

thermal cycling region is isolated as 90% of the maximum scalar RMS in the branch line, in

Figure 4-18. The region of thermal cycling is a contiguous saddle shape which is nearly

symmetric across the z = 0 mid-plane and located against the downstream wall of the branch.

The proximity of the region of highest scalar RMS to the wall of the conduit is important to

recognize as a point of concern for the conduit integrity.

Figure 4-18. Region of > 0.9 max

RMS in the branch line in the z = 0 plane (left) and looking

towards the T-junction (right). The region found to be contiguous and lies along the

downstream wall of the branch in the neighborhood of y = 3Db.

The uniformity index (UI) is typically accredited to Weltens (1993), and is calculated

as follows,

Page 94: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

77

∑ | |

| | ∑ , (4-4)

where θ is the scalar quantity, in this case the scalar average or RMS, θf is the cell face value,

and Af is the area of the face [184]. A uniform scalar distribution would have a UI of unity.

The index helps to characterize the recirculation zone in terms of degree of average mixing

within the pipe cross section. In this way zones in which the turbulent mixing is poorly

distributed or incomplete are identified. The UI was computed based on cross sectional data

in steps Δy = 1 mm though the entire branch.

Scalar fluctuations at the outer skin of the mesh can be interpreted as the boundary

condition for the pipe inner wall since only the fluid in the pipe has been simulated. While it

is important not to lose sight of the physical reality: the strong damping effect of conjugate

heat transfer, between a steel conduit and the fluid, on the fluctuations compared to the

adiabatic simulation herein. Note, however, that the location of peak RMS at the inner wall in

cross-flow mixing studies have been shown to be the same in both cases (in the presence or

absence of conjugate heat transfer) [82]. Therefore, in an effort to better evaluate the

locations at which the greatest non-uniform near-wall mixing occurs a weighted UI was

calculated for sections of the wall. The branch line wall has been divided into circumferential

bands 1 mm in axial length. The UI of the scalar RMS is calculated over each band

individually. As a means of giving greater weight to those bands with high UI and high RMS

a weighting factor w, which is equal to the maximum of the scalar RMS in the region of

interest is multiplied by the UI such that, , see Figure 4-19.

Figure 4-19. Uniformity Index (UI) of the scalar average in cross-sectional slices and the

weighted UI (UIw) of the scalar RMS over 1 mm thick circumferential sections of the branch.

The PSD is invaluable in interpreting turbulent mixing in the context of the thermal

fatigue problem. It is possible that the scalar fluctuations could be threatening at first glance,

i.e. of large amplitude, but have a frequency for which NPP piping steels are impervious.

The PSD has been calculated as follows,

*∫

+

(4-5)

Page 95: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

78

where θ(t) is the signal from each crossing point and tend is the duration of the signal. The

duration and frequency of the WMS measurements allow for more confidence in PSD

calculations due to the ability to ensemble average the signal while maintaining the most

important frequencies (roughly 0.1 Hz and higher). Zero padding of the PSD is avoided by

cutting additional signal collected beyond the nearest 2n data points. For example in the case

of 170 second measurements at 1600 Hz, less than 7 seconds are cut from the end of the

measurements for the purpose of computing the PSD with exactly 218

data points such that

there is no interpolation in the frequency domain.

The normalized PSD in the branch line at crossing points with high RMS at Ur = 50,

for example, at various depths in the branch line, all show a similar tendency towards high

amplitude at low frequencies, see Figure 4-20. Notable is the damped amplitude of high

frequency fluctuations at the deepest WMS position at this velocity ratio, 2Db. The same

phenomena is present in the higher velocity ratio case, Ur = 200; there it can be seen that near

the extent of the turbulent penetration, at 4Db in the branch line, high frequency fluctuations

are an order of magnitude weaker in the PSD.

Figure 4-20. PSD at WMS crossing points (t,r) of high RMS for various branch line locations

at a velocity ratio of 50.

Calculations discussed and shown previously in Section 1.1 and in Figure 4-12,

instabilities forming in the shear layer at the precipice of the branch line that are known to

contain dominant frequencies in the velocity and scalar spectra. As a means of tracing the

lineage of these characteristic frequencies as the shear layer grows and the instabilities are

redirected into the branch line, the PSD of the scalar field has been calculated (based on the

LES simulation at Ur = 100) along an L-shaped path, see Figure 4-21.28

The frequency domain is non-dimensionalized by introducing the Strouhal number,

calculated based on the branch line diameter and main flow velocity , where

Um = 1.53 m/s. Ensemble averaging of the signal spectra helps in discerning peaks and other

28

Details of the L-shaped path are as follows: with a step size of Δx = 1 mm, the PSD is computed in the z = 0

plane from (x,y) coordinate (0.012, 0.026) (1 mm past the upstream wall of the branch line and 1 mm into the

branch) to (-0.012, 0.026) (1 mm away from the downstream wall of the branch line and 1 mm into the branch).

At this point the stepping is made into the branch such that Δy = 1 mm starting from (-0.0128, 0.026) ending at

(-0.0128, 0.091), or y = 3.5Db.

Page 96: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

79

regions of interest; for this purpose Bartlett's method has been executed in generating the

following PSD [185]. The full length signal is cut in to K-segments of duration tk, with 0%

overlap and without windowing,

∑ . (4-6)

The PSD is then computed for each segment of duration tk and all K spectra are then averaged

to create the final, ensemble averaged spectrum,

*∫

+

, (4-7)

⟨ ⟩

. (4-8)

It is furthermore observed that there is a somewhat dominant frequency of scalar mixing at

the extent of turbulent penetration in the branch line, in the region of highest scalar RMS.

The normalized scalar PSD along the path of turbulent penetration into the branch,

shown in Figure 4-21, indicates a drastic weakening of the initially dominant peaks, centered

around St = 0.95 and 1.38 (55.9 Hz and 81.2 Hz), in the region of instabilities across the

branch mouth. Retardation of the main flow and mixing within the branch brings the scalar

fluctuation frequencies into the range relevant to thermal fatigue, St = 0.0017 – 0.17 (0.1 to

10 Hz), by the time the upstream progress of main flow in the branch fully stalls. Unluckily,

this is also the location of highest scalar RMS; the thermal cycling region.

Figure 4-21. The normalized, ensemble averaged (K = 4) scalar PSD vs. St number calculated

from LES results plotted a series of locations along a path (shown in sub-figure) from the

separation point of the main flow, across the mouth of the branch line and into the branch

along the downstream wall to a depth of 3.5Db. The darkest plotted line color indicates the

first path location while the lightest indicates the last in the branch.

Page 97: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

80

There is a sweeping transition from high frequency scalar mixing in the turbulent

main pipe to low frequency fluctuations in the turbulent penetration plume in the branch. The

fractional contribution to the total RMS of three St bins is shown in Figure 4-22. As the shear

layer induces scalar mixing upon interaction with the branch fluid, between 70 and 80% of

the scalar RMS is the result of high frequency (St > 0.8) scalar fluctuations. The mid

frequency, 0.2 < St < 0.8 (11.7 to 47 Hz) component of the PSD grows steadily in influence

at the expense of the high frequency fluctuations, such that half of the scalar RMS is due to

fluctuations in this frequency rage near the junction edge at the downstream wall of the

branch. The trend upon moving from the start of the branch to the end of the path at a depth

of 3.5Db indicates a steady transition of power from the high and mid frequencies into the low

frequencies. Nearly 100% of the total scalar variance near the downstream wall of the branch

at 3.5Db is the result of low frequency scalar fluctuations. Furthermore, the scalar RMS,

normalized to the maximum along the path reaches its maximum when fluctuations are

completely within the low St bin. As the last of the scalar spectrum power exits the high and

middle frequency range in the branch line, the RMS increases rapidly to its largest value

before decreasing; the scalar RMS remains the result of low frequency mixing.

Figure 4-22. Fractional contribution to the scalar variance of fluctuations from three Strouhal

bins, high (St > 0.8), medium (0.2 < St < 0.8), and low (St < 0.2) and normalized scalar RMS

at each position along the L-shaped path, calculated from LES results.

For the purpose of comparison the LES results were interpolated onto a 16-by-16 grid

in the branch cross-section identical to that of the experimental data from the WMS. The LES

results are found to systematically under predict the average scalar across the branch at 2Db,

as is shown in Figure 4-23. Both experiment and LES show a weak gradient in the average

scalar from the left side of the branch to the right. The LES result accurately predicts the

maximum RMS value in the cross section, however, the location and trend across the branch

are both under predicted and over predicted depending on the location.

Page 98: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

81

Figure 4-23. Plots of the average scalar (left) and RMS of the scalar (right) across the middle

of the branch pipe at 2Db at each WMS crossing point location from experiment and LES

simulation.

The PSD near the downstream wall of the branch line at 2Db measured by the WMS

compares well with the findings of the LES simulation. While the magnitude of the RMS is

not identical, when the spectra are plotted on top of each other, as in Figure 4-24, both

experiment and simulation anticipate a broad peak or plateau in the spectra around 6-7 Hz at

the location of crossing point (8,15). The LES simulation shows less power in the high

frequencies due to the filtering performed by the governing equations, whereas the WMS

may measure additional fluctuations at high frequencies due to electronic noise. A similar

agreement was also found in other locations.

Figure 4-24. Ensemble averaged PSD at a position near the wall (shown on the right) in the

branch line at a depth of 2Db from WMS measurements (K = 128) and LES simulation

(K = 16).

Page 99: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

82

4.5 Flow characteristics in the branch line (100 < Ur < 3000)

Turbulent penetration at higher velocity ratios, Ur = 100 to 400, appears to be similar

in behavior to that at lower velocity ratios with the exception of the final, deepest

measurement cross section, see Figure 4-25. At 2Db a continuation of the trend seen at lower

velocity ratios, in which the main flow average concentration is highest at the downstream

wall, is present. In this case, however, the concentrations of tap water are significantly higher

at the same axial position, between 60 and 80% across the entire cross section for the Ur =

200 case, for example. The scalar RMS at 2Db shows the typical crescent shape around the

region of highest average.

By 4Db a breakdown of the scalar average symmetry about the pipe mid-plane has

occurred and the main flow appears in highest average concentration along the bottom of the

pipe rather than remaining vertically centered near the right side wall as for lower velocity

ratios. This shift results also in a disturbance of regularity in the RMS profiles. What can be

noted, however, is the elevated levels of the RMS, to nearly 0.35 across a sizeable portion of

the top wall of the branch line for a velocity ratio of 400. The sudden break in symmetry here

is believed to be attributable to two factors. The first is swirling of the flow and the second is

related to a weak buoyancy due to the different working fluids. Petrov (2015) has shown

evidence of buoyancy effects in mixing between deionized and tap water [186]. Here the

buoyancy may also be the result of small temperature differences between the main and

branch flows.

Figure 4-25. Normalized average scalar (left) and RMS (right) in a 50 mm diameter branch

(Dr = 1) for three velocity ratio cases up to 400, and two WMS sensor positions.

Page 100: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

83

Figure 4-26. Scalar average and RMS profiles measured by a WMS at three branch positions

during mixing at Ur = 2000 in a 26 mm branch and rounded edge T-junction.

Figure 4-26 shows results from WMS measurements at a velocity ratio of 2000 at

three branch depths. Interpretable is strong evidence of swirling. Rather than the highest

concentration appearing at the downstream wall of the branch line, the scalar is concentrated

in a very similar way at the upstream wall, at 3.73Db. Deep in the branch at 7.73Db the

average scalar is very low at the bottom of the branch but RMS values remain high,

indicating transient turbulent penetration at this depth. The swirl direction may be due to

small density difference between the tap water and deionized water (as discussed above),

caused by slight inclinations in the section, or a real effect independent of the idiosyncrasies

of the experimental setup and based solely on the geometry and Reynolds number29

[105].

Additional experiments been carried out showing the manner in which increased

turbulent penetration is with increasing Reynolds number for a fixed velocity ratio (Figure

4-27) and vice versa for a fixed Reynolds number with increasing velocity ratio (Figure

4-28). In these WMS measurements, shown at 5.73Db the average scalar profile appears to be

influenced, in part, by the laminar velocity profile of the incoming deionized water flow

reducing the average scalar in the center of the conduit.

29

Robert (1992) notes the example of non-symmetrical flow separation in certain symmetric geometries as a

potential cause of turbulent penetration in the branch line.

Page 101: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

84

Figure 4-27. Scalar profiles measured by a WMS at 5.73Db indicates increased turbulent

penetration with main flow Reynolds number for a fixed velocity ratio.

Figure 4-28. Scalar profiles measured by a WMS at 5.73Db indicates increased turbulent

penetration with increasing velocity ratio for a fixed main flow Reynolds number.

Page 102: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

85

4.6 Global instabilities introduced by pump oscillations30

4.6.1 Influence of pulsation amplitude

When comparing the behavior of turbulent penetration for a main volume flow rate

oscillation amplitude, a, ranging from ±0% to ±20% at 1.5Db (see Figure 4-29), a few

observations are apparent. The average of the scalar, while maintaining the tendency of

penetration at the right wall of the branch line, tends to increase in magnitude with increasing

amplitude of the velocity oscillations. The RMS of the scalar transitions from a profile

indicative of a steady turbulent penetration (Qm ± 0 to 5%) to a much higher RMS at the wall

indicating that the region of thermal cycling encompassed a region which includes 1.5Db (Qm

± 10 to 20%). In general, in the branch line the average scalar tends to decrease while the

RMS increases with increasing a.

The penetration depth, which has been shown in this work, among others, to be an

increasing function of the velocity ratio, cycles between larger and larger ranges of depth in

the branch line with increasing main flow velocity oscillation amplitudes which results in the

higher scalar RMS values [141].

Figure 4-29. Contour plots at 1Db of the normalized time-averaged scalar and scalar RMS

measured by a WMS for the steady case of Ur = 34.5 and for Qm oscillations ranging in

amplitude from ±5% to ±20% at 0.1 Hz frequency.

4.6.2 Influence of pulsation frequency

Wire-mesh sensor data recorded for main flow oscillation frequency from 0.1 Hz to

0.3 Hz (see Figure 4-30) predicts trends in the scalar and RMS as function of f. The

corresponding Strouhaul (St) and Womersley (α) numbers are reported in the test matrix in

30

Recall the experimental procedure for these tests, described in Section 3.1.2. All tests are carried out in a

sharp edged T-junction with equal branch ratio, Dm = Db = 50 mm.

Page 103: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

86

Table 3-3. Within this range of Womersley numbers there is no indication of a change in

regime in terms of a mixing pattern. The measurements indicate that the average scalar

converges on the steady (f = 0) Qm profile from below with increasing Womersley number

(corresponding to increasing pulsation frequency). The scalar maintains the same profile in

the branch line but with a decreasing amplitude for decreasing Qm pulsation frequencies. The

opposite trend appears in the scalar RMS profiles in which it is evident that the RMS is

elevated for any Qm oscillations and tends to decrease towards the steady-Qm case from above

(i.e. from higher RMS values shrinking down to the stead- Qm profile) as the oscillation

frequency grows.

Figure 4-30. Contour plots at 1.5Db of the normalized time-averaged scalar and scalar RMS

for the steady case of Ur = 34.5 and for Qm oscillations ranging in frequency from 0.1 to 0.3

Hz with ±10% amplitude.

Measurements show that given the inlet condition of a pulsatile main flow, a sizeable

fraction of the total variance present in the scalar mixing PSD is concentrated at the same

frequency f, as the main flow velocity oscillations, especially in regions of high scalar RMS.

A peak in the mixing spectrum is present for all test cases, while for some it is more

prominent than others. For example, the peak magnitude in the PSD grows considerably

between the cases Qm ± 5% to 20%, regardless of the frequency. It is interesting to visualize

the change in peak magnitude for regions of high RMS and low RMS. Figure 4-31 shows a

ensemble averaged (K = 4), normalized PSD of a crossing point signal near the downstream

wall in the branch line, in the region of high RMS, and near the upstream wall, in a region of

low RMS, for the case of a steady main flow and oscillating main flow at f = 0.1, 0.2 and 0.3

Hz. The PSD shows peaks at the respective main flow pulsation frequencies. Hence a

transition between the velocity domain in the main flow (input) and scalar domain in

turbulent penetration mixing (output) takes place in the branch line. A weak dampening of

the peak amplitude in the PSD from the 0.1 Hz to 0.3 Hz case is apparent. When comparing

the spectrum from the region of high RMS (Figure 4-31 right) and low RMS (Figure 4-31

left) for the same flow conditions, a significant dampening of the peaks is seen. That is, in

Page 104: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

87

regions of high scalar RMS, a major portion of the scalar RMS magnitude is the result of

scalar fluctuations at the Qm oscillation frequency while in regions of low scalar RMS a

smaller fraction of the RMS is derived from fluctuations at f Hz.

Figure 4-31. Normalized PSD at a crossing point near the upstream wall (left) near the

downstream wall (right) of the branch line from the WMS measurements in which the main

flow inlet oscillation frequency was varied. Flow condition, Ur = 34.5 Qm ± 10% at 1.5Db.

A plot of the peak amplitude in from an ensemble averaged (K = 16), normalized PSD

(Ur = 34.5, Qm ± 10%, f = 0.3 Hz, at 1.5Db) at every crossing point horizontally across the

branch line from left to right helps to show the steady increase of the peak magnitude with

increasing RMS, see Figure 4-32. This indicates the power delivered to the main flow

oscillation frequency is amplified in regions of higher RMS in the turbulent penetration

region, akin to a resonance, since the normalization of the PSD renders the spectrum

independent of the RMS magnitude. Also notable is the presence of a second harmonic peak,

at 0.6 Hz, appearing about half way across the pipe as the peak at 0.3 Hz, and the scalar RMS

both grow larger.

Figure 4-32. Magnitude of the ensemble averaged (K = 16) PSD (from the WMS

measurement plane at 1.5Db) at 0.3 Hz, 0.6 Hz and 0.9 Hz (left axis), along with the scalar

RMS (right axis), for each crossing point across the middle of the branch line, from (16,1) to

(16,32). Flow condition Ur = 34.5, Qm ± 10%, f = 0.3 Hz.

Page 105: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

88

4.7 Influence of T-junction edge geometry

For the case of a rounded edge at the intersection of the conduits at the T-junction the

main flow boundary layer separation point upon reaching the branch line is no longer

necessarily located at a fixed point in space31

as in the sharp edge geometry (i.e. at the edge).

When a sharp edged junction is in place at high velocity ratios the shear layer forming after

the separation point at the upstream wall of the branch has been shown in time resolved CFD

simulations to result in large scale instabilities of with characteristic frequencies which are

then sheared and injected into the branch line along the downstream wall (see Section 1.1)

[151]. The flow behavior at a T-junction with curved edge is less straightforward.

Four WMS measurements are presented from two different branch positions in a

sharp edged T-junction and at the same two locations in a T-junction with a curved edge of

radius 10 mm, see Figure 4-33. At 3.73Db the average scalar is high (θ > 0.85) and the RMS

low (θRMS < 0.1) in the entire cross-section. Little if any difference is detected between the

mixing in the sharp edged and curved edge T-junction at this position. Upon closer

inspection, however, there is a slightly higher scalar concentration in a crescent shape along

the left side of the branch. At the following WMS position, at 5.73Db the symmetry across

the midline has been broken and a higher concentration of tap water is found more often in

the bottom of the branch.

Figure 4-33. The scalar average (left) and RMS (right) across the branch cross-section at two

positions in each T-junction geometry, measured by a WMS. Flow condition, Ur = 2000, Um

= 3.92 m/s, Rem = 6.5E4.

At 5.73Db the highest RMS values appear between the wall and the center of the pipe

with values of 0.2 to 0.25+. When the T-junction is rounded the same behavior is exhibited at

both measurement positions. In the case of a rounded edge, however, significantly less

turbulent penetration is detected, along with larger scalar fluctuations. The PSD has been

calculated at crossing points of high RMS in both T-junction geometries and WMS branch

positions; see Figure 4-34. Closer to the T-junction high frequencies dominate the spectra and

little effect is seen from curved junction edge. Most of the scalar fluctuations at 5.73Db are

31

The separation point on the curved wall is due to an adverse pressure gradient (strong flow deceleration) in

the boundary layer.

Page 106: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

89

occurring at low frequencies for both T-junction geometries. At the same position, two peaks

in the spectrum are visible at this crossing point only in the case of the fileted edge, at

approximately 17 and 23 Hz.

Figure 4-34. Normalized PSD

32 at crossing points of high RMS at 3.73Db (10,10) and 5.73Db

(5,14) in a sharp edge and curved edge (r = 10 mm) T-junction, Ur = 2000.

32

In this case estimated by Welch‘s method using no overlapping and a Hann window.

Page 107: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

90

4.8 Influence of density stratification

Unlike turbulent penetration with iso-dense fluids in the branch line and main pipe,

where large time-averaged scalar gradients are seen only deep in the branch line, density

differences (between main and branch fluids) results in the branch flow being present,

unmixed or weakly mixed in high concentrations much closer to the T-junction. Furthermore,

instead of reaching a maximum near the penetration depth, scalar fluctuations may exhibit

high values across a wide range of axial distances, the result of flow stratification. Note that

the Froude number based on the branch line velocity is always subcritical in the following

experiments.

Figure 4-35. Schematic of the Farley-Tihange phenomenon as observed at the LKE T-

junction facility.

Time histories of these experiments provide evidence of the Farley-Tihange

phenomena (see Figure 4-35), phenomena responsible for through-wall cracks in US and

European NPPs, previously described in Section 1.2.4. The heavy branch flow may be forced

to recede by the turbulent penetration at regular intervals such that at the bottom of the

conduit cyclic thermal shock, large amplitude variations in the scalar, occurs. Figure 4-36

demonstrates a scalar time series measured at the bottom of the branch line in which regular,

large scale pulsations are present. These pulsations are due to the heavy branch flow

occasionally reaching forward to the WMS position (3.73Db). Furthermore, the PSD in the

cross-section of the pipe, Figure 4-37 indicates that the pulsations at the bottom of the conduit

have a dominant frequency of 0.1 Hz. This mixing behavior represents a thermal fatigue

threat as demonstrated in the Farley and Tihange coolant leak accidents.

Figure 4-36. Signal time series from a crossing point near the bottom of the branch line in the

case of turbulent penetration with weak, dense in-flow (right). Present are large amplitude

fluctuations (highlighted in red). Data from a WMS measurement in a sharp edged junction at

3.73Db, ε = 0.059.

Page 108: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

91

Figure 4-37. Normalized PSD at every crossing point from the bottom (blue) to the top of the

branch (orange). Calculated from a WMS measurement in a sharp edged junction at 3.73Db,

ε = 0.059.

Figure 4-38 shows the average scalar as measured by the WMS for density

differences ranging from 0.024 to 0.088 at positions 3.73Db and 5.73Db. Firstly, at the nearer

position to the junction increasing density stratification introduces a significantly lower

average scalar concentration at the bottom of the branch as compared to iso-dense mixing.

For weak stratification, the case of mixing with ε = 0.024, the average scalar may reach

values approaching 0.5 at the bottom of the pipe. As the branch fluid density is increased, a

larger zone of low average scalar forms in the bottom of the branch and eventually, at ε =

0.088, a zone of average scalar between 0.05 and 0.15 is found in the bottom third of the

branch. The density stratification has introduced areas of large average scalar gradients, as

high as 0.09 θ/mm at the interface between the fluids. In iso-dense mixing cases the average

scalar is almost uniform across the entire branch line cross section. At 5.73Db we see sharply

stratified flow in the cross-section leading to very strong average scalar gradients of nearly

0.12 θ/mm. The exception is for the small density difference of 0.024 where there is still

noticeable blurring of the average scalar across the interface between the fluids.

Figure 4-38. The scalar average is plotted across the branch cross-section at two positions for

each density difference.

Page 109: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

92

Figure 4-39. The RMS of the scalar is plotted across the branch cross-section at two positions

for each density difference.

The RMS of the scalar indicates negligible scalar fluctuations at 3.73Db when no

density difference between the fluids is present. As the branch line fluid density is increased

the scalar RMS is found to be as high as 0.3 at the bottom of the pipe at the turbulent

interface between the two fluids as seen in Figure 4-39. Already for the case of ε = 0.051 the

interface exists independent of the bottom wall of the branch; deionized water is running

more or less steadily underneath the penetrating tap water. The trend is as follows: the scalar

RMS is reduced slightly at the bottom of the branch, where the average scalar is near 0 and

the maximum of the RMS is found rather just above the bottom, again at the interface

between the two fluids, and finally at the edges of the interface near the left and right walls at

ε = 0.088.

Deeper in the branch line (5.73Db) where in iso-dense mixing experiments elevated

RMS values were present throughout the cross-section, the introduction of even small density

differences again has a drastic effect. For small Δρ between the working fluids, ε = 0.024, the

mixing is strongly damped and reduced in spatial extent and although the maximum of the

RMS is still sizeable at 0.25 it exists at such elevated values only at a few crossing points on

one side of the interface between the fluids. In the remainder of the cross-section RMS values

are near zero. With increasing density difference the scalar RMS at 5.73Db is elevated in only

a single row of crossing points across the branch, i.e. ≤ 1.63 mm in vertical expanse. In this

row RMS values may reach above 0.15. Although scalar fluctuations are negligible above

and below the mixing layer, it should be noted that an order of magnitude difference is seen

between the RMS in the penetrating main flow on top of the interface and the heavy branch

flow below. For example, above the interface RMS values are typically in the range 0.02 to

0.05 while below the interface they are below 0.005.

Figure 4-40 shows the development of the stratified scalar interface with increasing

branch fluid density. The curves are the result of averaging the middle two crossing point

columns of the WMS. It is evident that even weak density stratification strongly affects the

behavior of turbulent penetration in the horizontal T-junction when compared to flows with

ρr = 1, especially deeper in the branch. At that position we see already at a density difference

of 0.051 the average scalar interface is smaller than the spatial resolution of the WMS, hence

the overlapping profiles. At 3.73Db an S-curved scalar profile is only present for the largest

density difference, 0.088.

Page 110: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

93

Figure 4-40. Vertical scalar average profiles (through the center of the WMS measurement

plane, two column-average) indicating flow stratification for each experiment at 3.73Db (left)

and 5.73Db (right).

An increasing density difference causes an increase in the maximum average scalar

gradient vertically through the center of the cross-section at 5.73Db, see Figure 4-41. The

gradient, ⁄ approaches 0.08 θ/mm at the highest density differences (Ri ≈ 5800)

while the largest change in ⁄ is observed between Ri ≈ 1550 and 3350 after which

increasing Ri has a lesser effect on the scalar gradient. Near the branch at 3.73Db, the trend as

a function of Richardson number is inconclusive based on the four density difference cases.

What is clear is a reduction in thermal gradient with increase main flow Reynolds number,

owing to an increase in turbulent penetration and therefore turbulent kinetic energy which

acts to disrupt the formation of stratification between main and branch flows, thereby

reducing the scalar gradient in the time-averaged sense.

Figure 4-41. Maximum average scalar gradient ⁄ across the center of the pipe

cross-section (along the z-axis) vs. Richardson number with respect to branch line position

and main flow Reynolds number.

A density difference between the fluids gives rise to a power spectrum with

significantly more amplitude in the low frequencies near the T-junction where RMS values

are also significantly higher. At 5.73Db where mixing is relegated primarily to the narrow,

Page 111: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

94

wavy interface between the fluids, some preferred mixing frequencies arise, described below.

Figure 4-42 shows the PSD at the crossing point of highest RMS for each experiment; when a

density difference is present, this point always lies somewhere in the interface between the

fluids.

Figure 4-42. Ensemble averaged PSD at crossing points of highest scalar RMS, wherever

they may be in the cross-section, for each experiment at 3.73Db and 5.73Db.

Two areas of interest deserve mention regarding the spectrum at 5.73Db. Firstly, there

is a large plateau around 6-9 Hz, which does not exist in the absence of density difference.

This plateau precipitates a steepening of the slope of the spectra for all four artificial density

cases. At higher frequencies, there is a peak in the PSD for all experiments in the vicinity of

28 Hz which is more pronounced with higher density difference. In the absence of density

difference the spectrum shows no major changes in slope at this position.

Page 112: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

95

4.9 Conclusion

Experiments have been carried out in a new adiabatic T-junction facility at velocity

ratios up to 3000 to better understand the behavior of turbulent penetration in T-junctions

with in-flow in a nuclear-safety context as it relates to HCTF. The main findings and

takeaways include,

Database of adiabatic experimental results with high spatial and temporal resolution

downstream of the T-junction and in the branch line for a wide range of high velocity

ratio, turbulent-penetration-inducing boundary conditions.

Comprehensive description of turbulent penetration with in-flow including elucidation

of the Farley-Tihange phenomena, the effect of density stratification, velocity ratio,

pulsatile flow and T-junction geometry on turbulent mixing by means of adiabatic

experiments and a large eddy simulation.

Region of largest scalar fluctuations is concentrated in a wall-attached zone in the

case of iso-dense turbulent penetration in T-junctions at velocity ratios around 100.

Shear layer formed across the mouth of the branch line due to separation of the main

flow plays an important role in turbulent penetration mixing, especially with regards

to critical mixing frequencies.

Recognition of characteristic mixing frequencies related to the Farley-Tihange

phenomena in T-junction branch lines with in-flow.

Portions of the experimental and theoretical work discussed in this chapter has been

or is expected to be published in peer-reviewed journal publications, peer-reviewed

conference proceedings, or in the theses of Master‘s students supervised by myself at the

ETH Zurich LKE. Specifically,

1. Master‘s thesis of Valori (2012): ―T-junction mixing with low side flows‖ [187]

2. Master‘s thesis of Schmidt (2013): ―Experimental and Numerical Investigation of

Turbulent Penetration Mixing in a Low Side Branch Flow Rate‖ [188]

3. Master‘s thesis of Trinca (2014): ―Thermal Loads Determination For Fatigue

Assessment In Low Side Flow T-junctions‖ [189]

4. Kickhofel, J., Valori, V., and Prasser, H.-M., ―Turbulent Penetration as a Thermal

Fatigue Problem in low Side Flow T-Junctions,‖ Nuclear Thermal Hydraulics and

Safety 8 (NTHAS8), Beppu, Japan, December 9-12, 2012.

5. Kickhofel, J. and Prasser, H.-M., ‖Large Eddy Simulation of Turbulent Penetration in

a T-junction,‖ International Congress on Advances in Nuclear Power Plants (ICAPP

2014), Charlotte, U.S.A., April 6-9, 2014.

6. Kickhofel, J., Trinca, C., and Prasser, H.-M., ‖The Influence of Density Stratification

and T-junction Geometry on Turbulent Penetration,‖ International Topical Meeting on

Page 113: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

96

Nuclear Thermal Hydraulics, Operation and Safety (NUTHOS-10), Okinawa, Japan,

December 14-18, 2014.

7. Kickhofel, J. and Prasser, H.-M, ‖Turbulent Penetration in T-junction Branch Lines

with Leakage Flow,‖ Nuclear Engineering and Design 276, 43–53, 2014; doi.org/t4g.

8. Kickhofel, J. and Prasser, H.-M., ―Large Eddy Simulation of Turbulent Penetration in

a Horizontal Adiabatic T-junction with Leakage Flow,‖ Nuclear Engineering and

Design, in review, 2015.

Page 114: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

97

5 Mesh sensor package

5.1 Introduction

Experimental measurements of single and two-phase flows at high temperatures and

pressures are invaluable sources of information for theoreticians, researchers and

professionals in the field of nuclear engineering. A great deal of research in the nuclear

thermal hydraulic domain is carried out at low temperature or in adiabatic test facilities at

atmospheric pressure, such as the turbulent penetration studies described in Chapter 4. The

transition from data gathered from such tests to a clearer understanding of thermal hydraulic

or fluid dynamic behavior at the working conditions of an NPP, known as scaling, is

nontrivial. The challenges of scaling could be avoided by the capturing of more experimental

data, of a higher density in space and time than is traditionally possible, from existing test

facilities (e.g. integral-type33

) which are constructed to operate nearer to or at the full

temperature and pressure conditions found in NPPs [190-192]. This is achievable only with

innovative instrumentation.

To this end, it is the intention of this chapter to introduce a new high-temperature

high-pressure resistant mesh sensor34

―package‖ design which is relatively uncomplicated

compared to the current state-of-the-art and whose sealing methodology and materials allow

for potential expansion or modification of the internal sensor structure [193]. The sensor

design potentially greatly expands on the breadth of operating temperatures and pressures

previously possible with mesh sensor technology to encompass not only the thermal-

hydraulic domain of BWRs but also that of PWRs.

The first full-scale prototype of the new mesh sensor has been designed and built for

the measurement of T-junction cross-flow mixing at high temperatures (up to 280°C) and

pressures (8 MPa) at the University of Stuttgart T-junction facility, described along with

results from the new mesh sensor in Chapter 6. The sensor utilizes the same electronic

principle and data acquisition system as traditional WMS technology developed by Prasser

(1998), and implemented at the LKE T-junction facility (Section 2.2.1) [163]. The sensor

design incorporates some classic and other, more innovative engineering materials. The

testing of these materials is described in Section 1.1 followed by the detailed description of

the prototype sensor (Section 1.1) and a discussion of its maximum potential operating

conditions (Section 1.1). The appendix of this chapter explores alternate sensor designs and

suitable alternate materials which were investigated to various degrees during the sensor

research and development process.

33

Examples of which include PWR PACTEL in Finland, PKL in Germany, and ROSA/LSTF in Japan. 34

Since the sensor does not utilize wires, the term is dropped when referring to this sensor.

Page 115: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

98

5.2 The challenge in brief

The engineering effort exhausted in designing a high-temperature-capable mesh

sensor is in large part dedicated to materials considerations. Recall the words of engineering

material guru Michael F. Ashby,

“There has never been an era in which the evolution of materials was faster and the sweep of

their properties more varied. The menu of materials has expanded so rapidly that designers

can be forgiven for not knowing that half of them exist. Yet not-to-know is, for the designer, to

risk failure: innovative design, is enabled by the imaginative exploitation of new or improved

materials” [194].

Indeed a great deal of time has been spent searching for appropriate materials (including the

utilization of Ashby plots) during the development of the mesh sensor package with the goal

of leaving no stone unturned, so as not to miss a potentially game-changing ceramic, alloy or

adhesive, etc. The material attributes of primary importance include some combination of

thermal expansion, electrical resistivity, Young‘s modulus and yield strength, depending on

the particular component. And while Ashby estimates (circa 2009) that, ―[t]here are maybe

more than 50,000 materials available to the engineer,‖ whole classes of those materials are

unavailable at temperatures near or above 300°C [195]. Essentially, only metals, ceramics

and a handful of exotic materials are suitable. That already dwindling list of options is shrunk

considerably when considering the corrosive operating environment: deionized water, vapor

and even boiling. Above 300ºC the use of electrically insulating materials such as PTFE,

PEEK or epoxies is impossible at worst or ill-advised at best. Above 327ºC, flexible,

electrically resistive perfluoroelastomers such as DuPont™ Kalrez® also reach their stated

limits.

Thermal expansion of materials must be accounted for where assemblies comprising a

variety of materials each with different thermal expansion coefficients (TEC) are expected to

encounter elevated temperatures and large thermal gradients. As a first step towards a high-

temperature mesh sensor which may be capable of measuring at temperatures and pressures

approximately equivalent to those found in BWRs, the aim has been to construct a prototype

mesh sensor package for up to 280ºC and 8 MPa. At these ambient conditions, some

thermoplastics, elastomers and epoxies remain usable.

Two primary challenges are faced when designing a mesh sensor for high-temperature

high-pressure flows. First among them, dictating a great deal of the overall design, is the

choice of a suitable electrode material. Due to thermal expansion the stainless steel wires

typically used in mesh sensors would go slack and droop at high temperatures, even if pre-

tensioned to near their yield strain. Such wires do not exhibit the necessary elasticity to

compensate their axial thermal expansion at high temperatures. Truth be told, none of the

dozens of wire materials investigated, be them elemental or alloy, have the necessary

combination of corrosion resistance and elasticity. The option then becomes one of using a

stiff electrode material or stainless steel wires with individual springs for maintaining tension.

The second challenge is sealing the sensor, maintaining the internal pressure of the conduit in

Page 116: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

99

which it is installed while allowing for the passage of signal carriers between the sensor and

the DAQ box. Previously published technology relied on epoxies which sometimes required

active cooling.[196]

Page 117: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

100

5.3 Prior state of the art

Prior state of the art is detailed primarily in publications by Prasser (2008), Dudlik

(2008), Pietruske (2007) as well as in patents and reports originating from the

Forschungszentrum Rossendorf (FZR, now HZDR) in Germany [196-201]. These

publications demonstrate two distinct high-temperature mesh sensor designs, of which both

have relied on high-temperature epoxies, rated to 180ºC, to ensure an effective pressure

barrier. A review of mesh sensor technology for high-temperature high-pressure flows is

shown in Table 5-1.

Table 5-1. Prior state of the art and the new mesh sensor package.

The rod-electrode type design utilized large electrodes, by WMS standards, of 5 mm

in height and 1.5 mm width (with a leaf-shape cross section to reduce pressure drop) which

were fixed on one end and allowed to thermally expand freely, along their axis, on the other.

This design, first implemented in the MTLoop facility at the Forschungs Zentrum

Rossendorf (FZR, now HZDR), went on to be manifest in successful measurement campaigns

at the PPP facility (in DN50 and DN100) of UMSICHT in Germany and the PMK-2 test

facility (in DN80) at the KFKI-AEKI in Hungary as a part of the OECD/NEA Primary

Coolant Loop Test Facility (PKL-2) Project35

[203]. These sensor ranged in size from a

DN100 version with 16 transmitter and 16 receiver electrodes rated for 250ºC at 7 MPa, to a

DN50 8 x 8 electrodes version rated for 180ºC also at 7 MPa. The extension in sensor

temperature resistance from 180ºC to 250ºC was achieved with the help of an integrated

water cooling circuit, outside of the pressure barrier but within the sensor flanges, to ensure

the integrity of the epoxy sealing.

A spring-based mesh sensor design, described by Pietruske (2007), benefits from thin

0.25 or 0.12 mm diameter stainless steel wire electrodes while the overall construction is

admittedly, ―extremely complicated‖ [198]. Thermal expansion and contraction is managed

by a system of compression or tensile springs in concert with ceramic insulation pearls,

requiring extensive machining of the sensor flanges. Two sensor constructions were

manufactured, assembled, and tested in steam-water and air-water multiphase flows at the

TOPFLOW facility at FZR, both with 3 mm wire pitch and 2 mm between the two layers.

35

In the framework of the EURATOM project WAHALoads.

Electrode

material/type

Design

Temperature

[°C]

Tested

Temperature

[°C]

Tested

Pressure

[MPa]

Conduit

size

[DN]

Electrodes

Manera (2001)

[202]

Stainless steel

wires 120 116 .12 40 16-by-16

Pietruske

(2007) [198]

Stainless steel

wires 286 280 6.5

200

50

64-by-64

16-by-16

WAHALoads

Project [196,

197, 199]

Stainless steel

lentil-rods

180

180

250

180

180

250

1

1

4

50

100

80

8-by-8

16-by-16

12-by-12

Present Work

[2]

Stainless steel

capillaries 350 265 7-8 80 16-by-16

Page 118: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

101

The smaller sensor was of 16-by-16 electrodes for a DN50 conduit, the larger, comprised of

64 x 64 electrodes was installed in a DN200 pipeline. The maximum temperature and

pressure ratings for both sensors were declared to be 7 MPa and the corresponding saturation

temperature of 286ºC. Unlike the lentil-rod sensors, these sensors measured flows at their

temperature and pressure limits. The larger of the two sensors prohibited the installation of all

12 bolts in the DN200 flanges, only 8 could be utilized, which meant a special licensing

procedure had to be carried out with the authority responsible for pressurized equipment.

Page 119: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

102

5.4 Material testing

Extensive material testing has been performed in the context of the new mesh sensor

development. A new autoclave (Büchi AG versoclave Typ 3E) has been integrated into the

LKE laboratories for the purpose of material and prototype testing in water/steam

environments at high temperatures and pressures. The autoclave is capable of heating a 3 liter

volume to a maximum temperature of 300ºC at up to 10 MPa, see Figure 5-1.

Figure 5-1. Schematic and photograph of Büchi AG autoclave used for material testing

related to the mesh sensor package development.

For testing, 750 ml of deionized water was placed in the volume along with the

material or assembly to be trialed. Heating then proceeded, typically as in Figure 5-2, such

that as the temperature rises beyond 100ºC nucleate boiling occurs in the water. The

autoclave was then vented briefly to expel air, resulting in a saturated liquid-vapor mixture

which tended towards superheated vapor rather than to water due to the chosen specific

volume. Conditions seen by the tested materials was extreme (i.e. conservative) in that they

were tested in water, steam, and in the presence of boiling. Material tests were performed at

temperatures up to 285ºC at saturation pressure where the maximum temperature was held for

minutes or many hours depending on the test. All materials which went into the prototype

design for which some uncertainty existed regarding suitability for use in high-temperature

water or steam (e.g. ceramics or insulation materials) were tested for at least 1 hour at around

280ºC and saturation pressure (approx. 6.9 MPa).

Page 120: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

103

Figure 5-2. A typical temperature profile measured by a PT100 thermocouple in the autoclave

during a material test.

Page 121: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

104

5.5 Prototype design and in-house testing

The mesh sensor package is notable for a separation of the pressure barrier

construction from the mesh sensor itself. The result is a package design which allows for a

variety of potential sensor designs to be incorporated. Two steel housing flanges comprise the

bulk of the sensor package, they act to provide a cavity within the pressure barrier of the

facility which is large enough that a mesh sensor and associated wiring may be mounted. The

two flanges, the ‗primary‘ and ‗lid‘ flange, surround and secure the mesh sensor sandwich, as

in a sketch of the mesh sensor package in Figure 5-3.

Figure 5-3. CAD drawing of the mesh sensor package installed in DN80 PN100 flanged

pipeline of inner diameter 71.8 mm.

The housing flanges allow for transport of the mesh sensor as a pre-assembled unit

which needs only to be installed on-site between standard flanges in the pipeline. The design

maintains the ability to execute the usual sealing strategy for the flanged pipeline in which

the sensor is installed, in this case that for DN80 PN100 flanges (eight M24 bolts). The

primary housing flange accepts the mesh sensor, centering pins and compression springs. The

lid flange acts as an adapter to seal the sensor inside of the primary housing flange while also

providing the appropriate facing, e.g. tongue and groove, to the pipeline flange. Two 10.7

mm diameter conduits are machined into the primary flange perpendicular to the axis of the

flow conduit and to one another. It is through these conduits that the 16 transmitter and 16

receiver signal carriers (‗transmission wires‘) are passed, separately, from the mesh sensor to

the multi-element feedthrough glands, there passing through the pressure barrier. The wires

are then connected to a D-SUB 25 terminal which is linked to the analog-to-digital DAQ box.

Between the primary flange and the feedthrough glands, stainless steel hoses 20 cm in length

are installed. The purpose of the pressure hoses is to provide greater physical separation

between the sealing glands, which have a temperature rating of 260ºC, and the heat source, be

it the fluid itself or external heating mats. The glands are rated for pressures up to 8.3 MPa.

A consequence of the additional space in the conduit provided by the housing flanges

is that the gasket which ensures sealing between the two flanges is of larger diameter than

Page 122: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

105

those found in the standard pipeline connections. The gasket located between the housing

flanges is of inner/outer diameter (ID/OD) Ø116/140 mm such that the sealing area is similar

to that of the standard Ø90/120 mm gaskets found in a tongue and groove DN80 PN100

flange. This means non-negligible stress in the form of torsion acts on both housing flanges.

CATIA V5 FEM has been performed based on a 300°C ambient with 100 MPa internal

pressure and conservative forces on the gasket surfaces were estimated based on 200 N·m

bolt torque. Results show maximum von Mises stresses of 290 MPa in the primary housing

flange and 350 MPa in the lid flange. These values are well above the Rp0.2 offset yield

strength of standard stainless steels at 300°C. Therefore, the stainless martensitic chromium-

nickel-molybdenum steel 1.4418 was chosen for the sensor package housing flanges,

although alternatives do exists, see Table 5-2. With a yield strength of 580 MPa at 300ºC,

high Cr content, and the addition of Molybdenum, 1.4418 is an ideal candidate for the

requirements of the sensor package in high-temperature steam-water environments. Another

potential material is 17/4 PH, a commercially common martensitic chromium-copper

precipitation hardened steel with high strength and good corrosion resistance [204].

Steel Rp 0.2% at

300°C (MPa)

Heat

Treatment Classification

1.4571 (316Ti)

X6 CrNiMoN 22-5-3 145 Annealed Austenitic

1.4550

X6 CrNiNb 18-10 136 Annealed Austenitic

1.4542 (17/4 PH)

X5 CrNiCuNb 17-4 650 P960 Martensitic

1.4418

X4 CrNiMo 16-5-1 580 QT900 Martensitic

Table 5-2. Comparison of steel yield strengths at elevated temperature. Options for HTWMS

housing flanges include 1.4418 and 1.4542. Yield strength data from Key to Steel (Verlag

Stahlschlüssel) [204].

Figure 5-4. Calculated von Mises stresses on the lid flange from a CATIA V5 FEM

simulation.

The sensor itself is comprised of three ceramic plates which are held in place by two

1.5 mm diameter centering pins and compressed together by four small springs housed in the

primary flange. The ceramic plates are made of Photoveel II, a proprietary-blend machinable

nitride ceramic manufactured by Ferrotec Europe GmbH, which has a rated thermal shock

Page 123: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

106

resistance of 600 K, making it suitable for environments where large temperature gradients

are expected such as in stratified flows or thermal shock scenarios [205]. More information

on Photoveel II as well as SHAPAL Hi-M SOFT, a suitable alternative ceramic, are shown in

Table 5-3. Two 3 mm thick ceramic plates act to hold and position the receiver and

transmitter electrodes, respectively, by employing sixteen 0.6 mm wide 0.6 mm deep grooves

per plate. The result is a minimum separation of approximately 2.4 mm between the electrode

layers. A third, 1.5 mm thick ceramic plate, completes the stack, ensuring electrical isolation

of the uppermost layer of electrodes from the lid flange.

Table 5-3. Ceramic materials suitable for a high-temperature mesh sensor operating in

steam-water environments with large thermal gradients.

The 16 receiver and 16 transmitter electrodes are positioned with a pitch of 4.49 mm

such that 208 crossing points lie within the 71.8 mm diameter pipe. A schematic and photo of

the substrate and a photograph of the sandwich construction in-situ with electrodes is shown

in Figure 5-5. The electrodes are EN 1.4404 (ANSI 316L) stainless steel capillaries of mean

ID/OD 0.31/0.57 mm. Splicing between the electrodes and transmission wires is made with

EN 1.4404 (ANSI 316L) stainless steel crimps of ID/OD 0.67/0.90 mm. The transmission

wires are PTFE insulated stranded wire with nineteen 0.10 mm OD silver coated copper

conductors.

Name Ceramic blend

TEC

[E-6 mm

/mm°C]

Thermal shock

resistance ΔT [K]

Bending Strength @

25°C [MPa]

Ferrotec Photoveel II Proprietary nitride

ceramic 1.4 600 440

SHAPAL™ Hi-M SOFT¹

Proprietary

aluminum nitride

ceramic

4.8 400 300

Page 124: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

107

Figure 5-5. Sketch of a single ceramic substrate for positioning 16 electrodes (top).

Photograph of the Photoveel II ceramic rings and electrodes installed in the housing flanges

(bottom).

Testing of the sensor at the LKE was performed in a purpose-built test-section

facility, shown in Figure 5-6. It is comprised of a 0.5 m long EN 1.4404 (ANSI 316L) pipe

welded on one end to a hemispherical cap, and on the other to a DN80 PN100 flange. A

second DN80 PN100 flange is welded directly to an end cap. Swagelok® connectors were

welded to the top and bottom end caps (3x and 1x, respectively). The prototype sensor is

flanged between the two components. The elongated nature of the setup is intended to allow

for efficient heating when wrapped with a heating tape.

Page 125: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

108

Figure 5-6. Photograph of test section (left) and drawing with values in mm (right) for

pressure and temperature testing of the mesh sensor package prototype.

The section has a 3.1 liter internal volume, similar to that of the autoclave. Cold

pressure tests were performed with a hand pump while the section was full of water. Nitrogen

may also be used to pressurize the facility. For hot tests, the setup includes resistance heating

tapes with PID control and glass fiber insulation. Pressure is read manually from a

manometer.

Page 126: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

109

5.6 Capabilities in theory

By selecting different materials within the same design concept, a second generation

sensor would be capable, in theory, of measuring steam-water flows in conduits at up to

22 MPa and 350ºC. Sealing the package at higher pressures, above those found in

commercial PWR power plants, requires only an upgraded feedthrough gland. Commercially

available multi-element feedthrough glands are built to withstand up to 22 MPa pressure. The

graphite gaskets typically used in flanged pipelines, already present in the prototype design,

are already able to withstand very high pressures (25 MPa). The greater challenge in high-

temperature mesh sensor design tends to be achieving temperature resistance. It remains

necessary to insulate the electrical conductors in the sensor from one another in the absence

of the traditional high-temperature insulator, PTFE which cannot withstand long term

operation above 300ºC. The solution is polyimide insulation, often classified as a type of

enamel insulator for metal conductors. Polyimide insulated wiring is commercially available

in a wide range of sizes and conductor types (at far higher prices than PTFE) and, according

to discussions with manufacturers, can withstand operating temperatures of at least 350ºC in

steam-water environments. Figure 5-7 shows, schematically, past high-temperature mesh

sensor designs relative to the current state of the art developed as a part of this thesis work

within a temperature-pressure-space assumed to be present in a water ambient.

Figure 5-7. Diagram (not to scale) showing the temperature and pressure capabilities of prior

state-of-the-art (left) and the current theoretical capabilities of the new mesh sensor package

(right, in red).

Aside from a discussion of design modifications for the increased pressure and

temperature resistance of the sensor, it is also worth describing ways in which the sensor

design could be expanded to incorporate additional instrumentation. The ability to easily

house two mesh sensors or a three layer mesh sensor opens the door to the measurement of

important flow properties, in addition to EC, at high temperatures and pressures. Prasser

(2013), inspired by Cox (1977), has described a generalized cross-correlation technique for

the reconstruction of fluctuating velocities well suited for data from a pair of mesh sensors

installed in a turbulent flow [206, 207] Furthermore, the a method for reconstruction of

interfacial area concentration, a valuable parameter in 1D nuclear safety thermal hydraulic

codes, has also been develop and improved for data from mesh sensor pairs [208-210].

Page 127: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

110

Furthermore, Shaban (2014) has developed a novel flow rate measurement method based on

void fraction signals from a WMS, this method benefits significantly from the use of dual

WMS units versus one alone [211]. Note that mesh sensor technology has also seen

successful implementation as a capacitance sensor for two-phase gas-oil flow distributions

[212, 213].

Figure 5-8. Sketches of a mesh sensor package incorporating a pair of 32-by-32 mesh sensors

along with eight multi-element feedthroughs in a DN80 PN100 housing. Adapted from

Kickhofel (2014) [193].

The additional space necessary for more sensor layers is at hand. A shortening of the

lid flange tongue is required, see Figure 5-8 (right). The strength of a flange of this geometry

has not been analyzed by way of FEM, however, it is foreseeable that it could be made

bulkier in order to increase its torsion resistance. The additional signal carriers can be

provided for by more boreholes in the primary housing flange along with additional multi-

element feedthroughs, if necessary, see Figure 5-8 (left). A single multi-element feedthrough

can pass up to 60 wires through the pressure barrier.

Page 128: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

111

5.7 Conclusion

A new mesh sensor construction has been designed for operation in high temperature

(potentially up to 350°C) and pressure (upper limit of 22 MPa) pipelines by implementing

novel materials, assemblies and a new sealing methodology. The mesh sensor package design

is scalable so as to be compatible with standard flanged pipelines of size estimated to be <

DN100 and > DN10. The invention solves the problems discussed related to prior state of the

art while at the same time allowing the sensor to operate at higher temperatures and to be

manufactured and assembled more easily and more cheaply. The pressure barrier is no longer

guaranteed by an epoxy, as in past designs by other researchers, but rather by a standard

graphite gasket between the housing flanges, the same type that would normally be sealing a

bolted flange joint, and commercially available threaded sealing glands. While still

maintaining space for the standard number of bolts for a given flange size, such sealing

glands could conceivably allow for a very high number of electrode signals to be extracted.

The ability to carry a large number of conductors through the pressure barrier, in addition to

the housing flange design, enables the installation of multiple mesh sensors, three layer

sensors, or other instrumentation such as thermocouples, enabling the reconstruction of

additional flow properties from measured data.

A 16-transmitter, 16-receiver prototype sensor has been constructed for a DN80

pipeline; its individual components and some assemblies thereof have been tested in an

autoclave at temperatures up to 285ºC in a liquid water-vapor environment with boiling at

saturation pressure. The prototype has also undergone pressure tests at a facility built at the

LKE for commissioning the sensor prior to proof-of-concept testing at the University of

Stuttgart.

Portions of the work discussed in this chapter have been patented or are included in

project reports of Master‘s students supervised by myself at the ETH Zurich LKE.

Specifically,

1. Kickhofel, J. and Prasser, H.-M., ‖Grid Sensor Package,‖ European Patent

Application EP14198142, ETH-Invention-No. 2014-053, Priority date December 16,

2014.

2. Semester project report of Ayer (2014): ―Design, Construction, and Implementation

of High Temperature, High Pressure Wire Mesh Sensors‖

3. Semester project report of Andermatt (2015): ―Capillary Based Wire Mesh Sensor for

High Temperature and High Pressure Applications‖

Page 129: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

112

5.8 Appendix

5.8.1 Alternate electrode materials

The stainless steel capillaries currently incorporated in the mesh sensor design would

likely deflect unacceptable amounts, such that the quality of the measurement is disturbed,

should they face high velocity flows. Added stiffness could be achieved with a material with

higher Young‘s modulus, such as Chromium. Within the housing flange design, an early

prototype of the mesh sensor sandwich incorporated stiff electrodes fixed on both ends to the

ceramic substrate by a high-temperature non-EC epoxy which performed well in autoclave

tests (EPO-TEK® OE188). The electrodes were made of twisted, 1K tow carbon fiber yarn

with a 120 µm layer of nickel deposited by electroless nickel plating, for improved stability

and EC. The electrodes were fixed without fear of damage because carbon fiber exhibits a

very small (assumed to be negligible) negative, transitioning to slightly positive, axial

thermal expansion between 20°C and 300°C. Induction brazing spliced the electrodes and

transmission wires using silver as a filler metal. The sensor design based on nickel plated

carbon fiber electrodes was tested in-house and at the University of Stuttgart T-junction

facility, where it failed due to epoxy liftoff which resulted in the displacement of the

electrodes.36

The liftoff was deemed to be the result of a combination of a TEC mismatch

between the epoxy (with TEC 68E-6 mm/mm above the glass transition temperature) and the

Photoveel II ceramic (TEC 1.4E-6 mm/mm) and vibrations in the facility reaching the bond

line.

To reach the small electrode diameters of 0.12 to 0.25 mm found in spring-based

sensor designs it is unlikely that there is any suitable material in rod or capillary form which

would not deflect unacceptably in a turbulent flow across the pipe diameters of interest. The

extension of the sensor to multiphase flows would benefit from a modified sensor sandwich

design integrating thinner electrodes. The transition to a spring based design can be

conceptualized within the current housing design. Ideally, such a design would rely on

custom-stamped springs significantly smaller in size than traditional compression or tension

coil springs. Short of a spring-based design, some testing has been performed with the

widespread shape-memory-alloy NiTi (also known as nitinol) which is manufactured in wire

form in many diameters [214].

An intriguing property of SMAs, for the mesh sensor designer, is their

pseudoelasticity (the result of a stress-induced crystal phase transformation) which allows for

the reversible deformation of large strains (>5%); strains which would otherwise result in the

plastic deformation of typical mesh sensor electrode materials. The yield strain of SMAs is

more than sufficient to compensate for electrode thermal expansion in high-temperature mesh

sensors [215]. Unfortunately, the pseudoelasticity of NiTi is lost long before reaching the

temperatures of interest (i.e. 250°C to 350°C). New SMAs based on NiTi with ternary

element addition, typically Palladium, extend the transition temperature, and therefore the

operating temperature limit of the material, to approximately 220°C [216, 217]. Such

materials may already be acquired commercially, albeit at high cost, from DYNALLOY Inc.

36

In this version the ceramic substrates were smooth rings without electrode-housing grooves.

Page 130: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

113

As the maturity of these ternary SMAs advances, they may represent an ideal material for a

fixed-wire high-temperature mesh sensor.

Page 131: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

114

6 Cross-flow mixing at high ΔT

6.1 Introduction

Collaboration between the ETH Zurich LKE and the University of Stuttgart IKE

undertaken as a part of this thesis has seen the Stuttgart T-junction facility taken to its highest

operating temperature, resulting in a bulk ΔT of 232 K between main and branch flows at

7 MPa pressure, during the proof-of-concept testing of the high-temperature high-pressure

mesh sensor package described in Chapter 5. This chapter explores the results of those

measurements in detail. Section 1.1 describes the T-junction facility at the University of

Stuttgart followed, in Section 1.1, by a detailed description of the mesh sensor module

installed for the purpose of mixing measurements. The test matrix, four different flow inlet

temperature boundary conditions, and experimental procedure at the facility during and

between measurements is found in Section 1.1. Results and analysis are divided into

measurements downstream and upstream of the T-junction in Sections 1.1 and 1.1,

respectively.

Page 132: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

115

6.2 Facility

At the epicenter of the T-junction facility at the University of Stuttgart, known as the

Fluid-Structure Interaction (FSI) test facility, is a sharp edged forged EN 1.4550 (ANSI 347,

X6 CrNiNb 18-10) austenitic stainless steel T-junction with reduced carbon content in

accordance with the German KTA 3201.1 [73]. The facility is shown schematically in Figure

6-1 along with the properties of the working fluid (deionized water with EC between 5 and 10

μS/cm at 20°C) and flow conditions. A main pipe with inner diameter Dm = 71.8 mm carrying

hot water at a constant mass flow rate of 0.4 kg/s (turbulent flow) encounters a perpendicular,

horizontal branch pipe of inner diameter Db = 38.9 mm with room temperature water flowing

at a rate of 0.1 kg/s (transition flow). Downstream of the T-junction the flow is then split

according to the mass flow rate ratio and recycled back to the T-junction; the flow headed for

the branch line first passes through a heat exchanger. The mass flow rate ratio at the T-

junction is such that at low ΔT = Tm – Tb a wall-jet type flow pattern is manifest downstream

of the T-junction at these flow rates. At higher ΔT, the cold branch line flow quickly sinks to

the bottom of the mixing pipe downstream of the T-junction and only weakly mixes with the

hotter, lighter main flow. More information about the flow behavior at the facility beyond the

findings articulated in this cahpter is described in the work of Kuschewski (2011, 2013,

NWLED-IF experiments), Klören (2012, 2013, LES simulations), and, most recently, Selvam

(2014, 2015, LES simulations) [59, 60, 88, 89, 218, 219]. The hot flow is measured once per

minute 1015 mm upstream of the T-junction in the center of the main pipe by a thermocouple

(XTM3) while the temperature in the branch flow is measured with the same frequency by a

thermocouple in the center of the branch line flow 2940 mm upstream of the T-junction

(XTM4).

Figure 6-1. Piping and instrumentation diagram of the FSI test facility at the University of

Stuttgart adapted from Kuschewski (2013) [60]. Temperature and velocity in the main pipe

represent maximum, bulk values.

Page 133: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

116

The facility is operated at temperatures up to 256°C (measured by XTM3) and

pressures up to 8 MPa. Resistance heating pads along the length of the main pipe inlet run are

regulated by a PID controller with reference thermocouples located between the mats and the

outer walls of the conduit. Heating of the circulated working fluid proceeds in the main pipe

at a typical rate of 40 K/hour. Meanwhile, the room temperature flow from the branch line

never ceases, such that mixing is always occurring at the T-junction. The flange system used

in the main pipe and mixing pipe are standard, specifically DN80 PN100, with the exception

of the inner diameter (71.8 mm) which is custom. With the exception of the T-junction, the

piping is comprised of austenitic stainless steel EN 1.4571 (ANSI 316Ti, X6CrNiMoTi17-12-

2). Flanged modules 340 mm in length neighbor the T-junction. Two modules in the main

pipe upstream of the T-junction and three in the mixing pipe downstream allow for the

installation of measurement instrumentation at a variety of locations.

At high main flow temperatures, stratification in the flow results in large thermal

stresses in the facility which induce a bending moment in the pipeline, which is not

anchored.[220] This behavior has been investigated in FEM simulations by the MPA

Stuttgart based on temperature profiles from LES with conjugate heat transfer [221]. Two

simulations were performed based on main flow temperatures of Tm = 120 and 280ºC. At

120ºC, before the onset of cold upstream flow, modules downstream of the T-junction rises to

a level slightly higher than the T-junction module. Significant upstream flow at 280ºC results

in the T-junction module along with the first two upstream modules rising significantly such

that downstream modules are no longer horizontal to gravity but slightly sloped. It cannot be

excluded that this phenomenon may have an effect on mixing behavior at the T-junction.

Page 134: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

117

6.3 Instrumentation

6.3.1 Mesh sensor module

The mesh sensor package prototype, described in Chapter 5, was fixed to a flanged

‗spacer‘ module of reduced length manufactured for adapting the sensor to the Stuttgart T-

junction facility. Four thermocouples were installed at four circumferential positions on the

outer wall of this shortened module. A photograph of the mesh sensor module installed

upstream of the T-junction is shown in Figure 6-2. The orientation of the mesh when installed

is shown in Figure 6-3 where the viewing orientation is looking away from the T-junction.

The free cross-section across both electrode layers is approximately 77% and the pressure

drop coefficient, again estimated from Idel'chick (1966), is 0.26 [168].

Figure 6-2. Mesh sensor package (red box) installed upstream of the T-junction at -3.5Dm,

thermocouples on the outer wall of the mesh sensor module spacer (green box), and

additional thermocouples (blue boxes), are shown installed at the FSI test facility.

Figure 6-3. Sensor sandwich comprising three ceramic plates and stainless steel capillary

electrodes. The orientation of the sensor is shown as if viewed from the T-junction, the

Cartesian mesh is tilted 22.5º relative to gravity.

Page 135: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

118

The calibration of the mesh sensor relies on a nearby thermocouple to compute a

relationship between the raw mesh sensor signal (which is proportional to the EC of the fluid)

at a given crossing point and temperature. Since mixing downstream of the T-junction cannot

be excluded (nor can mixing upstream at higher temperatures), calibration of some crossing

points near the bottom of the pipe is a challenge. Full details of the calibration methodology

is presented in Appendix Section 6.10.1. The resolution of the sensor in terms of EC is high

and is limited primarily by electronic noise. Signal post-processing as a means to filter noise

is described in Appendix Section 6.10.2. A linear function is used to describe mesh sensor

signal as a function of temperature across the range of interest (roughly 60 to 200ºC). The

gain-magnitude frequency response of the mesh sensor and thermocouples are discussed in

Appendix Section 6.10.3.

At four positions (0º, 90º, 180º and 270º) on the mesh sensor module spacer 1 mm K-

type thermocouples were installed. These thermocouples had a measurement frequency of 0.5

Hz. Additional thermocouples at the facility at various in-wall and in-flow depths

(measurement frequency 100 Hz) as well as those related to the heating mats may be utilized

to track the flow behavior. All thermocouples, as well as the associated acquisition device,

were provided and managed by researchers at the University of Stuttgart.

Page 136: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

119

6.4 Test Matrix

Two measurement campaigns were performed at the Stuttgart T-junction facility; both

saw the temperature in the main pipe rise from room temperature to 256°C in the center of

the main flow, as measured by the XTM3 thermocouple, over the course of roughly 8 hours

at 7 MPa pressure. During the first campaign the mesh sensor measurement plane was located

at y = 3.5Dm. The second campaign saw the mesh sensor installed upstream of the T-junction

at -3.5Dm with additional thermocouples placed on the mesh sensor module spacer at -5.5Dm.

Both configurations are shown in

Figure 6-4. Configuration of the Stuttgart T-junction facility during the first (top) and second

(bottom) measurement campaigns. Main flow travels right to left, gravity is directed into the

page.

Page 137: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

120

Figure 6-5. Temperature history recorded once per minute in the center of the main pipe

(XTM3) and branch line (XTM4) upstream of the T-junction, dotted lines represent analyzed

fluid temperature test cases.

The heating was occasionally stabilized for some minutes, such that longer

measurements with better statistics could be acquired. Four test cases have been isolated for

deeper analysis; they are main flow temperatures of 165±1ºC, 181±1ºC, 201±1ºC and

256±1ºC with a branch flow temperature of 22±1ºC. The main pipe and branch line

temperature histories are shown in Figure 6-5 for both measurement campaigns. Due to

instability in the main flow temperature in the second measurement campaign, the Tm =

201±1 ºC case is left unanalyzed. The main and branch mass flow rates were equal for all

three test cases: = 0.4 kg/s and = 0.1 kg/s. The fluid conditions in all four test cases

are enumerated in Table 6-1. The Froude number is defined, along the lines of Eq. 3-15, as,

√ , (6-1)

where ′ is the reduced gravity and the mixture velocity Umix is defined as the sum of

superficial velocities,

, (6-2)

where A is the main pipe flow area. The Prandlt number is defined as,

⁄ , (6-3)

where cp is the specific heat, μ is the dynamic viscosity, and k is the thermal conductivity. It

represents the ratio of viscous diffusion rate to thermal diffusion rate. In the branch fluid, Pr

is between 6.4 and 6.9 for all test cases while in the far less viscous main flow, in the highest

temperature case, Pr is around its minimum in water, 0.83.

Herein the Richardson number is calculated as,

, (6-4)

Page 138: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

121

where ⁄ is the density difference. This definition of the Richardson

number is similar to that found in Chapuliot (2005) with the difference being the use of the

mixture velocity rather than the branch line velocity [1]. Finally, the momentum ratio is

defined as in Kamide (2009) as [46],

. (6-5)

Test case Tm

[ºC]

ΔT*

[K]

ρm

[kg/m3]

ρr*

[-] ε

*

Ur*

[-]

pr

[-]

Rem

[E4]

Rer

[-]

Fr

[-]

Ri

[-]

#1 165 143 906.3 0.91 0.094 1.30 3.58 4.27 12.12 1.62 3.7

#2 181 159 889.9 0.89 0.11 1.32 3.64 4.71 13.37 1.52 4.2

#3 201 179 867.6 0.87 0.13 1.35 3.74 5.26 14.94 1.41 4.9

#4 256 232 793.0 0.79 0.21 1.48 4.10 6.81 19.35 1.22 6.6

Table 6-1. Fluid conditions for each experimental case including Reynolds, Froude and

Richardson numbers. Subscript r represents the ratio between main and branch flow water.

*Based on Tb = 22-24ºC, ρb = 1000.6 kg/m3, Ub = 0.084 m/s.

Page 139: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

122

6.5 Regarding the orientation of presented results

Both downstream (Campaign #1) and upstream (Campaign #2) mesh sensor data is

presented from the perspective of the T-junction in Sections 1.1, 1.1 and 1.1. A schematic of

the data viewing orientation with coordinate axes is shown in Figure 6-6.

Figure 6-6. Representative position of the mesh sensor and data viewing orientation (always

looking away from the T-junction) for mesh sensor installation at 3.5Dm downstream of the

T-junction in the mixing pipe (Campaign #1, left) or -3.5Dm upstream in the main pipe

(Campaign #2, right).

By interpolating the mesh sensor data onto a polar coordinate space, results are analyzed

along a perimeter half of the mesh sensor pitch (2.24 mm) inside the inner pipe wall, i.e. at a

radius of 33.66 mm. Via this interpolation, the position of the stratified layer on the left-side

and right-side pipe wall may be reported as angular position along the arc with length S, see

Figure 6-7. Furthermore, the maximum gradient of the temperature along the arc

⁄ , is also calculated both as time-averaged values as well as time-resolved values.

See Appendix Section 6.10.4 for details.

Figure 6-7. Sketch of variables related to stratified flow in the T-junction. Angular positions

α1 and α2 represent the location of maximum thermal gradient ⁄ along the circular

perimeter (in red) with r = 33.66 mm

Page 140: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

123

6.6 Downstream results

The mesh sensor module positioned at 3.5Dm elucidates the and TRMS profiles within

the cross-section of the pipeline. It is a considerable advantage of the mesh sensor compared

to traditional thermocouples that local gradients are resolved far better due to the dense

matrix of measuring location. Figure 6-8 shows plots of both mean and fluctuating

temperature fields as measured by the mesh sensor for all test cases. Density differences

between the hot main flow and cold branch flow are seen to play a strong role in the mixing

behavior downstream of the T-junction. As the main flow temperature increases (and thereby

the water density decreases) the mixing zone downstream of the T-junction, as measured by

the extent of elevated temperature fluctuations in the cross-section, tends to sharpen. As the

mixing zone shrinks the amplitude of temperature fluctuations increases to around TRMS = 10

K. All tests show similar profiles in which a mixing layer between hot main flow and cold

branch flow spans the inner diameter of the pipe. The mixing layer in the profile is oriented

perpendicular to gravity except in the first test case where the layer appears to be oriented

approximately 20º relative to the horizontal. Below the mixing layer, cold fluid temperatures

(< 75ºC) are measured even in mixing at the highest ΔT (Tm = 256ºC). The measured flow

profiles cannot be categorized within the traditional low-ΔT cross-flow mixing nomenclature

of Kamide (2009) [46].

The temperature profiles measured in the third and fourth test cases are unreliable in

the hot fluid region at the top of the pipeline due to the ambiguous dependency of the EC of

water versus temperature (see Appendix Section 6.10.1). In the important lower portions of

the pipe cross-section, where temperatures are below 200ºC, the temperature profile can be

reported with confidence knowing that temperatures in the bottom of the pipe have risen to

these levels. The zone of average cold fluid at the bottom of the pipe is greatly reduced in

area, especially at Tm = 256ºC while the minimum average temperatures, of 70 to 80ºC, are

only marginally elevated compared to lower ΔT cases. Of all test cases, temperature

fluctuation amplitudes are highest in the mixing zone found in the fourth test and,

furthermore, the fluctuations in this case are the most localized.

Page 141: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

124

Figure 6-8. Time-averaged (top row) and RMS (bottom row) of temperature in the cross-

section of the pipeline measured by the mesh sensor 3.5Dm downstream of the T-junction.1

Viewing orientation is looking away from the T-junction.

Figure 6-9. Ensemble averaged (K = 8) PSD at crossing-points (right) of highest TRMS for

each test case.

The PSD, computed at the crossing point measuring the highest TRMS for each test

case (see Figure 6-9) does not indicate a preferred frequency of mixing except in the fourth

test, in which frequencies between 1.6 Hz and 2.4 Hz show a plateau of high amplitude. The

spectrum from Test #1 is weighted slightly more towards the high frequencies relative to the

1 Technical difficulties with the 6

th transmitter electrode resulted in an intermittent signal which also affected its

neighboring transmitters. Therefore, the average and RMS plots in those three rows of crossing points are

compiled not from the whole 180 s measurement but from the portion in which the 6th

transmitter was

functioning properly. Nonetheless, it is evident a signal of questionable validity remains in the 5th

, 6th

, and 7th

transmitter electrodes and therefore quantitative discussion related to those positions is avoided.

Page 142: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

125

other tests. Apart from this, the four spectra are similar in shape. Power spectra at some

positions measured by the mesh sensor show similarities with spectra from downstream in-

flow (2 mm from inner wall) thermocouples at different positions. For example, in Figure

6-10 the thermal mixing spectrum from near the wall in the mesh sensor at a circumferential

position of 219° (based on a conventional polar coordinate system in the cross-section) is

shown to be similar to the spectrum 1.5Dm further downstream at 324.5° yet is dissimilar to

the spectrum downstream at nearly the same circumferential position, 234.5°. This

asymmetry may be the result of the swinging motion of the heavy cold flow as it progresses

down the main pipe, described by both Klören (2011) and Selvam (2015) in mixing cases

with lower ΔT [82, 218]. This swinging motion may also explain the tilted layer of high

temperature gradient see in the profile in Figure 6-8.

Figure 6-10. PSD measured near the wall by the mesh sensor in Test #4 along with that from

a downstream thermocouple measurement.

Page 143: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

126

6.7 Upstream results

Upstream flow in the T-junction, the result of a slow turbulent main flow and large

density differences between the main and branch flows generating a stratified countercurrent

flow, is a significant feature of the flow behavior at high ΔT at the facility. The formation and

progression of upstream flow with increasing ΔT, starting at the T-junction, may be tracked

by a combination of thermocouples in the walls of the main pipe, on the outside of the main

pipe, and by the mesh sensor.1 Specifically, measurement points are present at -0.96 (1.5 mm

from inner wall), -1.58 (1.8 mm from inner wall), -3.5 (mesh sensor module), -5.9 (outer

wall), -10.1 (outer wall) and –40.7Dm (outer wall). The onset of upstream flow has been

found to be more akin to a thermal shock than a slow, ramped cooling of the bottom of the

conduit. Therefore, thermocouples, even those on the outside wall of the main pipe, measure

a sudden plummet in the temperature which can help to localize in time the moment upstream

flow reached the measurement location. The thermal inertia of the wall naturally results in a

slightly delayed indicator. This delay is neglected, for simplicity, since the time-scales of the

upstream flow progression are large relative to that of heat transfer within the steel pipe wall.

The penetration depth of the upstream flow into the main pipe extends more than

2.92 m based on thermocouple readings. The upstream penetration of the cold branch flow as

a function of ΔT is shown in Figure 6-11. The position of the mesh sensor (-3.5Dm) appears to

represent the approximate vicinity of a sharp reduction in the rate of upstream cold flow

penetration with increasing ΔT. The reproducibility of these measurements in another facility

may be hindered by the heating of the upstream flow by the heating pads which are installed

around the main pipe starting at approximately -13.6Dm. This heating could conceivably slow

the upstream flow progression by introducing an additional heat source.

Figure 6-11. Data points represent the ΔT (= Tm - Tb) at which upstream flow was detected (as

a drop in temperature) at thermocouples measurement locations or the mesh sensor during

both the first and second measurement campaigns. Vertical dotted line represents the location

at which heating pads surrounding the pipe end.

1 Recall that all thermocouple instrumentation was the property of the University of Stuttgart.

Page 144: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

127

The upstream flow onset, as measured by the mesh sensor, includes a number of

important features. First and foremost is the rate with which the temperature is seen to

decrease at the bottom of the main pipe at -3.5Dm. A number of crossing points at the bottom

of the pipe detect a sudden drop in temperature for which the cold branch line water is

responsible. Part of the initial decrease in temperature2 includes a portion in which the fluid

temperature drops from 140 to 100ºC at a rate of approximately -130 K/s [30]. Near the wall,

the temperature fluctuations recorded at two crossing points near the pipe-bottom are

TRMS(16,11) = 6.3 K and TRMS(15,11) = 8.4 K in the first 10 seconds after the onset of

upstream flow. The fluctuations are greatly reduced in magnitude some 10‘s of seconds later,

between 30 and 40 s after the onset of upstream flow to TRMS(16,11) = 3.1 K and TRMS(15,11)

= 5.2 K.

Figure 6-12. Temperature history at three mesh sensor crossing points (see locations relative

to mesh sensor sandwich, right) at the bottom of the main pipe (location indicated Figure 6-3)

as cold upstream flow is detected at -3.5Dm.

Figure 6-13. Temperature history at thermocouples placed on the outside wall of the mesh

sensor spacer section (see locations relative to pipe-wall cross-section, right) at -5.5Dm as

cold upstream flow is detected at the bottom of the pipe.

2 The temperature history is reminiscent of thermal shock experienced in thermal stratification tests in a mockup

of horizontal feedwater piping at the HDR-facility in the 1980‘s, discussed in Section 1.2.4. Specifically, Figure

7 of Wolf (1989).

Page 145: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

128

Detection of the onset of upstream flow outside of the pipeline is shown in Figure

6-13. Outside of the 8.55 mm thick steel wall at the bottom of the conduit (T7), due to the

sudden presence of cold upstream flow from the branch line, a sudden decrease in

temperature at a maximum rate more than two orders of magnitude lower than in the flow is

recorded. The thermal inertia of the pipeline limits the maximum rate of temperature decrease

is approximately -1.14 K/s at -5.5Dm in this thermal shock situation. The temperature

stabilizes approximately 15 minutes after the onset of upstream flow after decreasing from

154 to 94ºC. Note that the stabilization of the temperatures on the side (T11) and top (T13) of

the conduit are not the result of upstream flow but due to the PID heating program regulating

the main flow heating, preparing for test case #1.

Figure 6-14. Time-averaged (top row) and RMS (bottom row) of temperature in the cross-

section of the pipeline measured by the mesh -3.5Dm sensor upstream of the T-junction.

Viewing orientation is looking away from the T-junction.3

Mean and fluctuation temperature and fluctuation profiles are shown in Figure 6-14

for Test #1, #2 and #4 upstream of the T-junction. In Test #1 the mesh sensor measures a

small amount of upstream flow at the very bottom of the pipe with temperature of

approximately 100ºC. Neighboring crossing points see a mean temperature increase of over

30ºC. The highest temperature fluctuations are found in the thin layer between the cold

upstream flow and the hot main flow. The remainder of the cross-section shows negligible

temperature fluctuations. With higher ΔT in Test #2, the amount of upstream flow in the pipe

3 Technical difficulties with the 6

th transmitter electrode resulted in an intermittent signal which also affected its

neighboring transmitters. Therefore, the average and RMS plots in those three rows of crossing points are

compiled not from the whole 180 s measurement but from the portion in which the 6th

transmitter was

functioning, in the case of Test #1. Nonetheless, it is evident a signal of questionable validity remains in the 5th

,

6th

, and 7th transmitter electrodes and therefore quantitative discussion related to those positions is avoided.

Insufficient up-time was found in the 6th

transmitter electrode in Test #2 and #4, it has therefore been left blank.

Page 146: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

129

cross section is increased and the temperature gradient between the zone of upstream flow

and the hot flow is considerably larger. The RMS of the temperature, however, is decreased

compared to the lower ΔT case. Finally, by including interpretation of the and TRMS profiles

from the fourth test case it can be concluded that the fraction of cold fluid in the cross section

is increasing with increasing ΔT, this trend is quantified later in this chapter.

The maximum time-averaged thermal gradient near the pipe wall ⁄ ,

increases drastically after the onset of upstream flow, which occurs at the sensor position (-

3.5Dm) between Ri = 3.53 and 3.65, to values of around 4 K/mm at both the left and right

walls, see Figure 6-15. The average thermal gradient at the edges of the stratified layer

increases sharply as a function of Richardson number, thereafter. Around Ri = 5 the

magnitude of ⁄ plateaus at between 9 and 11 K/mm at the right and left walls,

respectively.

Figure 6-15. Maximum average thermal gradient ⁄ near the wall -3.5Dm upstream

of the T-junction vs. Richardson number. The onset of upstream flow is apparent in the step

change (indicated by dotted red arrow) in thermal gradient around Ri = 0.48.

For a perspective of the time-dependent behavior of the cold flow upstream of the

T-junction, the angular position of the stratified layer at the right (α1) and left walls (α2) of the

pipe is plotted as a function of time during test cases #1, #2 and #4 in Figure 6-16. The

angular position of the stratified layer upstream of the T-junction, a zone of countercurrent

stratified flow, fluctuates in a range sometimes larger than 20º within one second at the right

and left wall, respectively, during the first test case (Ri = 3.7). Already at Ri = 4.2 in Test #2

the interface is far more stable, showing few disturbances during 10 s. Similarly, at the

highest ΔT, in the final test case with a Richardson number of 6.6, while having risen in the

pipe the interface shows fluctuations of only 5 to 6º at each wall.

Page 147: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

130

Figure 6-16. Position of maximum thermal gradient ⁄ over the course of 10

seconds along the right (α1) and left walls (α2) pipe upstream of the T-junction at -3.5Dm.

Although the location of largest thermal gradient with respect to the near-wall

arc in the flow does not exude a dynamic nature, the magnitude of that thermal gradient is not

steady in time, even in Test #4. Figure 6-17 shows the time history of the thermal gradient

⁄ at locations α1 and α2 for the same 10 s measurement as in the previous figure.

Note that the instantaneous thermal gradient near the wall is significantly higher than the

average gradients shown in Figure 6-15, the result of movement of the interface blurring the

average gradient.

Figure 6-17. Maximum thermal gradient ⁄ over the course of 10 seconds along

the left (bottom) and right (top) walls in the pipe

Figure 6-18 shows the PSD for two test cases at three crossing points near the bottom of the

main pipe (recall Figure 6-3 right). At Tm = 165ºC (Test #1) the highest RMS is measured at

(15,11), the PSD at this position shows a plateau at frequencies below 0.7 Hz without

evidence of a preferred mixing frequency. A larger portion of the RMS of the temperature

arises from low frequencies in the mixing layer (15,11) than above (14,11) or below it (16,11)

. In the case of a larger temperature difference, in Test #2, the crossing point (14,11) lies

Page 148: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

131

within the mixing zone between the cold upstream flow and the hot main flow. The mixing

spectrum at this crossing point also exhibits a plateau at lower frequencies with a peak at 0.1

Hz. Considering the materials and geometries of interest, 0.1 to 1 Hz is typically cited as the

range of interest related to thermal fatigue [222]. Timperi (2010) has conducted an LES

simulation of an HDR thermal stratification experiment in which a horizontal DN400 mockup

of a feedwater line is exposed to stratified flow between hot water at 210 and cold water at

60; the most dominant frequency found in the mixing layer is reported to be ―a little less than

1 Hz‖ [31].

Figure 6-18. Ensemble averaged PSD (K=4) at three crossing points near the bottom of the

main pipe in Test #1 (left) and #2 (right).

A regularly extending and receding upstream flow could result in thermal fatigue in

the main pipe analogous to the case of turbulent penetration in the branch line, e.g. the

Farley-Tihange phenomena. As discussed in Section 4.8, that mixing scenario sees a cold

tongue of water impinging into a branch line filled from the main pipe with much hotter

water [223]. In the case of the measurement campaigns at the Stuttgart T-junction facility,

since the heating program calls for a continuous rise in Tm, the penetration depth of the

upstream flow only extends and does not appear to extend and recede in a transient manner,

within the resolution of the available instrumentation at the facility.

Page 149: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

132

6.8 Combined analysis

A compilation of mesh sensor data from measurements at the T-junction facility

indicate that the cold branch flow is being split between the mixing pipe and main pipe in

larger and larger fractions with increasing density difference, explaining the decreasing cold

flow area with increasing ΔT in the mixing pipe along with the concurrent increase in the

main pipe. The trend towards symmetric stratified profiles upstream and downstream may be

exacerbated by the fluid-structure interaction occurring at the facility, in the sense that the T-

junction becomes the highest point in the facility at high temperature. The mixing scenarios

are categorized according to ΔT in Figure 6-19. The experiments provide the opportunity to

investigate temperature distributions in thermally stratified flows with presumably different

velocity gradients, upstream versus downstream of the T-junction, within the same mixing

scenario.

Figure 6-19. Schematic description of flow behavior at the Stuttgart T-junction facility at

low, mid, and high ΔT.

Made possible by the mesh sensor is the analysis of the thermal gradient, both in the

center of the pipe and near the walls, as well as the area of cold flow within the pipe cross-

section. Downstream of the T-junction the time-averaged temperature in the cross-section, as

viewed from the side of the pipe in Figure 6-20 (left), shows the somewhat linear behavior of

the temperature profile at the bottom of the pipe with steepening slope as ΔT increases

between Test #1 to #4. Meanwhile, upstream of the T-junction in Figure 6-20 (right), where

the interfacial friction between cold and hot fluids is presumably stronger due to the

countercurrent flow pattern, the thermal profile takes on a distinct S-shape with steeper

profile than found downstream.

Page 150: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

133

Figure 6-20. Thermal gradient profile viewed in the plane of the mesh sensor. Each semi-

transparent bar represents time-averaged temperature measured by a crossing point of the

sensor. Dotted boxes approximately isolate the mixing zone.

The maximum thermal gradient in the center of the cross-section (see Appendix

Section 6.10.4) downstream of the T-junction is found to be a linearly increasing function of

Richardson number over a large range from Ri = 1.78 to 6.57 (ΔT = 78 to 232ºC), see Figure

6-21. In absolute terms, the maximum thermal gradient more than doubles from around 4

K/mm to 9.5 K/mm at 3.5Dm. Each data point represents the gradient of temperature from a

5-second time-averaged mesh sensor signal extracted from each consecutive full 180 s

measurement during the heating of the facility. Upstream of the T-junction the thermal

gradient prior to upstream flow onset is low, around 1 K/mm and increases slowly with Ri

number. The onset of upstream flow results in a jump in the maximum thermal gradient

found in the centerline (shown with dotted red arrow Figure 6-21) to around 8 K/mm. The

stratified layer sharpens further resulting in a maximum thermal gradient of around 14.5±1

K/mm between Ri = 4.5 and 6.6, significantly higher than found downstream of the T-

junction.

Page 151: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

134

Figure 6-21. Maximum average thermal gradient ⁄ across the stratified layer in

the center of the pipe cross-section vs. Richardson number. The onset of upstream flow is

apparent in the step change (indicated by dotted red arrow) in thermal gradient around

Ri = 3.6.

As deduced previously in a qualitative sense from profiles measured by the mesh

sensor (recall Figure 6-8), the cold holdup area appears to decrease 3.5Dm downstream of the

T-junction with increasing ΔT. Figure 6-22 shows quantitatively the trend in cold holdup as

% of the total flow area (see Appendix Section 6.10.5) versus Ri number. Indeed it happens

that the cold flow area below the stratified layer shrinks with increasing Ri number while,

concurrently, the same holdup area -3.5Dm upstream of the T-junction increases with Ri

number. As the Richardson number increases beyond 6 the holdup area upstream of the T-

junction becomes larger than downstream at the same non-dimensional positions (±3.5Dm).

As Ri increases, it has steadily less influence on the proportional area of heavier cold flow in

the cross-section both upstream and downstream of the T-junction.

Figure 6-22. Cold holdup % in the pipe cross section vs. Richardson number 3.5Dm

downstream -3.5Dm and upstream of the T-junction.

The results indicate a steady transition to an increasingly stratified flow as the main

pipe flow temperature increases. The time-averaged location of the stratified layer may be

discerned both downstream and upstream of the T-junction by localizing the largest time-

Page 152: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

135

averaged near-wall thermal gradient. Since the stratified layer spans the conduit, it is possible

to discern two peaks in ⁄ around the inner perimeter of the pipe; one peak corresponds

to one side of the thermal discontinuity at the border between hot and cold fluids, the other

peak corresponding to the other side of this layer spanning the cross-section, recall Figure

6-7. The position of these maxima gives some indication of pipe wall locations prone to

cyclic loading due to instabilities in the stratified layer. Figure 6-23 shows the angular wall

positions α1 and α2, of maximum average thermal gradient ⁄ , downstream (left,

subscript ‗d‘) and upstream (right, subscript ‗u‘) of the T-junction. The stratified layer moves

downwards with Richardson number downstream of the T-junction, and moves upwards

upstream of the T-junction. The shift is small, however, with α2d moving from 214º to 225.4º

and α1d to as little as 310.4º. Upstream, the position of the interface between cold branch flow

moving underneath and counter to the hot main flow rises slightly between Ri = 3.65 and 6.57

to positions 327.4º and 216.9º.

Figure 6-23. Location of maximum average thermal gradient near the wall ⁄ ,

3.5Dm downstream (left) and -3.5Dm upstream (right) of the T-junction vs. Richardson

number.1

1 Missing data for α1d is the result of a malfunctioning 6

th transmitter electrode initially discussed with regards to

Figure 6-8.

Page 153: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

136

6.9 Conclusion

The measurements described in this chapter are party to a series of proof-of-concept

tests performed at the University of Stuttgart T-junction facility of a new mesh sensor

package for high-temperature and high-pressure flows, especially those related to nuclear

safety, described in Chapter 5. T-junction mixing with reactor-like temperature differences

between main pipe and branch line flows has been investigated at bulk temperatures of up to

256ºC. Instrumentation included thermocouples installed on the pipe outer wall, in-wall and

in-flow as well as a novel mesh sensor capable of measuring high-temperature and high-

pressure flows at high frequency (in this case 10 kHz) with high spatial resolution (in this

case 208 measurements points in the flow with a pitch of 4.49 mm). The mesh sensor

package, measuring EC, is used here to measure single phase mixing in the T-junction

without a salt tracer by means of a calibration routine based on the temperature dependence

of the EC of water. The mesh sensor was installed both downstream and upstream of the T-

junction where it elucidated the flow average and fluctuating temperature profiles. Four

boundary conditions, in particular, were investigated for extended durations of up to 10

minutes. These included mixing at main pipe flow temperatures ranging from 165ºC to

256ºC. The mixing behavior downstream of the T-junction cannot be described within the

usual categorizations of cross-flow mixing in T-junctions while upstream mixing appears

typical of countercurrent stratified flows found in other contexts.

The mesh sensor indicates a thermally stratified layer between cold fluid in the

bottom portion of the pipeline and hot fluid on top. As the Richardson number increases, the

interface between the main flow and branch flow in the mixing pipe becomes sharper and

contains larger amplitude temperature fluctuations. Downstream, the area of cold fluid in the

cross-section is seen to decrease with increasing Tm such that the highest temperature

fluctuations near the wall at the mesh sensor position move downwards along the walls from

the middle portion of the pipeline.

The upstream flow phenomena, cold flow penetrating more than 3 meters into the

main pipe underneath the incoming hot main flow has been documented in detail thanks to

the mesh sensor. The result of a stratified flow with high Richardson number, the cold flow

rests in a stable way on the very bottom of the pipe with temperature fluctuations residing in

a thin layer not thicker than the pitch of the mesh sensor. The thermal shock which occurs

when the cold upstream flow arrives at a particular location at the pipe bottom, in this case

measured at -3.5Dm upstream (in the flow, by the mesh sensor) and at -5.5Dm (outer wall, by

a thermocouple) of the T-junction was quantified. Furthermore, the cold holdup area of the

upstream flow was found to increase with Richardson number.

Portions of the work discussed in this chapter are to be published in peer-reviewed

conference proceedings as,

1. Kickhofel, J., Selvam, K., Huber, H., Laurien, E., and Prasser, H.-M.. ―Mesh Sensor

for High Temperature and High Pressure Applications,― 16th International Topical

Meeting on Nuclear Reactor Thermalhydraulics (NURETH-16), Chicago, USA,

August 30 - September 4, 2015.

Page 154: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

137

6.10 Appendix

6.10.1 Mesh sensor calibration methodology

The mesh sensor measures the EC of a small volume of fluid at each crossing point

between a transmitter and a receiver wire of the mesh. Typically, single phase mixing

experiments, such as studies of turbulent mixing at T-junctions, would utilize a salt tracer to

differentiate the main flow from the branch flow and therefore capture the mixing process as

in mixing studies at the LKE T-junction discussed in Chapter 4. The addition of a tracer was

not permitted at the Stuttgart T-junction facility and therefore the measurements rely on the

varying EC of water with temperature. The ion product Kw = [H+][OH-] (also referred to as

the dissociation constant), of pure water has been measured by Marshall (1981) over a wide

range of temperatures and pressures and subsequently accepted by the International

Association for the Properties of Water and Steam (IAPWS) [224]. Figure 6-24 (left) shows

the EC of pure water versus temperature at 7 MPa based on the correlation and coefficients of

Marshall (1981) [224].

Figure 6-24. EC vs. temperature (left) according to the correlation of Marshall (1981) for

pure water, and EC vs. temperature (right) from experiments with NaCl solutions at high

pressure from Quist (1967) [224, 225].

Truly pure water, however, does not exist in an engineering context. Measurements

by a handheld meter at the T-junction facility found the circulated water to have EC of 7-9

μS/cm1, which results in a different shape of the ion product versus temperature curve. It has

been shown, for electrolyte solutions such as NaCl, for example, that EC is essentially linear

to 200ºC over a wide range of pressures, see Figure 6-24 (right) [225]. The mobility of ions in

the water are the dominating contribution, compared to the self-dissociation of H2O, to the

EC as a function of temperature especially at elevated temperatures in the range of interest.2

The peak in EC is the result of ionic association counterbalancing the ionic mobility as the

1 A slight increase in working fluid room temperature EC of approximately 2 μS/cm was noted in campaign

before and after measurements, presumably due to corrosion products. 2 The sharp decrease in viscosity of water vs temperature as it is heated above room temperature is the reason

for increased ionic mobility.

Page 155: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

138

density of water continues to decrease with temperature. The nonlinearity of EC as a function

of temperature, seen especially below 100ºC, is lost with an increase in impurities. Therefore,

a linear behavior of the EC of the water over a range between approximately 60 and 200ºC

has been assumed.

The calibration of the mesh sensor at the Stuttgart T-junction is performed using the

raw data collected from the full day of heating and cooling the facility, with the sensor

installed upstream of the T-junction. For the calibration of each crossing point, a single

reference thermocouple was chosen. The reference thermocouple, XTM3, described earlier in

Section 1.1, lies in the center of the flow 765 mm upstream of the mesh sensor. It would be

incorrect to assume that the temperature at XTM3 is equal to the temperature across the

whole of the mesh sensor plane. A number of factors are at play; firstly, there is no heating of

the flow between XTM3 and the mesh sensor plane; some heat losses and mixing can be

expected. Secondly, and more importantly, at a Reynolds number of 54,000 (at 200ºC) in the

main pipe the flow is not sufficiently turbulent such that a flat temperature profile is achieved

in the pipe cross section. A top-to-bottom temperature gradient in the main pipe is expected

to exceed 10-15 K at high water temperatures based on thermocouple measurements at the

facility. Indeed it was observed that the signals from the crossing points close to the top of the

pipe cross-section change from increasing to decreasing tendency earlier during the heat-up

process than those close to the bottom. Therefore, a least squares regression was performed to

best estimate the reference temperature at each mesh sensor crossing point Tt,r, as a function

of reference thermocouple temperature Tr, at time tmax of highest crossing-point signal. This

approach takes benefit from the assumption that the maximum EC is always found at a

temperature of 240°C. The reference temperature Tr measured by XTM3 is different from

240°C in the same instant. By knowing the relationship between the reference temperature

Tr(tmax,t,r) and Tt,r(tmax,t,r) at the time of maximum signal at each crossing point (when

Tt,r(tmax,t,r) = 240°C) that relationship then helps to convert Tr(t) measured at any other time to

an estimated crossing point temperature Tt,r(t). The available data was analyzed to create a

table containing the reference temperature Tr(tmax) at the moment the turning point is reached

for each crossing point (t,r) of electrodes at its known spatial coordinates (z,x).

A least squares regression seeks the planar fitting to these 3D points of the form (z, x,

Tr(tmax)). The plane Tr + A + Bz + Cx describes the distribution of temperature in the pipe

cross section at the location of the mesh sensor which is known to exhibit a gradient from top

to bottom at high temperatures. The x-component is included to account for the possibility of

upstream swirl influencing the temperature profile along this axis, although a flow

straightening device is installed. The minimum of the function,

∑ ∑

, (6-6)

is sought via,

, (6-7)

Page 156: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

139

where Tt,r(tmax,t,r) is defined for all (t,r) crossing points as 240ºC: the temperature

corresponding to the maximum signal, and therefore maximum EC, at each crossing point.

The reference thermocouple reading, Tr at the time of maximum signal at crossing point (t,r),

tmax,t,r varies. In this way a function for the estimated temperature at each crossing point at

any given time, Tt,r = Tr(t) + A + Bzt,r + Cxt,r, may be generated assuming a tilted-planar

temperature profile in the cross section at 240ºC. The regression finds the following values, A

= 1.8139 [K], B = 701.8 [K/m] and C = 17.8 [K/m] indicating a top-to-bottom gradient in the

pipe of up to approximately 49 K and very little temperature change from left-to-right . The

coefficients B and C are only representative for the conditions of a hot facility at about

240°C. It is plausible to assume that the temperature gradients are proportional to the

difference between the reference and the ambient temperature. Therefore, a further step is

taken to scale the coefficients B and C with main flow temperature:

′ (

) (

), (6-8)

and similarly for C‟(t). With this correlation, the estimated temperature at the mesh sensor

was corrected for all available data points recorded during the heat-up phase.

An example of the relationship between the corrected reference temperature, Tt,r, and

the mesh sensor signal, θ, is shown in Figure 6-25. Visible is the peak in the EC of the water

at 240ºC. The region in which the crossing points are calibrated is where the signal shows

linear dependence to the temperature. A linear fit is made in this region for each crossing

point which becomes function which converts the raw measured signals to temperatures.

Figure 6-25. Plot of raw mesh sensor signal vs. estimated temperature at a crossing point near

the top of the pipe. The function which describes the linear fit, shown in red, is that which

converts signal values to temperature at this crossing point for any given measurement

dataset.

The presence of cold upstream flow at the sensor, present sometimes even during the

heat-up process, precludes the generation of complete calibration data for the crossing points

of the sensor at the bottom of the pipe, as in the case of the downstream calibration. We

therefore rely on the linear behavior of the EC of the fluid as a function of temperature in

Page 157: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

140

calibrating these crossing points for temperatures that we could otherwise not verify being

experienced at those locations at the bottom of the pipe where mixing interferes.

6.10.2 Noise filtering

Care has been taken to filter noise from the mesh sensor signal while maintaining

integrity of the signal as it relates to real fluid temperature fluctuations. Wire-mesh sensors

are known to be somewhat susceptible to electrical perturbations, which leads to a

superposition of the signals both with noise and glitches. Measurements recorded at 10 kHz

enable a variety of filtering measures to be implemented while maintaining more than

adequate sampling rates. Grounding of the pipeline to the DAQ system and shielded cables

was performed to reduce the noise as much as reasonably possible at the Stuttgart T-junction

facility. In order to further reduce electronic noise visible in the sensor signal, a median filter

approach has been implemented.

The 1-D median filter replaces each value with the median of a windowed range

around that value. We have chosen a window size of 14 values (a ‘14-order‘ median filter) in

order to eliminate large amplitude high frequency electronic noise and glitches observed in

some chosen reference signals. More information, such as boundary effects when

implementing the filter can be found in textbooks [226]. Figure 6-26 shows a sample signal

with and without filtering. Only the very high frequency portions of the spectrum are affected

by the filtering.

Figure 6-26. A comparison between one seconds of raw signal recorded by the mesh sensor,

and the corresponding PSD, in the mixing layer downstream of the T-junction with and

without median filtering.

By integrating the PSD we find that θRMS post-filtering is reduced by 3.7%. The

power spectrum remains largely unchanged at lower frequencies with 99.3% of the scalar

variance retained for a frequency domain of up to 500 Hz. Given the Reynolds number of the

flow and eddy turnover times we do not expect fluctuations in the fluid temperature at

frequencies above 500 Hz to be meaningful. No noise filtering has been performed on

thermocouple signals.

Page 158: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

141

6.10.3 Gain-magnitude frequency response

The frequency response of K-type thermocouples at the Stuttgart T-junction facility

has been investigated in the past by researchers at the IKE. The well-known low-pass

filtering effect present in the thermocouples is the product of their thermal inertia. The gain-

magnitude frequency response of the mesh sensor, based on the earlier discussion related to

the LKE T-junction WMSs in Appendix Section 2.3.3, is shown along with the thermocouple

response in Figure 6-27.

Figure 6-27. Gain-magnitude frequency response of thermocouples at the facility and of the

mesh sensor for a number of fluid velocities.

6.10.4 Calculation of instantaneous and maximum average thermal gradient

For higher resolution in the cross-section, mesh sensor data was passed through a 2D

cubic interpolation function in MATLAB onto a 64 x 64 Cartesian grid such that the new data

is C2 continuous. Two columns, corresponding to a width of 2.24 mm, were then averaged

row-wise such that a top to bottom temperature profile in the center of the conduit was

generated from the calibrated, median filtered mesh sensor data. Within the lower half of this

profile a maximum of the temperature derivative , was found and is reported as the

maximum thermal gradient. The mesh sensor data used in computing average values over a

wide range of Richardson numbers were calculated from 10 kHz signals, 5 seconds in

duration extracted from each full 180 s measurement.

The calculation of thermal gradient along an arc near the wall , has been

calculated based on mesh sensor data interpolated (again with a cubic spline interpolation)

onto a polar space with 128 angular positions at a radius half of the sensor electrode pitch less

than the pipe radius, i.e. 33.66 mm. Once again, average values are based on 5-second long

signals, one from each measurement file.

6.10.5 Calculation of cold holdup area

Page 159: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

142

Cold (branch) flow holdup was calculated as follows. Crossing points measuring

below a threshold temperature,

(6-9)

where Tm is the most recent main flow temperature recorded by reference thermocouple

(XTM3) and is the minimum average temperature calculated from the five second

mesh sensor signal were labeled as cold flow and given a value of 1 in the 16-by-16 2D

holdup matrix. Values above the threshold were set to 0. The 2D holdup matrix is multiplied

element-wise by a weight matrix, W, which takes into account the reduced measurement area

of crossing points near the wall of the conduit.

(∑ ∑

) (6-10)

where A is the pitch of the sensor squared, 20.14 mm2. An example of a 2D holdup matrix is

shown for an arbitrary time step in Figure 6-28. Note that the plot is not oriented with gravity

pointing downwards.

Figure 6-28. Instantaneous weighted cold holdup at Ri = 0.98 3.5Dm downstream of the

T-junction.

Page 160: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

143

7 Outlook

The experiments related to turbulent penetration herein represent a stepping stone

from which future studies can build. They provide a large amount of data to investigate and

interpret in the context of future, related endeavors. More experiments are necessary to better

understand the effects of additional parameters, such as fluid viscosity, on the mixing

behavior in the branch line. Simulations are an ideal tool to more comprehensively

understand the phenomena at work. This mixing scenario remains on the periphery of (and

beyond) that which time-resolved CFD simulations can today feasibly capture. The CFD

simulation of such mixing scenarios where long simulations times are necessitated remains a

challenge which will only become easier with enhanced computational means. Especially

valuable is to strive for improve simulation techniques and grid-generation practices for this

complex case of turbulent-laminar flow interaction based on the simulation and findings

presented herein.

The value of the mesh sensor as a tool for investigating high-temperature high-

pressure flows should not be understated. The sensor is not limited in applicability to cases

relevant to thermal fatigue in T-junctions but could be implemented in a wide range of safety-

relevant flows, including two-phase flows and those outside of the field of nuclear

engineering. An enhanced design including multiple mesh sensors, thermocouples or other

complementary instrumentation would result in a powerful measurement suite within one

device capable of exploring a wide-breadth of NPP safety-relevant flows. The proof-of-

concept measurements and sensor design herein represent a significant milestone towards a

more straightforward, affordable and accessible means of capturing cross-sectional data at

high resolution at reactor operating conditions. The observed mixing phenomena in the

presence of strong density stratification should bring about a reconsideration of cross-flow

mixing classifications and awareness of new zones, specifically upstream of the T-junction,

where thermal fatigue may pay a role.

Page 161: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

144

8 Curriculum vitae

John Kickhofel

Date of Birth: October 20, 1986

Place of Birth: San Mateo, CA, U.S.A.

Nationality: American

Email: [email protected]

Education

02/2011 – 09/2015

08/2013 - 12/2013

08/2008 - 05/2010

07/2004 - 04/2008

Work Experience

06/2009 - 12/2009

05/2008 - 09/2008

Doctor of Science candidate

Laboratory of Nuclear Energy Systems, Institute of Energy Technology

Department of Mechanical and Process Engineering

ETH Zurich, Switzerland

Supervised by Prof. Dr. Horst-Michael Prasser

Diplôme d‘université in International Nuclear Law

Université Montpellier 1, France

Master of Science in Nuclear Engineering

Joint degree ETH Zurich - EPF Lausanne, Switzerland

Bachelor of Science in Physics

Georgia Institute of Technology, U.S.A.

Research, Engineering, Manufacturing and Sustaining (REMS) Intern

Schlumberger Oilfield Services, Sugar Land, TX, U.S.A.

Supervised by Dr. Dean Homan

Page 162: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

145

9 Publications

9.1 Patents

1. Kickhofel, J. and Prasser, H.-M. (2014), ‖Grid Sensor Package,‖ European Patent

Application EP14198142, ETH-Invention-No. 2014-053, Priority date December 16,

2014.

9.2 Peer-Reviewed Journal articles

1. Kickhofel, J. and Prasser, H.-M., ―Large Eddy Simulation of Turbulent Penetration in

a Horizontal Adiabatic T-junction with Leakage Flow,‖ Nuclear Engineering and

Design, in review, 2015.

2. Kickhofel, J. and Prasser, H.-M, ‖Turbulent Penetration in T-junction Branch Lines

with Leakage Flow,‖ Nuclear Engineering and Design 276, 43–53, 2014; doi.org/t4g.

3. Zboray, R., Kickhofel J., Damsohn, M., and Prasser, H.-M., "Cold-neutron

tomography of annular flow and functional spacer performance in a model of a

boiling water reactor fuel rod bundle," Nuclear Engineering and Design 241, 3201 –

3215, 2011; doi:10.1016/j.nucengdes.2011.06.029

4. Kickhofel, J., Zboray, R., Damsohn, M., Kaestner, A., Lehmann, E.-H., Prasser, H.-

M., ―Cold Neutron Tomography of Annular Coolant Flow in a Double Subchannel

Model of a Boiling Water Reactor,‖ Nuclear Instruments and Methods in Physics

Research Section A 651, 297-304, 2011; doi:10.1016/j.nima.2011.01.062

5. Kickhofel, J., Mohamide, A., Jalfin, J., Gibson, J., Thomas, P., Minerbo, G., Wang,

H., and Homan, D., ―Inductive Conductivity Tensor Measurements for Flowline or

Material Samples,‖ Review of Scientific Instruments 81, 075102, 2010;

doi:10.1063/1.3449320

9.3 Peer-Reviewed Conference papers

1. Kickhofel, J., Selvam, K., Huber, H., Laurien, E., and Prasser, H.-M., ―Mesh Sensor

for High Temperature and High Pressure Applications,― 16th International Topical

Meeting on Nuclear Reactor Thermalhydraulics (NURETH-16), Chicago, USA,

August 30 - September 4, 2015.

2. Kickhofel, J., Trinca, C., and Prasser, H.-M., ‖The Influence of Density Stratification

and T-junction Geometry on Turbulent Penetration,‖ International Topical Meeting on

Page 163: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

146

Nuclear Thermal Hydraulics, Operation and Safety (NUTHOS-10), Okinawa, Japan,

December 14-18, 2014.

3. Kickhofel, J. and Prasser, H.-M., ‖Large Eddy Simulation of Turbulent Penetration in

a T-junction,‖ International Congress on Advances in Nuclear Power Plants (ICAPP

2014), Charlotte, U.S.A., April 6-9, 2014.

4. Kickhofel, J., Valori, V., and Prasser, H.-M., ―Turbulent Penetration as a Thermal

Fatigue Problem in low Side Flow T-Junctions,‖ Nuclear Thermal Hydraulics and

Safety 8 (NTHAS8), Beppu, Japan, December 9-12, 2012.

5. Kickhofel, J., Fokken, J., Kappula, R., and Prasser, H.-M., ‖Steady State RANS

Simulations of Temperature Fluctuation in a Single Phase Turbulent Mixing,‖

International Congress on Advances in Nuclear Power Plants (ICAPP 2012), Chicago,

U.S.A., June 24-28, 2012.

Page 164: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

147

10 Bibliography

[1]. S. Chapuliot, C. Gourdin, T. Payen, J. P. Magnaud and A. Monavon, "Hydro-thermal-

mechanical analysis of thermal fatigue in a mixing tee," Nuclear Engineering and Design 235 (5), pp.

575-596 (2005).

[2]. J. Kickhofel, K. Selvam, H. Huber, L. Eckhart and H.-M. Prasser, "Wire Mesh Sensor for

High Temperature and High Pressure Applications," In: Proc. of the 16th International Topical

Meeting on Nuclear Reactor Thermalhydraulics, Chicago, USA, (2015).

[3]. E. Paffumi, V. Radu and K. F. Nilsson, "Thermal fatigue striping damage assessment from

simple screening criterion to spectrum loading approach," International Journal of Fatigue 53, pp. 92-

104 (2013).

[4]. E. Paffumi, K. F. Nilsson and N. G. Taylor, "Thermal frequency response studies of a hollow

cylinder subject to loads of different amplitude and shape," Nuclear Engineering and Design 240 (6),

pp. 1355-1362 (2010).

[5]. V. Radu, E. Paffumi, N. Taylor and K.-F. Nilsson, "Assessment of thermal fatigue crack

growth in the high cycle domain under sinusoidal thermal loading: An application – Civaux 1 case,"

2007.

[6]. V. Radu, N. Taylor and E. Paffumi, "Development of new analytical solutions for elastic

thermal stress components in a hollow cylinder under sinusoidal transient thermal loading,"

International Journal of Pressure Vessels and Piping 85, pp. 885–893 (2008).

[7]. V. Radu, E. Paffumi, N. Taylor and K.-F. Nilsson, "A study on fatigue crack growth in the

high cycle domain assuming sinusoidal thermal loading," International Journal of Pressure Vessels

and Piping 86, pp. 818–829 (2009).

[8]. S. Taheri, "Some Advances on Understanding of High Cycle Thermal Fatigue Crazing,"

Journal of Pressure Vessel Technology 129 (3), pp. 400 (2007).

[9]. A. Sabitti and S. Taheri, "Crack arrest in high cycle thermal fatigue crazing," Nuclear

Engineering and Design 240 (1), pp. 30-38 (2010).

[10]. E. Molinié, N. Monteil, S. Delatouche, S. Roux, N. Robert and C. Pages, " Caractérisation des

tronc¸on RRA 900-1300 Mwe deposés: synthèse des enseignements acquis," In: Proc. of the

International Symposium Fontevraud 5, France, (2002).

[11]. E. Paffumi, K.-F. Nilsson and N. G. Taylor, "Simulation of thermal fatigue damage in a 316L

model pipe component," International Journal of Pressure Vessels and Piping 85, pp. 798–813

(2008).

[12]. M. H. C. Hannink and F. J. Blom, "Numerical methods for the prediction of thermal fatigue

due to turbulent mixing," Nuclear Engineering and Design 241 (3), pp. 681-687 (2011).

[13]. S. Taheri, L. Vincent and J.-C. Le-roux, "A new model for fatigue damage accumulation of

austenitic stainless steel under variable amplitude loading," Procedia Engineering 66, pp. 575-586

(2013).

[14]. O. C. Garrido, S. E. Shawish and L. Cizelj, "A novel approach to generate random surface

thermal loads in piping," Nuclear Engineering and Design 273, pp. 98–109 (2014).

Page 165: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

148

[15]. W. Häfner and J. H. Spurk, "Formation of Waves in Thermal Stratified Flow," In: Proc. of the

International Conference on Physical Modeling of Transport and Dispersion, Cambridge, U.S.A.,

(1990).

[16]. V. K. Dhir, R. C. Amar and J. C. Mills, "A One Dimensional Model for the Prediction of

Stratification in Horizontal Pipes Subjected to Fluid Temperature Transient at Inlet," Nuclear

Engineering and Design 107, pp. 307-314 (1988).

[17]. IAEA, "Assessment and Management of Ageing of Major Nuclear Power Plant Components

Important to Safety: Primary piping in PWRs," IAEA IAEA-TECDOC-1361, Vienna, Austria, 2003.

[18]. M. Dahlberg, K.-F. Nilsson, N. Taylor, C. Faidy, U. Wilke, S. Chapuliot, D. Kalkhof, I.

Bretherton, M. Church, J. Solin and J. Catalano, ""Development of a European Procedure for

Assessment of High Cycle Thermal Fatigue in Light Water Reactors: Final Report of the NESC-

Thermal Fatigue Project ", 2007.

[19]. B. Lydell and J. Riznic, "OPDE—The international pipe failure data exchange project,"

Nuclear Engineering and Design 238 (8), pp. 2115-2123 (2008).

[20]. P. R. Huebotter, "Report of the National Task Force on Thermal Striping in LMFBRs," 1978.

[21]. J. E. Brunnings, "LMFBR thermal-striping evaluation," 1982.

[22]. IAEA, "Validation of fast reactor thermomechanical and thermohydraulic codes: Final report

of a co-ordinated research project 1996–1999," IAEA-TECDOC-1381, Vienna, Austria, 2002.

[23]. O. Gelineau, C. Escaravage, J. P. Simoneau and C. Faidy, "High Cycle Thermal Fatigue:

Experience and State of the Art in French LMFRs," In: Proc. of the Transactions of 16th International

Conference on Structural Mechanics in Reactor Technology (SMiRT-16), Washington DC, USA,

(2001).

[24]. A. L. Lund, "US NRC Regulatory perspective on unanticipated thermal fatigue in LWR

piping," In: Proc. of the Specialists Meeting on Experience with Thermal Fatigue in LWR Piping

Caused by Mixing and Stratification, Paris, France, (1998).

[25]. USNRC, "Bulletin 88-08: Supplement 1, Thermal Stresses in Piping Connected to Reactor

Coolant Systems," OMB No.: 3150-0011, Washington, D.C., U.S.A., 1988.

[26]. USNRC, "Bulletin 88-08: Supplement 2, Thermal Stresses in Piping Connected to Reactor

Coolant Systems," Washington, D.C., U.S.A., 1988.

[27]. USNRC, "Bulletin 88-08: Supplement 3, Thermal Stresses in Piping Connected to Reactor

Coolant Systems," Washington, D.C., U.S.A., 1989.

[28]. L. Wolf, W. Häfner, M. Geiss, E. Hansjosten and G. Katzenmeier, "Results of HDR-

experiments for pipe loads under thermally stratified flow conditions," Nuclear Engineering and

Design 137 (3), pp. 387-404 (1992).

[29]. A. Talja and E. Hansjosten, "Results of thermal stratification tests in a horizontal pipe line at

the HDR-facility," Nuclear Engineering and Design 118 (1), pp. 29-41 (1990).

[30]. L. Wolf, M. Geiss, E. Hansjosten and A. Talja, "Elvaluation of the Thermal Stratification

Tests at the HDR-Facility," In: Proc. of the 4th International Topical Meeting on Nuclear Reactor

Thermal Hydraulics (NURETH-4), Karlsruhe, Germany, (1989).

Page 166: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

149

[31]. A. Timperi, "Simulation of thermal striping in a HDR experiment. In: SAFIR2010. The

Finnish Research Programme on Nuclear Power Plant Safety 2007–2010, Interim Report. Puska, E.K.

(Ed.)." VTT Report No. VTT Research Notes 2466, Espoo, Finland, 2010.

[32]. H. C. Rezende, A. A. C. Santos, M. A. Navarro and E. Jordão, "Verification and Validation of

a thermal stratification experiment CFD simulation," Nuclear Engineering and Design 248, pp. 72-81

(2012).

[33]. H. C. Rezende, M. A. Navarro, E. Jordão, A. Augusto and C. d. Santos3, "Experiments on

One-phase Thermally Stratified Flows in Nuclear Reactor Pipe Lines " In: Proc. of the 2009

International Nuclear Atlantic Conference - INAC 2009, Rio de Janeiro, Brazil, (2009).

[34]. E. Jud, A. Crua and U. R. Blumer, "Experimental investigation of the stratified flow in the

horizontal pipework of nuclear reactors," Nuclear Engineering and Design 153, pp. 173-181 (1995).

[35]. S. Qiao, H. Gu, H. Wang, Y. Luo, D. Wang, P. Liu, Q. Wang and Q. Mao, "Experimental

investigation of thermal stratification in a pressurizer surge line," Annals of Nuclear Energy 73, pp.

211-217 (2014).

[36]. I. Boros and A. Aszódi, "Analysis of thermal stratification in the primary circuit of a VVER-

440 reactor with the CFX code," Nuclear Engineering and Design 238 (3), pp. 1265-1274 (2008).

[37]. Y. Sibamoto, T. Yonomoto, Y. Anoda and Y. Kukita, "Experimental investigation on

similarity between velocity and density profiles in density-stratifed countercurrent flow in reactor

horizontal leg," Nuclear Engineering and Design 201, pp. 83-98 (2000).

[38]. C. Faidy, "Thermal fatigue in French RHR system," In: Proc. of the Proceedings of

International conference on Fatigue of Reactor Components, Napa, California, USA, (2000).

[39]. J. Galpin and J. P. Simoneau, "Large Eddy Simulation of a thermal mixing tee in order to

assess the thermal fatigue," International Journal of Heat and Fluid Flow 32 (3), pp. 539-545 (2011).

[40]. C. Faidy, "High Cycle Thermal Fatigue: Lessons Learned From Civaux Event," In: Proc. of

the Materials Reliability Program: Second International Conference on Fatigue of Reactor

Components (MRP-84), Snowbird, Utah, (2002).

[41]. C. Faidy, "Thermal Fatigue in Mixing Tees: Status and Justification of French Assessment

Method," In: Proc. of the 12th International Conference on Nuclear Engineering (ICONE12),

Arlington, Virginia, U.S.A., (2004).

[42]. C. Faidy, "French Procedure for Ageing management program of safety components," In:

Proc. of the ASME Pressure Vessel & Piping Conference (PVP2003), Cleveland, Ohio, USA, (2003).

[43]. E. Blondet and C. Faidy, "High Cycle Thermal Fatigue in French PWR," In: Proc. of the 10th

International Conference on Nuclear Engineering (ICONE10), Arlington, VA, U.S.A., (2002).

[44]. A. Fissolo, C. Gourdin, O. Ancelet, S. Amiable, A. Demassieux, S. Chapuliot, N. Haddar, F.

Mermaz, J. M. Stelmaszyk and A. Constantinescu, "Crack initiation under thermal fatigue: An

overview of CEA experiencePart II (of II): Application of various criteria to biaxial thermal fatigue

tests and a first proposal to improve the estimation of the thermal fatigue damage," International

Journal of Fatigue 31 (7), pp. 1196-1210 (2009).

[45]. A. Fissolo, S. Amiable, O. Ancelet, F. Mermaz, J. Stelmaszyk, A. Constantinescu, C.

Robertson, L. Vincent, V. Maillot and F. Bouchet, "Crack initiation under thermal fatigue: An

Page 167: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

150

overview of CEA experience. Part I: Thermal fatigue appears to be more damaging than uniaxial

isothermal fatigue," International Journal of Fatigue 31 (3), pp. 587-600 (2009).

[46]. H. Kamide, M. Igarashi, S. Kawashima, N. Kimura and K. Hayashi, "Study on mixing

behavior in a tee piping and numerical analyses for evaluation of thermal striping," Nuclear

Engineering and Design 239 (1), pp. 58-67 (2009).

[47]. B. L. Smith, J. H. Mahaffy and K. Angele, "A CFD benchmarking exercise based on flow

mixing in a T-junction," Nuclear Engineering and Design 264, pp. 80-88 (2013).

[48]. A. Sakowitz, M. Mihaescu and L. Fuchs, "Effects of velocity ratio and inflow pulsations on

the flow in a T-junction by Large Eddy Simulation," Computers & Fluids 88, pp. 374–385 (2013).

[49]. V. S. Naik-Nimbalkar, A. W. Patwardhan, I. Banerjee, G. Padmakumar and G. Vaidyanathan,

"Thermal mixing in T-junctions," Chemical Engineering Science 65 (22), pp. 5901-5911 (2010).

[50]. C. Walker, M. Simiano, R. Zboray and H.-M. Prasser, "Investigations on mixing phenomena

in single-phase flowin a T-junction geometry," Nuclear Engineering and Design 239, pp. 116-126

(2009).

[51]. O. Braillard, P. Quemere and V. Lorch, "Thermal Fatigue in Mixing Tees Impacted by

Turbulent Flows at Large Gap of Temperature: The FATHER Experiment and Numerical

Simulation," In: Proc. of the 15th International Conference on Nuclear Engineering (ICONE15),

(2007).

[52]. O. Braillard and D. Edelin, "Advanced Experimental Tools Designed for the Assessment of

the Thermal Load Applied to the Mixing Tee and Nozzle Geometries in the PWR Plant," In: Proc. of

the First International Conference on Advancements in Nuclear Instrumentation Measurement

Methods and their Applications (ANIMMA), Marseille, France, (2009).

[53]. M. Igarashi, M. Tanaka, N. Kimura and H. Kamide, "Study on Fluid Mixing Phenomena for

Evaluation of Thermal Striping in a Mixing Tee," In: Proc. of the 10th International Topical Meeting

on Nuclear Reactor Thermal Hydraulics (NURETH-10), Seoul, Korea, (2003).

[54]. M. Igarashi, M. Tanaka, S. Kawashima and H. Kamide, "Experimental Study on Fluid Mixing

for Evaluation of Thermal Striping in T-pipe Junction," In: Proc. of the 10th International Conference

on Nuclear Engineering (ICONE10), Arlington, VA, USA, (2002).

[55]. H. Ogawa, M. Igarashi, N. Kimura and H. Kamide, "Experimental Study on Fluid Mixing

Phenomena in T-pipe Junction with Upstream Elbow," In: Proc. of the 11th International Topical

Meeting on Nuclear Reactor Thermal Hydraulics (NURETH-11), Avignon, France, (2005).

[56]. N. Kimura, H. Ogawa and H. Kamide, "Experimental study on fluid mixing phenomena in T-

pipe junction with upstream elbow," Nuclear Engineering and Design 240, pp. 3055–3066 (2010).

[57]. OECD/NEA, "Report of the OECD/NEA-Vattenfall T-Junction Benchmark exercise," 2011.

[58]. B. L. Smith, J. H. Mahaffy and K. Angele, "A CFD Benchmarking Exercise Based on Flow

Mixing in a T-Junction," In: Proc. of the 14th International Topical Meeting on Nuclear Reactor

Thermal Hydraulics (NURETH-14), Toronto, Ontario, Canada, (2011).

[59]. M. Kuschewski, R. Kulenovic and E. Laurien, "Experimental Setup for the Investigation of

Fluid-Structure Interactions in a T-junction," In: Proc. of the 14th International Topical Meeting on

Nuclear Reactor Thermal Hydraulics (NURETH-14), Toronto, Ontario, Canada, (2011).

Page 168: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

151

[60]. M. Kuschewski, R. Kulenovic and E. Laurien, "Experimental setup for the investigation of

fluid–structure interactions in a T-junction," Nuclear Engineering and Design 264, pp. 223-230

(2013).

[61]. M.-S. Chen, H.-E. Hsieh, Y.-M. Ferng and B.-S. Pei, "Experimental observations of thermal

mixing characteristics in T-junction piping," Nuclear Engineering and Design 276, pp. 107-114

(2014).

[62]. M.-S. Chen, Y.-M. Ferng, H.-E. Hsieh and B.-S. Pei, "Study of Thermal Mixing

Characteristics with T-Junction Piping System," In: Proc. of the ANS Annual Meeting, Chicago,

Illinois, USA, (2012).

[63]. Z.-Y. Zhang, Y.-M. Ferng, K.-H. Lin, M.-S. Chen and H.-E. Hsieh, "Study of Thermal

Mixing Characteristics in T-junction," In: Proc. of the International Congress on Advances in Nuclear

Power Plants (ICAPP), Nice, France, (2015).

[64]. S. Kuhn, O. Braillard, B. Ničeno and H.-M. Prasser, "Computational study of conjugate heat

transfer in T-junctions," Nuclear Engineering and Design 240 (6), pp. 1548-1557 (2010).

[65]. NUGENIA, http://s538600174.onlinehome.fr/nugenia/portfolio/mother-project/, Accessed:

27.10.2014.

[66]. R. Beaufils and S. Courtin, "Analysis of the Father Experiment with an Engineering Method

Devoted to High Cycle Thermal Fatigue," In: Proc. of the Proceedings of the ASME 2011 Pressure

Vessels & Piping Conference (PVP 2011), Baltimore, Maryland, USA, (2011).

[67]. S. Courtin, "High Cycle Thermal Fatigue Damage Prediction in Mixing Zones of Nuclear

Power Plants: Engineering Issues Illustrated on the FATHER Case," Procedia Engineering 66, pp.

240-249 (2013).

[68]. S. Courtin, J. A. L. Duff, B. Tacchini, A. Fissolo, L. Vincent, J. M. Stephan and R. Tampigny,

"High Cycle Thermal Fatigue FATHER Experiment: Non Destructive and Metallographic

Examinations," In: Proc. of the International Conference on Structural Mechanics in Reactor

Technology (SMiRT21), India, (2011).

[69]. S. Courtin and R. Beaufils, "Engineering Analyses on the High Cycle Thermal Fatigue

FATHER Case," In: Proc. of the International Conference on Structural Mechanics in Reactor

Technology (SMiRT21), India, (2011).

[70]. K.-J. Metzner and U.Wilke, "European THERFAT project—thermal fatigue evaluation of

piping system ―Tee‖-connections," Nuclear Engineering and Design 235, pp. 473-484 (2005).

[71]. "Assessment of Computational Fluid Dynamics (CFD) for Nuclear Reactor Safety Problems,

NEA/CSNI/R(2007)13," O. N. E. Agency, 2008.

[72]. JSME, "Guideline for Evaluation of High-Cycle Thermal Fatigue of a Pipe (in Japanese),"

JSME S017-2003, Tokyo, Japan, 2003.

[73]. D. K. A. KTA, "KTA 3201.1 Komponenten des Primärkreises von Leichtwasserreaktoren -

Teil 1: Werkstoffe und Erzeugnisformen," 1998-06.

[74]. J.-p. Simoneau, J. Champigny and O. Gelineau, "Applications of large eddy simulations in

nuclear field," Nuclear Engineering and Design 240 (2), pp. 429-439 (2010).

Page 169: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

152

[75]. G. Grötzbach and M. Wörner, "Direct numerical and large eddy simulations in nuclear

applications," International Journal of Heat and Fluid Flow 20, pp. 222-240 (1999).

[76]. M. Scheuerer, M. Heitsch, F. Menter, Y. Egorov, I. Toth, D. Bestion, S. Pigny, H. Paillere, A.

Martin, M. Boucker, E. Krepper, S. Willemsen, P. Muhlbauer, M. Andreani, B. Smith, R. Karlsson,

M. Henriksson, B. Hemstrom, I. Karppinen and G. Kimber, "Evaluation of computational fluid

dynamic methods for reactor safety analysis (ECORA)," Nuclear Engineering and Design 235 (2-4),

pp. 359-368 (2005).

[77]. L.-W. Hu and M. S. Kazimi, "LES benchmark study of high cycle temperature fluctuations

caused by thermal striping in a mixing tee," International Journal of Heat and Fluid Flow 27 (1), pp.

54-64 (2006).

[78]. T. Lu, D. Attinger and S. M. Liu, "Large-eddy simulations of velocity and temperature

fluctuations in hot and cold fluids mixing in a tee junction with an upstream straight or elbow main

pipe," Nuclear Engineering and Design 263, pp. 32-41 (2013).

[79]. T. Lu, S. M. Liu and D. Attinger, "Large-eddy simulations of structure effects of an upstream

elbow main pipe on hot and cold fluids mixing in a vertical tee junction," Annals of Nuclear Energy

60, pp. 420-431 (2013).

[80]. A. K. Kuczaj and E. M. J. Komen, "An Assessment of Large-eddy Simulation Toward

Thermal Fatigue Prediction," Nuclear Technology 170, (2010).

[81]. T. Ming and J. Zhao, "Large-eddy simulation of thermal fatigue in a mixing tee,"

International Journal of Heat and Fluid Flow 37, pp. 93-108 (2012).

[82]. D. Klören and E. Laurien, "Coupled Large-Eddy Simulation of Thermal Mixing in a T-

Junction," In: Proc. of the 14th International Topical Meeting on Nuclear Reactor Thermal Hydraulics

(NURETH-14), Toronto, Ontario, Canada, (2011).

[83]. CEA, http://www-cast3m.cea.fr/, Accessed: 27.10.2014.

[84]. T. Morii, Y. Onishi, K. Hirakawa and I. Nakamori, "CFD Simulations of a Turbulent Flow in

a T-Junction," In: Proc. of the 14th International Topical Meeting on Nuclear Reactor Thermal

Hydraulics, NURETH-14, Toronto, Ontario, Canada, (2011).

[85]. A. K. Kuczaj, E. M. J. Komen and M. S. Loginov, "Large-Eddy Simulation study of turbulent

mixing in a T-junction," Nuclear Engineering and Design 240 (9), pp. 2116-2122 (2010).

[86]. Y. Odemark, T. M. Green, K. Angele, J. Westin, F. Alavyoon and S. Lundström, "High-cycle

Thermal Fatigue in Mixing Tees: New Large-Eddy Simulations Validated Against New Data

Obtained by PIV in the Vattenfall Experiment," In: Proc. of the 17th International Conference on

Nuclear Engineering (ICONE17), Brussels, Belgium, (2009).

[87]. T. Frank, C. Lifante, H. M. Prasser and F. Menter, "Simulation of turbulent and thermal

mixing in T-junctions using URANS and scale-resolving turbulence models in ANSYS CFX,"

Nuclear Engineering and Design 240 (9), pp. 2313-2328 (2010).

[88]. D. Kloeren, M. Kuschewski and E. Laurien, "Large-Eddy Simulations of Stratified Flows in

Pipe Configurations Influenced by a Weld Seam," In: Proc. of the CFD for Nuclear Reactor Safety

Applications (CFD4NRS-4) Workshop, Daejeon, Korea, (2012).

Page 170: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

153

[89]. K. Selvam, R. Kulenovic and E. Laurien, "Large Eddy Simulation of Fluid Mixing at High

Temperature Differences in a T-junction Piping System," In: Proc. of the International Congress on

Advances in Nuclear Power Plants (ICAPP '14), Charlotte, U.S.A., (2014).

[90]. M. Tanaka, H. Ohshima and H. Monji, "Thermal Mixing in T-Junction Piping System Related

to High-Cycle Thermal Fatigue in Structure," Journal of Nuclear Science and Technology 47 (9), pp.

790–801 (2010).

[91]. T. Muramatsu, "Frequency Evaluation of Temperature Fluctuation Related to Thermal

Striping Phenomena Using a Direct Numerical Simulation Code DINUS-3," ASME PVP-C 253, pp.

111-121 (1993).

[92]. M. Tanaka and H. Ohshima, "Numerical Simulations of Thermal-mixing in T-junction Piping

System Using Large Eddy Simulation Approach," In: Proc. of the CFD4NRS-3, Washington DC,

USA, (2010).

[93]. T. Muramatsu and H. Ninokata, "Development of thermohydraulics computer programs for

thermal striping phenomena," Nuclear Technology 113 (1), pp. 54-71 (1996).

[94]. R. Sau and K. Mahesh, "Dynamics and mixing of vortex ring in crossflow," Journal of Fluid

Mechanics 604, pp. 384-409 (2008).

[95]. H. Hibara, T. Muramatsu, N. Hirata and K. Sudo, "Flows in T-junction piping system: 1st

report, flow characteristics and vortex street formed by branch pipe flow," Transactions of the JSME

(in Japanese) 70 (693), pp. 1192-1200 (2004).

[96]. M. Tanaka, H. Ohshima and S. Murakami, "Numerical investigation of thermal striping near

core instruments plate around control rod channels in JSFR," In: Proc. of the 13th International

Topical Meeting on Nuclear Reactor Thermal Hydraulics (NURETH-13), Kanazawa, Japan, (2009).

[97]. S. Qian and N. Kasahara, "Large Eddy Simulation Analysis of Fluid Temperature

Fluctuations at a T-junction for Prediction of Thermal Loading," Journal of Pressure Vessel

Technology 137, (2015).

[98]. S. Qian and N. Kasahara, "LES Analysis of Temperature Fluctuations at T-junctions for

Prediction of Thermal Loading," In: Proc. of the Proceedings of the ASME 2011 Pressure Vessels &

Piping Division Conference (PVP2011), Baltimore, Maryland, USA, (2011).

[99]. S. Qian, S. Kanamaru and N. Kasahara, "High-accuracy Analysis Methods of Fluid

Temperature Fluctuations at T-junctions for Thermal Fatigue Evaluation," In: Proc. of the

Proceedings of the ASME 2012 Pressure Vessels & Piping Conference (PVP2012), Toronto, Ontario,

Canada, (2012).

[100]. A. Nakamura, T. Oumaya and N. Takenaka, "Numerical Investigation of Thermal Striping at

a Mixing Tee Using Detached Eddy Simulation," In: Proc. of the 13th International Topical Meeting

on Nuclear Reactor Thermal Hydraulics (NURETH-13) Kanazawa City, Japan, (2009).

[101]. A. Nakamura, H. Ikeda, S. Qian, M. Tanaka and N. Kasahara, "Benchmark Simulation of

Temperature Fluctuation Using CFD for the Evaluation of the Thermal Load in a T-junction Pipe," In:

Proc. of the The Seventh Korea-Japan Symposium on Nuclear Thermal Hydraulics and Safety

(NTHAS7), Chuncheon, Korea, (2010).

[102]. K. Miyoshi, A. Nakamura and N. Takenaka, "Numerical Evaluation of Wall Temperature

Measurement Methods Developed to Investigate Thermal Fatigue at T-Junction Pipe," In: Proc. of the

Page 171: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

154

Proceedings of the 2012 20th International Conference on Nuclear Engineering collocated with the

ASME 2012 Power Conference, Anaheim, California, USA, (2012).

[103]. M. Kamaya and A. Nakamura, "Thermal stress analysis for fatigue damage evaluation at a

mixing tee," Nuclear Engineering and Design 241 (8), pp. 2674-2687 (2011).

[104]. Y. Utanohara, A. Nakamura and K. Miyoshi, "Numerical Simulation of Long-period Fluid

Temperature Fluctuation Downstream from a Mixing Tee," In: Proc. of the 10th International Topical

Meeting on Nuclear Thermal-Hydraulics, Operation and Safety (NUTHOS-10), Okinawa, Japan,

(2014).

[105]. M. Robert, "Corkscrew Flow Pattern In Piping System Dead Legs," In: Proc. of the 5th

International Topical Meeting on Nuclear Reactor Thermal Hydraulics (NURETH-5)

Salt Lake City, UT, USA, (1992).

[106]. M. Robert and J. D. Mattei, "Thermal-Hydraulic Phenomena in a Zero Flowrate Pipe

Connected to a High Flowrate, High Temperature Circuit," In: Proc. of the Engineering Turbulence

Modelling and Measurements, Dubrovnik, (1990).

[107]. E. Deutsch, P. Montanari and C. Mallez, "Isothermal study of the flow at the junction

between an auxiliary line and primary circuit of pressurised water reactors," Journal of Hydraulic

Research 35 (6), pp. 799-812 (1997).

[108]. J. H. Kim, A. F. Deardorff and R. M. Roidt, "Thermal stratification in nuclear reactor piping

system," In: Proc. of the 1st JSME/ASME joint international conference on nuclear engineering,

Tokyo, Japan, (1991).

[109]. USNRC, "Bulletin 88-08: Thermal Stresses in Piping Connected to Reactor Coolant

Systems," Washington, D.C., U.S.A., 1988.

[110]. J. H. Kim, R. M. Roidt and A. F. Deardorff, "Thermal stratification and reactor piping

integrity," Nuclear Engineering and Design 139, pp. 83-95 (1993).

[111]. K. Tanimoto, T. Shiraishi, S. Suzuki, H. Hirayama, K. Shiina, Y. Minami, H. Isaka, T.

Fukuda, J. Mizutani, S. Moriya and H. Madarame, "Study On Relationship Between Thermal

Stratification And Cavity Flow In A Branch Pipe With A Closed End," In: Proc. of the 10th

International Conference on Nuclear Engineering (ICONE 10), Arlington, VA, USA, (2002).

[112]. Y. Kondo, K. Tanimoto, T. Shiraishi, S. Suzuki, K. Ogura, K. Shiina, T. Fukuda, N. Chigusa

and S. Moriya, "Study on thermal stratification by cavity flow in a branch pipe with a closed end

(Concept of Lsh evaluation method)," Transactions of the Japan Society of Mechanical Engineers 70

(689), pp. 178-191 (2004).

[113]. K. Shiina, T. Kawamura, M. Ohtsuka, T. Mizuno, M. Hisatsune, K. Ogura, K. Tanimoto, T.

Fukuda, Y. Minami, S. Moriya and H. Madarame, "Prediction of Vortex Penetration Depth at

Thermal Stratification by Cavity Flow in a Branch Pipe with Closed End (Effect of Heat Radiation

Condition on Temperature Fluctuations)," Heat Transfer—Asian Research 36 (1), pp. 38-55 (2007).

[114]. Y. Kondo, K. Tanimoto, T. Shiraishi, S. Suzuki, H. Hirayama, K. Shiina, Y. Minami, H.

Isaka, T. Fukuda, J. Mizutani, S. Moriya and H. Madarame, "Relationship Between Thermal

Stratification And Cavity Flow In A Branch Pipe With A Closed End (Development of Evaluation

Method for Penetration Length of Cavity Flow in Branch Pipe Consisted of Vertical and Horizontal

Pipe)," In: Proc. of the 11th International Conference on Nuclear Engineering (ICONE 11), Tokyo,

Japan, (2003).

Page 172: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

155

[115]. K. Shiina, T. Kawamura, M. Ohtsuka, T. Mizuno, M. Hisatsune, K. Ogura, K. Tanimoto, T.

Fukuda, Y. Minami, S. Moriya and H. Madarame, "Prediction of vortex penetration depth on thermal

stratification by cavity flow in a branch pipe with a closed end (Effect of heat leak condition on

temperature fluctuation)," Transactions of the Japan Society of Mechanical Engineers 71 (703), pp.

954-961 (2005).

[116]. Y. Ishikawa, Y. Okuda and N. Kasahara, "Numerical Estimation Of Flow Penetrating Depth

Into Stagnant Branch Pipes For Thermal Fatigue Prediction," In: Proc. of the ASME 2014 Pressure

Vessels & Piping Conference (PVP 2014), Anaheim, California, USA, (2014).

[117]. A. Nakamura, K. Miyoshi, T. Oumaya, N. Takenaka, S. Hosokawa, D. Hamatani, M. Hase,

D. Onojima, Y. Yamamoto and A. Saito, "Temperature fluctuation phenomena in a normally stagnant

pipe connected downward to a high velocity and high temperature main pipe," Nuclear Engineering

and Design 269, pp. 360-373 (2014).

[118]. T. Shiraishi, Y. Nabika and S. Suzuki, "Experiments on a vortex in a circular pipe with a

closed end," Transactions of the Visualization Society of Japan 20 (2), pp. 93-96 (2000).

[119]. N. Nakamori, T. Suzuta, T. Ueno, J. Kasahara, K. Hanzawa, K. Oketani and O. Ukai,

"Thermal Stratification in Branch Pipe," In: Proc. of the 30th National Heat Transfer Symposium,

(1993).

[120]. M. Robert, J. H. Kim, R. M. Roidt and A. F. Deardorff, "Letter to the Editor and Authors'

Response," Nuclear Engineering and Design 144, pp. 517-419 (1993).

[121]. http://www.kepco.co.jp/pressre/1999/0525-1j.html, Accessed: December 3, 2014.

[122]. A. Nakamura, K. Miyoshi and N. Takenaka, "Evaluation of Penetration Length in a Closed

Straight Branch Pipe Using Numerical Simulations," INSS JOURNAL 19, pp. 39-49 (2012).

[123]. K. Miyoshi, A. Nakamura, T. Tokura, K. Sugimoto and N. Takenaka, "Flow and temperature

fluctuation mechanism in a downward branch pipe with a closed end (4th. report, Influence of inner

diameter on penetration depth and fluctuation characteristics)," Transactions of the JSME 80 (810),

(2014).

[124]. T. Iguchi, A. Saito, N. Takenaka, K. Miyoshi and A. Nakamura, "Experimental Study on

Penetration Length in a Small Size Branch Pipe in a Nuclear Power Plant," In: Proc. of the 14th

International Topical Meeting on Nuclear Reactor Thermal Hydraulics (NURETH-14), Toronto,

Ontario, Canada, (2011).

[125]. K. Miyoshi, A. Nakamura, T. Tokura, K. Sugimoto and N. Takenaka, "Flow and temperature

fluctuation mechanism in a downward branch pipe with a closed end (5th. report, Investigation of

fluctuation mechanism for penetration depth)," Transactions of the JSME 80 (817), (2014).

[126]. USNRC, S. Chu and B. McGill, "NRC Public Meeting Summary Report (Thermal Fatigue

Management Guideline for Normally Stagnant Non-Isolable RCS Branch Lines)," 2012.

[127]. EPRI, "Materials Reliability Program: EdF Thermal Fatigue Monitoring Experience on

Reactor Coolant System Auxiliary Lines (MRP - 69)," Materials Reliability Program Doc. #1003082,

Palo Alto, CA, U.S.A., 2002.

[128]. EPRI, "Materials Reliability Program: Operating Experience Regarding Thermal Fatigue of

Un-isolable Piping Connected to PWR Reactor Coolant Systems (MRP-25)," Materials Reliability

Program Doc. #1001006, Palo Alto, CA, U.S.A., 2000.

Page 173: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

156

[129]. EPRI, "Materials Reliability Program Interim Report on Thermal Cycling Model

Development for Representative Unisolable Piping Configurations (MRP-81)," Materials Reliability

Program Doc. #1003527, Palo Alto, CA, U.S.A., 2002.

[130]. J. D. Keller, A. J. Bilanin, A. E. Kaufman, D. B. Bliss, S. T. Rosinski and K. Wolfe, "Thermal

Cycling Modeling and Phenomenological Testing of Un-isolable Lines Attached to Primary Coolant

System Part 1: Up-Horizontal Configuration," In: Proc. of the Materials Reliability Program: Second

International Conference on Fatigue of Reactor Components (MRP-84), Snowbird, Utah, USA,

(2002).

[131]. J. D. Keller, A. J. Bilanin, A. E. Kaufman, D. B. Bliss, S. T. Rosinski and K. Wolfe, "Thermal

Cycling Modeling and Phenomenological Testing of Un-isolable Lines Attached to Primary Coolant

System Part 2: Down-Horizontal Configuration," In: Proc. of the Materials Reliability Program:

Second International Conference on Fatigue of Reactor Components (MRP-84), Snowbird, Utah,

USA, (2002).

[132]. J. D. Keller, A. J. Bilanin and S. T. Rosinski, "Modeling Swirl Penetration And Thermal

Cycling In Un-isolable Branch Lines Of Pwr Plants," In: Proc. of the 2003 ASME Pressure Vessels

and Piping Conference (PVP 2003), Cleveland, Ohio, USA, (2003).

[133]. EPRI, "Thermal Cycling Screening and Evaluation Model for Normally Stagnant Non-

Isolable Reactor Coolant Branch Line Piping with a Generic Application Assessment (MRP-132),"

Materials Reliability Program Doc. #1009552, Palo Alto, CA, U.S.A., 2004.

[134]. EPRI, "Materials Reliability Program: Development of a Thermal-Cycling Model for Un-

Isolable Branch Line Piping Configurations (MRP-97)," Materials Reliability Program Doc.

#1003209, Palo Alto, CA, U.S.A., 2003.

[135]. EPRI, "Materials Reliability Program: Addendum to MRP-97/MRP-132 Thermal Cycling

Screening and Evaluation Model (MRP-251)," Materials Reliability Program Doc. #1018357, Palo

Alto, CA, U.S.A., 2008.

[136]. EPRI, "Materials Reliability Program: Management of Thermal Fatigue in Normally Stagnant

Non-Isolable Reactor Coolant System Branch Lines (MRP-146)," Materials Reliability Program Doc.

#1011955, Palo Alto, CA, U.S.A., 2005.

[137]. EPRI, "Materials Reliability Program: Management of Thermal Fatigue in Normally Stagnant

Non-Isolable Reactor Coolant System Branch Lines - Supplemental Guidance (MRP-146S),"

Materials Reliability Program Doc. #1018330, Palo Alto, CA, U.S.A., 2008.

[138]. EPRI, "Materials Reliability Program: MRP-170 EPRI Thermal Fatigue Evaluation per MRP-

146, Version 1.0," Materials Reliability Program Doc. #1013270, Palo Alto, CA, U.S.A., 2006.

[139]. EPRI, "Materials Reliability Program: Management of Thermal Fatigue in Normally Stagnant

Non-Isolable Reactor Coolant System Branch Lines (MRP-146, Revision 1)," Materials Reliability

Program Doc. #1022564, Palo Alto, CA, U.S.A., 2011.

[140]. EPRI, "Materials Reliability Program: Operating Experience Regarding Thermal Fatigue of

Piping Connected to PWR Reactor Coolant Systems (MRP-85)," Palo Alto, CA, U.S.A., 2003.

[141]. R. Zboray and H.-M. Prasser, "On the relevance of low side flows for thermal loads in T-

junctions," Nuclear Engineering and Design 241 (8), pp. 2881-2888 (2011).

[142]. Y. A. Hassan, "Experimental and computational thermal-hydraulic results of fluid and

thermal mixing tests," Nuclear Engineering and Design 74, pp. 215-221 (1982).

Page 174: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

157

[143]. N. Nakamori, K. Hanzawa, K. Oketani, T. Ueno, J. Kasahara and S. Shirahama, "Research on

Thermal Stratification in Unisolable Piping of Reactor Coolant Pressure Boundary," In: Proc. of the

Specialists Meeting on Experience with Thermal Fagtigue in LWR Piping Causted by Mixing and

Stratification, Paris, France, (1998).

[144]. S.-N. Kim, S.-H. Hwang and K.-H. Yoon, "Experiments on the Thermal Stratification in the

Branch of NPP," Journal of Mechanical Science and Technology 19 (5), pp. 1206-1215 (2005).

[145]. J. Kickhofel, V. Valori and H.-M. Prasser, "Turbulent Penetration as a Thermal Fatigue

Problem in Low Side Flow T-Junctions," In: Proc. of the Eighth Japan-Korea Symposium on Nuclear

Thermal Hydraulics and Safety (NTHAS8), Beppu, Japan, (2012).

[146]. J. Kickhofel, V. Valori and H.-M. Prasser, "Turbulent penetration in T-junction branch lines

with leakage flow," Nuclear Engineering and Design 276, pp. 43-53 (2014).

[147]. J. Kickhofel, H.-M. Prasser and C. Trinca, "The Influence of Density Stratification and T-

junction Geometry on Turbulent Penetration," In: Proc. of the 10th International Topical Meeting on

Nuclear Thermal-Hydraulics, Operation and Safety (NUTHOS-10), Okinawa, (2014).

[148]. M. Nakamori and K. Hanzawa, "Valve Maintenance Guideline to Prevent Thermal Fatigue

Damage of the Reactor Pressure Boundary," In: Proc. of the Joint ASME/JSME Pressure Vessels and

Piping Conference, Honolulu, Hawaii, U.S.A., (1995).

[149]. H. Ikeda, K. Nakada and T. Murofushi, "Prediction of thermal stratification phenomena in a

branch pipe using numerical simulation," In: Proc. of the 15th International Conference on Nuclear

Engineering (ICONE-15), Nagoya, Japan, (2007).

[150]. T. Lu, W. W. Han and H. Zhai, "Numerical simulation of temperature fluctuation reduction

by a vortex breaker in an elbow pipe with thermal stratificatio," Annals of Nuclear Energy 75, pp.

462-467 (2015).

[151]. J. Kickhofel and H.-M. Prasser, "Large Eddy Simulation of Turbulent Penetration in a T-

junction," In: Proc. of the International Congress on Advances in Nuclear Power Plants (ICAPP),

Charlotte, USA, (2014).

[152]. T. Oumaya, A. Nakamura, N. Takenaka and D. Onojima, "Thermal Stress Evaluation of a

Closed Branch Pipe Connected to Reactor Coolant Loop," In: Proc. of the 14th International

Conference on Nuclear Engineering (ICONE14), Miami, Florida, U.S.A., (2006).

[153]. N. Takenaka, S. Hosokawa, M. Hase, D. Onojima, A. Nakamura and T. Oumaya,

"Investigation of Flow Structure And Temperature Fluctuation in a Closed Branch Pipe Connected to

High Velocity and High Temperature Flow in a Main Pipe," In: Proc. of the Fourth Japan-Korea

Symposium on Nuclear Thermal Hydraulics and Safety (NTHAS4), Sapporo, Japan, (2004).

[154]. S.-H. Kim, J.-B. Choi, J.-S. Park, Y.-H. Choi and J.-H. Lee, "A Coupled CFD-FEM Analysis

On The Safety Injection Piping Subjected To Thermal Stratification," Nuclear Engineering and

Technology 45 (2), pp. 237-248 (2013).

[155]. EPRI, "Materials Reliability Program: Management of Thermal Fatigue in Normally Stagnant

Non-Isolable Reactor Coolant System Branch Lines - Product Set," Materials Reliability Program

Doc. #E231389, Palo Alto, CA, U.S.A., 2009.

[156]. A. Sakowitz, M. Mihaescu and L. Fuchs, "Turbulent flow mechanisms in mixing T-junctions

by Large Eddy Simulations," International Journal of Heat and Fluid Flow 45, pp. 135–146 (2014).

Page 175: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

158

[157]. G. Brown and A. Roshko, "On density effects and large structure in turbulent mixing layers,"

J. Fluid Meeh. 64, pp. 775-816 (1974).

[158]. J. Fokken, R. Kapulla, G. Galgani, O. Schib and H.-M. Prasser, "Stably Stratified Isokinetic

Turbulent Mixing Layers Investigations in a square flow channel," In: Proc. of the International Youth

Nuclear Congress, Cape Town, South Africa, (2010).

[159]. E. C. Eggertson, R. Kapulla, J. Fokken and H.-M. Prasser, "Turbulent Mixing and its Effects

on Thermal Fatigue in Nuclear Reactors," World Academy of Science, Engineering and Technology 5

(4), pp. 181-188 (2011).

[160]. H.-M. Prasser, M. Damsohn, J. Fokken and S. Frey, "Novel fluid dynamic instrumentation for

mixing studies developed at ETH Zurich and PSI," In: Proc. of the 13th International Topical Meeting

on Nuclear Reactor Thermal Hydraulics (NURETH-13), Kanazawa City, Ishikawa Prefecture, Japan,

(2009).

[161]. F. P. D‘Aleo and H.-M. Prasser, "Design, calibration and testing of a thin film temperature

gauge array for temperature and heat flux measurements in fluid mixing experiments," Flow

Measurement and Instrumentation 24, pp. 29-35 (2012).

[162]. F. P. D‘Aleo and H.-M. Prasser, "Transient heat flux deduction for a slab of finite thickness

using surface temperature measurements," International Journal of Heat and Mass Transfer 60, pp.

616-623 (2013).

[163]. H.-M. Prasser, A. Böttger and J. Zschau, "A new electrode-mesh tomograph for gas–liquid

flow," Flow Measurement and Instrumentation 9, pp. 111-119 (1998).

[164]. V. Vitagliano and P. A. Lyons, "Diffusion Coefficients for Aqueous Solutions of Sodium

Chloride and Barium Chloride," Journal of the American Chemical Society 78 (8), pp. 1549-1552

(1956).

[165]. t. R. gmbH, Accessed: May 3.

[166]. U. Rohde, S. Kliem, B. Hemström, T. Toppila and Y. Bezrukov, "The European project

FLOMIX-R: Description of the slug mixing and buoyancy related experiments at the different test

facilities," Final report on WP 2, FZR-430, 2005.

[167]. F. Bulk, "An Experimental Study on Cross-Flow Mixing in a Rod-Bundle Geometry using a

Wire-Mesh," Delft University of Technology Master's Thesis, 2012.

[168]. I. E. Idel'chick, "Handbook of Hydraulic Resistance: Coefficients of Local Resistance and of

Friction," Published for the U.S. Atomic Energy Commission and the National Science Foundation,

Washington D.C. by the Israel Program for Scientific Translations, 1966.

[169]. M. Y. Gundogdu and M. O. Carprilioglu, "Present state of art on pulsating flow theory (part

2: turbulent flow regimes)," JSME International Journal 42, pp. 398-410 (1999).

[170]. G. J. Brereton and R. R. Mankbadi, "Review of recent advances in the study of unsteady

turbulent internal flows," Applied Mechanics Reviews 48, pp. 189-212 (1995).

[171]. J. R. Womersley, "Method for the calculation of velocity, rate of flow and viscous drag in

arteries when the pressure gradient is known," Journal of Physiology 127 (3), pp. 553-563 (1955).

[172]. S. Pope, Turbulent Flows. (Cambridge Unviersity Press, 2000).

Page 176: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

159

[173]. P. Sagaut, Large Eddy Simulation for Incompressible Flows (Third ed.). (Springer, 2006).

[174]. I. B. Celik, "Introductory Turbulence Modeling," Lectures Notes, West Virginia University

Mechanical & Aerospace Engineering Dept., 1999.

[175]. A. Leonard, "Energy cascade in large-eddy simulations of turbulent fluid flows," Advances in

Geophysics A 18, pp. 237-248 (1974).

[176]. F. G. Schmitt, "About Boussinesq‘s turbulent viscosity hypothesis: historical remarks and a

direct evaluation of its validity," C. R. Mecanique 335, pp. 617-627 (2007).

[177]. J. Ma, F. Wang and X. Tang, "Comparison of Several Subgrid-Scale Models for Large-Eddy

Simulation of Turbulent Flows in Water Turbine," In: Proc. of the 4th International Symposium on

Fluid Machinery and Fluid Engineering, Beijing, China, (2008).

[178]. F. R. Menter, "Best Practice: Scale-Resolving Simulations in ANSYS CFD (Version 1.0)," A.

G. GmbH, 2012.

[179]. F. Nicoud and F. Ducros, "Subgrid-Scale StressModelling Based on the Square of the

Velocity Gradient Tensor," Flow, Turbulence and Combustion 62 (3), pp. 183-200 (1999).

[180]. "Star-CCM+ 8.04.007-R8 Manual," CD-Adapco, 2013.

[181]. Y. Addad, U. Gaitonde, D. Laurence and S. Rolfo, in Quality and Reliability of Large-Eddy

Simulations, edited by J. Meyers, B. J. Geurts and P. Sagaut (2008).

[182]. S. T. Jayaraju, E. M. J. Komen and E. Baglietto, "Suitability of wall-functions in Large Eddy

Simulation for thermal fatigue in a T-junction," Nuclear Engineering and Design 240 (10), pp. 2544-

2554 (2010).

[183]. Rowley C., Colonius T. and B. A. T., "On self-sustained oscillations in two-dimensional

compressible flow over rectangular cavities," J. Fluid Mech. 455, pp. 315-346 (2002).

[184]. H. Weltens, H. Bressler, F. Terres, H. Neumaier and D. Rammoser, "Optimisation of

Catalytic Converter Gas Flow Distribution by CFD Prediction," In: Proc. of the SAE Technical Paper

(1993).

[185]. M. S. Bartlett, "Smoothing Periodograms from Time-Series with Continuous Spectra," Nature

161, pp. 686–687 (1948).

[186]. V. Petrov, T. D. Nguyen, D. Nunez, A. Dave and A. Manera, "High Resolution Experiments

of Velocity and Concentration Fluctuations in a Jet Flow," In: Proc. of the 15th International Topical

Meeting on Nuclear Reactor Thermal Hydraulics (NURETH-15), Chicago, U.S.A., (2015).

[187]. V. Valori, "T-junction mixing with low side flows," ETH Zürich Labor für

Kernenergiesysteme Master's Thesis, 2012.

[188]. L. Schmidt, "Experimental and Numerical Investigation of Turbulent Penetration Mixing in a

Low Side Branch Flow Rate," ETH Zürich Labor für Kernenergiesysteme Master's Thesis, 2013.

[189]. C. Trinca, "Thermal Loads Determination For Fatigue Assessment In Low Side Flow T-

junctions," ETH Zürich Labor für Kernenergiesysteme Master's Thesis, 2014.

Page 177: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

160

[190]. K. Umminger, T. Mull and B. Brand, "Integral Effect Tests in the PKL Facility with

International Participation," Nuclear Engineering and Technology 41 (6), (2009).

[191]. H. Purhonen, M. Puustinen, V. Riikonen, R. Kyrki-Rajamäki and J. Vihavainen, "PACTEL

integral test facility – Description of versatile applications," Annals of Nuclear Energy 33 (11-12),

(2006).

[192]. H. Nakamura, T. Watanabe, T. Takeda, Y. Maruyama and M. Suzuki, "Overview of Recent

Efforts Through ROSA/LSTF Experiments," Nuclear Engineering and Technology 41 (6), (2009).

[193]. J. Kickhofel and H.-M. Prasser, "Grid Sensor Package," European Patent Application

EP14198142, (2014).

[194]. M. Ashby and K. Johnson, Materials and Design (Second Edition) (Elsevier Ltd., 2010).

[195]. M. F. Ashby and D. R. H. Jones, Engineering Materials I: An Introduction to Properties,

Applications, and Design, Fourth Edition. (Elsevier Ltd., 2012).

[196]. H.-M. Prasser, A. Böttger and J. Zschau, "Two-Phase Flow Water Hammer Transients and

Induced Loads on Materials and Structures of Nuclear Power Plants; Deliverable

D13 Revision 2: Documentation of Advanced Two-phase Flow Instrumentation Adapted to PPP and

PMK-2 Facilities," 2003.

[197]. H.-M. Prasser, G. Ézsöl, G. Baranyai and T. Sühnel, "Spontaneous Water Hammers In A

Steam Line In Case Of Cold Water Ingress," Multiphase Science and Technology 20 (3-4), pp. 265-

289 (2008).

[198]. H. Pietruske and H.-M. Prasser, "Wire-mesh sensors for high-resolving two-phase flow

studies at high pressures and temperatures," Flow Measurement and Instrumentation 18 (2), pp. 87-94

(2007).

[199]. A. Dudlik, H.-M. Prasser, A. Apostolidis and A. Bergant, "Water hammer induced by fast

acting valves—experimental studies at Pilot Plant Pipework," Multiphase Science and Technology 20

(3-4), pp. 239–263 (2008).

[200]. H.-M. Prasser, A. Böttger and J. Zschau, "Entwicklung von Zweiphasenmeßtechnik für

vergleichende Untersuchungen zur Beschreibung von transienten Strömungen in Rohrleitungen:

Abschlußbericht zum Vorhaben 11ZF9504/1," 1998.

[201]. H.-M. Prasser, J. Zschau and A. Boettger, "Gittersensor zur Bestimmung der

Leitfähigkeitsverteilung in strömenden Medien sowie Verfahren zur Gewinnung der Meßsignale,"

DE19649011 C2, (2000).

[202]. A. Manera, H.-M. Prasser, T. H. J. J. van der Hagen, R. F. Mudde and W. J. M. de Kruijf, "A

comparison of void-fraction measurements during flashing-induced instabilities obtained with a wire-

mesh sensor and a gamma-transmission set-up," In: Proc. of the 4th International Conference on

Multiphase Flow (ICMF-2001), New Orleans, U.S.A., (2001).

[203]. OECD/NEA, https://www.oecd-nea.org/jointproj/pkl-2.html, Accessed: Feburary 27 2015.

[204]. Key to Steel - Stahlschlüssel. (Verlag Stahlschlüssel Wegst GmbH, Marbach, Germany,

2013).

[205]. F. E. GmbH, https://ceramics.ferrotec.com/products/ceramics/materials/machinable/nitride/,

Accessed: February 3.

Page 178: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

161

[206]. H.-M. Prasser, "Generalized Cross-correlation Technique For The Measurement Of Time-

dependent Velocities," In: Proc. of the 15th International Topical Meeting on Nuclear Reactor

Thermal Hydraulics (NURETH-15), Pisa, Italy, (2013).

[207]. G. Cox, "Gas velocity measurement in fires by the cross-correlation of random thermal

fluctuations—A comparison with conventional techniques," Combustion and Flame 28, pp. 155-163

(1977).

[208]. T. Betschart, D. Suckow, V. Brankov and H.-M. Prasser, "The Hydrodynamics Of Two-phase

Flow In A Tube Bundle And A Bare Channel," In: Proc. of the International Congress on Advances in

Nuclear Power Plants (ICAPP '14), Charlotte, U.S.A., (2014).

[209]. H.-M. Prasser, "Evolution of interfacial area concentration in a vertical air–water flow

measured by wire–mesh sensors," Nuclear Engineering and Design 237, pp. 1608-1617 (2007).

[210]. H.-M. Prasser, S. Stucki, T. Betschart and J. Eisenberg, "Interfacial Area Density

Measurement Using A Three-Layer Wire-mesh Sensor," In: Proc. of the 16th International Topical

Meeting on Nuclear Reactor Thermal Hydraulics (NURETH-16), Chicago, U.S.A., (2015).

[211]. H. Shaban and S. Tavoularis, "The wire-mesh sensor as a two-phase flow meter,"

Measurement Science and Technology 26 (1), pp. (2014).

[212]. M. J. D. Silva, E. Schleicher and U. Hampel, "Capacitance wire-mesh sensor for fast

measurement of phase fraction distributions," Measurement Science and Technology 18 (7), (2007).

[213]. M. J. D. Silva, S. Thiele, L. Abdulkareem, B. J. Azzopardi and U. Hampel, "High-resolution

gas–oil two-phase flow visualization with a capacitance wire-mesh sensor," Flow Measurement and

Instrumentation 21 (3), pp. 191-197 (2010).

[214]. F. E. Wang, W. J. Buehler and S. J. Pickart, "Crystal Structure and a Unique Martensitic

Transition of TiNi," Journal of Applied Physics 36 (10), pp. 3232-3229 (1965).

[215]. S. A. Thompson, "An overview of nickel–titanium alloys used in dentistry," International

Endodontic Journal, pp. 297-310 (2000).

[216]. E. Kassab, L. Neelakantan, M. Frotscher, S. Swaminathan, B. Maaß, M. Rohwerder, J.

Gomes and G. Eggeler, "Effect of ternary element addition on the corrosion behaviour of NiTi shape

memory alloys," Materials and Corrosion 65 (1), pp. 18-22 (2014).

[217]. Glen S. Bigelow, Santo A. Padula Ii, Anita Garg, Darrell Gaydosh and R. D. Noebe,

"Characterization of Ternary NiTiPd High-Temperature Shape-Memory Alloys under Load-Biased

Thermal Cycling," Metallurgical And Materials Transactions A 41A, pp. 3065-3079 (2010).

[218]. P. K. Selvam, R. Kulenovic and E. Laurien, "Large eddy simulation on thermal mixing of

fluids in a T-junction with conjugate heat transfer," Nuclear Engineering and Design 284, pp. 238-

246 (2015).

[219]. D. Klören and E. Laurien, "Large-Eddy Simulations of Stratified and Non-stratified T-

junction Mixing Flows," High Performance Computing in Science and Engineering „12: Transactions

of the High Performance Computing Center, Stuttgart (HLRS), pp. 311-324 (2013).

[220]. H. D. Kweon, J. S. Kim and K. Y. Lee, "Fatigue design of nuclear class 1 piping considering

thermal stratification," Nuclear Engineering and Design 238 (6), pp. 1265-1274 (2008).

Page 179: In Copyright - Non-Commercial Use Permitted Rights ... · unprecedented resolution, but also countercurrent stratification developing upstream at high Richardson numbers. The new

162

[221]. X. Schuler, K.-H. Herter, S. Moogk, E. Laurien, D. Klören, R. Kulenovic and M.

Kuschewski, "Thermal fatigue: Fluid-structure interaction at thermal mixing events," In: Proc. of the

38th MPA-Seminar Power Generation and Energy Efficiency - Materials and Component Behaviour,

Stuttgart, Germany, (2012).

[222]. L.-W. Hu, J. Lee, P. Saha and M. S. Kazimi, "Numerical simulation study of high thermal

fatigue caused by thermal stripping," In: Proc. of the Third International Conference on Fatigue of

Reactor Components, Seville, Spain, (2004).

[223]. USNRC, "Bulletin 88-08: Thermal Stresses in Piping Connected to Reactor Coolant

Systems," OMB No.: 3150-0011 Washington, D.C., U.S.A., 1988.

[224]. W. L. Marshall and E. U. Franck, "Ion product of water substance, 0–1000°C, 1–10,000 bars

New International Formulation and its background," Journal of Physical and Chemical Reference

Data 10 (2), pp. 295 (1981).

[225]. A. S. Quist and W. L. Marshall, "Electrical Conductances of Aqueous Sodium Chloride

Solutions from 0 to 800 Degrees and at Pressures to 4000 Bars," Journal of Physical Chemistry 72

(2), pp. 684-703 (1967).

[226]. G. R. Arce, Nonlinear Signal Processing: A Statistical Approach. (Wiley, New Jersey,

U.S.A., 2005).