surgical design for the fontan procedure using computational fluid

201
UC San Diego UC San Diego Electronic Theses and Dissertations Title Surgical design for the Fontan procedure using computational fluid dynamics and derivative- free optimization Permalink https://escholarship.org/uc/item/86b0t9qv Author Yang, Weiguang Publication Date 2012-01-01 Peer reviewed|Thesis/dissertation eScholarship.org Powered by the California Digital Library University of California

Upload: ngodan

Post on 11-Feb-2017

220 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Surgical design for the Fontan procedure using computational fluid

UC San DiegoUC San Diego Electronic Theses and Dissertations

TitleSurgical design for the Fontan procedure using computational fluid dynamics and derivative-free optimization

Permalinkhttps://escholarship.org/uc/item/86b0t9qv

AuthorYang, Weiguang

Publication Date2012-01-01 Peer reviewed|Thesis/dissertation

eScholarship.org Powered by the California Digital LibraryUniversity of California

Page 2: Surgical design for the Fontan procedure using computational fluid

UNIVERSITY OF CALIFORNIA, SAN DIEGO

Surgical Design for the Fontan Procedure Using Computational FluidDynamics and Derivative-free Optimization

A dissertation submitted in partial satisfaction of the

requirements for the degree

Doctor of Philosophy

in

Engineering Sciences (Mechanical Engineering)

by

Weiguang Yang

Committee in charge:

Professor Alison L. Marsden, ChairProfessor Yuri BazilevsProfessor Juan C. del AlamoProfessor Jeffrey A. FeinsteinProfessor Juan C. LasherasProfessor Andrew D. McCullochProfessor Beth J. Printz

2012

Page 3: Surgical design for the Fontan procedure using computational fluid

Copyright

Weiguang Yang, 2012

All rights reserved.

Page 4: Surgical design for the Fontan procedure using computational fluid

The dissertation of Weiguang Yang is approved, and it is

acceptable in quality and form for publication on micro-

film and electronically:

Chair

University of California, San Diego

2012

iii

Page 5: Surgical design for the Fontan procedure using computational fluid

DEDICATION

To my parents

iv

Page 6: Surgical design for the Fontan procedure using computational fluid

TABLE OF CONTENTS

Signature Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

Dedication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

Table of Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiv

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

Vita and Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxi

Abstract of the Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . xxii

Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Single ventricle heart defects (SVHD) and surgical palli-

ations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . 21.3 Magnetic resonance imaging (MRI) for single ventricle

heart defects . . . . . . . . . . . . . . . . . . . . . . . . . 51.4 Computational fluid dynamics for single ventricle heart

defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81.5 In-vitro flow experiments for single ventricle heart defects 121.6 A novel Y-graft and optimal design . . . . . . . . . . . . 151.7 Outline of the thesis . . . . . . . . . . . . . . . . . . . . 18

Chapter 2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202.1 Finite element methods (FEM) for blood flow problems . 21

2.1.1 FEM for convection-dominated flow . . . . . . . 212.1.2 Stabilized FEM for Navier-Stokes equations . . . 262.1.3 Boundary conditions . . . . . . . . . . . . . . . . 29

2.2 Surrogate management framework (SMF) . . . . . . . . . 342.2.1 Surrogate models . . . . . . . . . . . . . . . . . . 372.2.2 Mesh adaptive direct search (MADS) . . . . . . . 41

Chapter 3 Constrained Optimization of an Idealized Y-graft Model . . . 433.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.1.1 Model construction and parameterization . . . . . 443.1.2 Flow simulation and boundary conditions . . . . . 483.1.3 Unconstrained optimization . . . . . . . . . . . . 50

v

Page 7: Surgical design for the Fontan procedure using computational fluid

3.1.4 Polling strategies . . . . . . . . . . . . . . . . . . 513.1.5 Constrained optimization . . . . . . . . . . . . . . 533.1.6 Choice of cost function and constraints for Fontan

optimization . . . . . . . . . . . . . . . . . . . . . 553.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.2.1 Unconstrained optimization . . . . . . . . . . . . 593.2.2 Polling comparison . . . . . . . . . . . . . . . . . 663.2.3 Constrained optimization . . . . . . . . . . . . . . 68

3.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 713.4 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . 763.5 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . 77

Chapter 4 Hemodynamic Evaluations for traditional and Y-graft FontanGeometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . 794.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

4.1.1 Geometrical model construction . . . . . . . . . . 804.1.2 Flow simulation and boundary conditions . . . . . 834.1.3 Determination of performance parameters . . . . 84

4.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 874.2.1 Hepatic flow distribution . . . . . . . . . . . . . . 874.2.2 SVC pressure . . . . . . . . . . . . . . . . . . . . 894.2.3 Power loss . . . . . . . . . . . . . . . . . . . . . . 924.2.4 Wall Shear Stress . . . . . . . . . . . . . . . . . . 924.2.5 Averaged results . . . . . . . . . . . . . . . . . . . 94

4.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 954.3.1 Hepatic flow distribution . . . . . . . . . . . . . . 964.3.2 Power loss . . . . . . . . . . . . . . . . . . . . . . 984.3.3 SVC pressure . . . . . . . . . . . . . . . . . . . . 984.3.4 Wall shear stress . . . . . . . . . . . . . . . . . . 994.3.5 Ranking . . . . . . . . . . . . . . . . . . . . . . . 100

4.4 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . 1024.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . 1034.6 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . 104

Chapter 5 Y-graft optimal design for improved hepatic flow distribution . 1055.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

5.1.1 Geometrical model construction . . . . . . . . . . 1075.1.2 Flow simulation and boundary conditions . . . . . 1125.1.3 Optimization algorithm . . . . . . . . . . . . . . . 114

5.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1155.2.1 Idealized cases . . . . . . . . . . . . . . . . . . . . 1155.2.2 Patient-specific cases . . . . . . . . . . . . . . . . 119

5.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 124

vi

Page 8: Surgical design for the Fontan procedure using computational fluid

5.3.1 Idealized cases . . . . . . . . . . . . . . . . . . . . 1245.3.2 Patient-specific cases . . . . . . . . . . . . . . . . 1285.3.3 Technical considerations for Y-graft implantation 130

5.4 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . 1315.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . 1335.6 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . 134

Chapter 6 Simulations and validation for the first cohort of Y-graft Fontanpatients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1366.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

6.1.1 Surgical technique and clinical data acquisition . . 1376.1.2 Model construction . . . . . . . . . . . . . . . . . 1386.1.3 Flow simulation and boundary conditions . . . . . 140

6.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1426.2.1 Simulation vs. lung perfusion . . . . . . . . . . . 1426.2.2 Longitudinal HFD . . . . . . . . . . . . . . . . . 1436.2.3 HFD estimation without in vivo flow conditions . 1446.2.4 Thrombus investigation . . . . . . . . . . . . . . . 144

6.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 1486.4 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . 1556.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . 1566.6 Acknowledgments . . . . . . . . . . . . . . . . . . . . . . 157

Chapter 7 Conclusions and future work . . . . . . . . . . . . . . . . . . . 1587.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . 1587.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . 161

7.2.1 Pre-operative prediction and assessment . . . . . 1617.2.2 Patient specific optimal design . . . . . . . . . . . 1627.2.3 Validation against 4D MRI . . . . . . . . . . . . . 163

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165

vii

Page 9: Surgical design for the Fontan procedure using computational fluid

LIST OF FIGURES

Figure 1.1: Staged surgical palliations for SVHD. (a) Norwood procedure:a modified BT-shunt is created to channel aortic flow to thepulmonary arteries (PAs) (b) Glenn procedure: the superiorvena cava (SVC) is disconnected from the heart and reimplantedinto the PAs. (c) Fontan procedure: the inferior vena cava(IVC) is connected to the PAs via a lateral tunnel (LT) or anextracardiac (EC) Gore-Tex tube. Figures are reproduced withpermission from Gaca et al., Radiology, 2008;247:617-631. . . . 3

Figure 1.2: An illustration of PAVMs in a Glenn patient. This figure isreproduced with permission from Duncan et al., Annals of Tho-racic Surgery, 2003;76:1759-1766. . . . . . . . . . . . . . . . . 6

Figure 1.3: A sketch for two novel designs. (a) A dual-bifurcation designproposed by Soerensen et al.1 bifurcates the IVC and SVC flow.(b) A Y-shaped graft proposed by Marsden et al.2 bifurcatesthe IVC flow only. . . . . . . . . . . . . . . . . . . . . . . . . . 16

Figure 2.1: Solutions for the 1D steady convection-diffusion problem (2.1)with f = 0, g(0) = 0, g(10) = 1 using Galerkin, exact arti-ficial diffusion (EAD) and streamline-upwind-Petrov-Galerkin(SUPG) schemes. . . . . . . . . . . . . . . . . . . . . . . . . . 25

Figure 2.2: Solutions for the 1D steady convection-diffusion problem (2.1)with f = −16 a

10

(−2 + 4 x

10

), g(0) = 0, g(10) = 1 using Galerkin,

exact artificial diffusion (EAD) and streamline-upwind-Petrov-Galerkin (SUPG) schemes. . . . . . . . . . . . . . . . . . . . . 27

Figure 2.3: A spatial domain is divided into a 3D domain Ω modeled byNavier-Stokes equations and a downstream Ω′ modeled by lumpparameter models. The DtN outflow boundary conditions areprescribed on the boundary ΓB that separates Ω and Ω′. . . . 30

Figure 2.4: A three element Winkessel model. . . . . . . . . . . . . . . . . 33Figure 2.5: Flowchart of SMF using MADS. Search and poll steps are exe-

cuted alternately according to whether a design point that im-proves the current best cost function is found. . . . . . . . . . 36

Figure 3.1: Model parametrization showing the six design parameters usedfor shape optimization (a), and the resting pulsatile IVC andSVC flow waveforms used for inflow boundary conditions (b). . 46

Figure 3.2: Pulsatile waveforms for the rest, 2X and 3X exercise cases. Tosimulate two exercise levels, the IVC flow rate was increased by2 and 3 times; SVC flow was unchanged. . . . . . . . . . . . . 50

viii

Page 10: Surgical design for the Fontan procedure using computational fluid

Figure 3.3: The shape optimization procedure is made up of a series ofautomated sub-steps from model construction to the input ofthe cost function value into the optimization algorithm. . . . . 51

Figure 3.4: Example of a filter for the constrained optimization problem.The filter shown in (a) is improved when a dominating point isfound, producing the filter shown in (b). . . . . . . . . . . . . 56

Figure 3.5: Time-averaged shear stress magnitude (dynes/cm2) of the opti-mal shape over one respiratory cycle during the rest conditionusing pulsatile waveform. . . . . . . . . . . . . . . . . . . . . . 58

Figure 3.6: Unconstrained optimization results using steady inflow condi-tions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

Figure 3.7: Convergence history for the unconstrained optimization understeady inflow conditions. . . . . . . . . . . . . . . . . . . . . . . 61

Figure 3.8: Mean pressure for the optimal shapes under the rest, 2X and3X exercise levels using pulsatile waveforms. . . . . . . . . . . . 63

Figure 3.9: Instantaneous velocity magnitude on the centerline cut plane ofthe optimal shapes using pulsatile waveforms for the rest, 2X,and 3X cases with unconstrained optimization. . . . . . . . . . 64

Figure 3.10: Velocity vectors at peak IVC inflow for the exercise and restoptimal shapes at exercise conditions. Compared with the ex-ercise optimal design (left), the large graft of the rest optimaldesign (right) results in more flow separation and causes moreenergy loss. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

Figure 3.11: Convergence history for 5 LTMADS and 1 OrthoMADS in-stances with poll only under the 3X steady inflow condition. . . 67

Figure 3.12: Convergence history for 5 LTMADS and 1 OrthoMADS in-stances with search and poll together under the 3X steady inflowcondition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

Figure 3.13: Final results of the constrained optimization plotted as costfunction J vs. constraint function H for the rest case. Themodel in the upper left corner is the best feasible design andthe model in right bottom corner is the design with highestenergy efficiency. Differences in shape among these models showa strong effect of the WSS constraint for the rest case. . . . . . 70

Figure 3.14: Final results of the constrained optimization plotted as costfunction J vs. constraint function H for the 2X exercise case.The number of feasible points is increased to 11. The bestfeasible, least infeasible and highest energy efficiency models arelisted from left to right. Different points in the filter plot havesimilar geometry. Results for the 2X exercise case show that theeffect of the WSS constraint is weakened as inflow rates increase. 71

ix

Page 11: Surgical design for the Fontan procedure using computational fluid

Figure 3.15: Final results of the constrained optimization plotted as costfunction J vs. constraint function H for the 3X exercise case.The number of feasible points is increased to 32. . . . . . . . . 72

Figure 4.1: Original Glenn models and variations of Fontan geometries forfive patients. The Y-graft includes a 20 mm trunk and two 15mm branches. The size of the tube-shaped graft is 20 mm. Pa-tients B and E have a stenosis in the LPA and RPA, respectively,denoted by arrows. . . . . . . . . . . . . . . . . . . . . . . . . 82

Figure 4.2: Based on conservation of mass, we have QRPA = QIV C · x +QSV C · y and QLPA = QIV C · (1 − x) + QSV C · (1 − y), wherex is the fraction of hepatic flow going to the RPA, and y is thefraction of SVC flow going to the RPA. . . . . . . . . . . . . . 86

Figure 4.3: Visualization of the particle tracking in the model Y-graft IIfor patient B. Particle tracking is terminated when particles arewashed from the model. . . . . . . . . . . . . . . . . . . . . . . 87

Figure 4.4: Left: Hepatic flow distribution at rest. Right: Differences (per-centage of the IVC flow) from the theoretical optima for eachdesign at rest and exercise. Note that the theoretical optimafor patient A at rest, 2X and 3X are 61/39, 70/30 and 72/28,respectively, and that a 50/50 split can not be achieved in the-ory. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

Figure 4.5: Time-averaged velocity vectors in the Y-graft and T-junctionmodels for patients A, D and E. In the T-junction design forpatient A, the SVC jet blocks the hepatic flow entering the LPA.In patient D, most SVC flow is directed to the RPA due to acurved SVC. In patient E, Y-graft II improves the hepatic flowdistribution by having a straight proximal branch for the RPA,in which the SVC jet blocks hepatic flow going to the RPA fromthe right branch, compared to Y-graft I. . . . . . . . . . . . . . 91

Figure 4.6: Hepatic flow distribution changes with variations in pulmonaryflow split. Patients’ original pulmonary flow splits are markedby the arrows at the x axis. The table shows the averageddeviations with respect to the original hepatic flow distributionfor a 25% change in pulmonary flow split. . . . . . . . . . . . . 92

Figure 4.7: Contours of time-averaged WSS (dynes/cm2) at rest for patientsA and C. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

Figure 4.8: Averaged differences from the theoretical optima and powerlosses over five patients. The best performing of the Y-graftand offset designs for patients B and E are used. The differencesbetween the Y-graft and T-junction designs are statistically sig-nificant (∗P < 0.05). . . . . . . . . . . . . . . . . . . . . . . . . 95

x

Page 12: Surgical design for the Fontan procedure using computational fluid

Figure 5.1: Model parameterization and flared SVC anastomosis. Upperleft: Design parameters and centerlines of an idealized Y-graftFontan model. Upper right: A representative Y-graft model.Parameters DL and DR allow two branches to vary indepen-dently. Bottom left: An LPA-flared SVC anastomosis with astraight junction for the RPA side. Bottom right: A curved-to-LPA SVC anastomosis. . . . . . . . . . . . . . . . . . . . . . . 109

Figure 5.2: 1. A patient-specific Glenn model. 2. In the semi-idealizedGlenn model, the PA is approximated by uniform circular seg-mentations and the pulmonary artery branches are neglected.The PA diameter is equal to the averaged diameter of the patient-specific PA. 3. A Y-graft is implanted forming a semi-idealizedFontan model for the same patient. The design parameters forthe Y-graft areXL, XR, LIV C andDbranch. When large branchesare anastomosed, the segmentation at the anastomosis is en-larged to the graft size. Then the rest of the PA segmentationsare enlarged linearly according to the distance to the closestanastomosis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

Figure 5.3: Optimal values for the HFD. Based on Equation (4.2), the the-oretical optimum for the HFD, defined as the value closest to50/50, is determined given an inflow ratio QIV C

QSVCand a pul-

monary flow split FRPA (% inflow to RPA). . . . . . . . . . . . 114Figure 5.4: A comparison of HFD and energy loss for optimal unequal and

equal-sized branches. HFDs for the unequal and equal-sizedbranches are 63/37 and 65/35 (IVC-RPA/IVC-LPA), respec-tively, but equal-sized branches perform better in reducing en-ergy loss. The pulmonary flow split is 79/21 (RPA/LPA). . . . 117

Figure 5.5: Optimal Y-grafts with equal-sized branches for a large range ofpulmonary flow splits. Theoretical optima given by Equation4.2 are achieved by using optimization. The difference from thetheoretical value is shown at each point. . . . . . . . . . . . . 117

Figure 5.6: Time-averaged flow fields of optimal Y-grafts for a straight SVC-PA junction and two types of flared SVC anastomoses. Thepulmonary flow split is 55/45 (RPA/LPA). Compared to themodel with a straight SVC-PA junction, the optimal Y-graftsfor two flared SVC anastomoses have a more distal anastomosisfor the LPA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

xi

Page 13: Surgical design for the Fontan procedure using computational fluid

Figure 5.7: HFD vs. QIV C

Qinflowfor an idealized model and a patient-specific

model (patient B). Patient B’s original inflow ratio QIV C

Qinflowis

marked by an arrow. Total inflow is kept constant in this com-parison. The idealized Y-graft is optimized for an IVC inflow-to-total inflow ratio of 45%. There is only 1% change in theY-graft model when the ratio is altered. However, the patientspecific model is more sensitive to the change of IVC inflow-to-total inflow ratio. . . . . . . . . . . . . . . . . . . . . . . . . . . 119

Figure 5.8: a) Time-averaged velocity vector fields in the semi-idealizedand patient-specific models for patient A. b) Particle snapshotstaken at T=3s for the non-optimized and optimal models. Thebar chart shows the semi-idealized model (upper left) has a sim-ilar hepatic flow split to the patient-specific model (upper right)for the same optimal Y-graft, and that the optimized Y-graft im-proves the HFD by 79%, compared to the original non-optimizeddesign (lower left). The optimal and non-optimized branch sizesare 12.9 and 15 mm, respectively. . . . . . . . . . . . . . . . . . 121

Figure 5.9: Time-averaged velocity vector fields in the semi-idealized andpatient-specific models with and without the RUL for patientB. The Y-graft is optimized for a HFD of 50/50. Due to theeffect of the RUL, the optimized Y-graft skewed the hepatic flowby around 15% after it was implanted into the patient-specificmodel. When the RUL is excluded from the patient-specificmodel, the HFD is consistent with the idealized model prediction.122

Figure 5.10: a) Time-averaged velocity vector fields and HFD for patient B.The Y-graft in the semi-idealized model (upper left) is opti-mized for a hepatic flow split of 65/35 (RPA/LPA) to accountfor the overestimation of the RPA hepatic flow in the semi-idealized model. b) Particle snapshots taken at T=3s for thenon-optimized and optimal models. The bar chart shows theoptimal Y-graft improves the performance by 94% achieving aneven HFD in the patient-specific model (upper right), comparedto the non-optimized design (lower left). The optimal and non-optimized branch sizes are 16 and 15 mm, respectively. . . . . . 123

Figure 6.1: Post-operative MRI/CT images and models. Since patientsYF5 developed thrombus in the left branch, an unblocked Y-graft was reconstructed for study. . . . . . . . . . . . . . . . . 139

Figure 6.2: a) Comparison between early post-operative simulation-derivedHFD and lung perfusion data for patients YF1, YF2 and YF3.HFD in the early post-operative stage was quantified by simu-lation and lung perfusion. b) Changes in HFD from the earlyto six-month post-operative stages derived from simulations. . 143

xii

Page 14: Surgical design for the Fontan procedure using computational fluid

Figure 6.3: Time-averaged WSS magnitude for patients YF1, YF2 YF3 andYF5 in the early post-operative stage. YF5 R14-L10 and YF5R12-L12 are two modified Y-graft designs for patient YF5. Inthe baseline model for patient YF5, a distal anastomosis for theleft branch and a highly skewed pulmonary flow split resultedin larger low WSS area in the left branch. In model R12-L12,the WSS in the left branch was enhanced due to a proximalanastomosis that allowed SVC flow to wash the left branch. . . 147

Figure 6.4: Percentage of low WSS region for two branches in patients YF1,YF2, YF3 and YF5. For each threshold value τ , the low WSSarea relative to each branch surface was computed. . . . . . . 147

Figure 6.5: Percentage of low WSS region for patient YF5’s modified Y-grafts. In model R12-L12, the low WSS area in the left branchcan be effectively minimized by using a proximal anastomosisin which the SVC jet impinged the wall and the impact of theSVC jet on the WSS was reduced with increasing LPA flow.Compared to the baseline model (Figure 6.4), model R14-L10has a similar percentage of low WSS area for the left branchin the early post-operative stage for threshold values below 2dynes/cm2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

xiii

Page 15: Surgical design for the Fontan procedure using computational fluid

LIST OF TABLES

Table 3.1: Bounds on the design parameters for the idealized model. Nega-tive values for ΔR and ΔL indicate inward convex branches andpositive values denote outward concave branches. Bounds werechosen to be consistent with MRI data from a typical patient. . 47

Table 3.2: Values of the four constant geometric parameters used in modelconstruction, taken from MRI data of a typical Fontan patient. 47

Table 3.3: Mean flow rates, Re in the IVC and SVC and resistance dropsat rest and two levels of simulated exercise. . . . . . . . . . . . . 49

Table 3.4: Optimal parameters, cost function values and number of evalu-ations for the unconstrained optimization using different inflowconditions. Parameters that lie on the boundary are in bold. . . 64

Table 3.5: Comparison results for 5 LTMADS and 1 OrthoMADS instanceswith poll only under the 3X steady inflow condition. Parametersthat lie on the boundary are in bold. . . . . . . . . . . . . . . . 67

Table 3.6: Comparison results for 5 LTMADS and 1 OrthoMADS instanceswith search and poll together under the 3X steady inflow condi-tion. OrthoMADS found the best solution among 5 instances ofLTMADS, with relative precision 0.1%. Parameters that lie onthe boundary are in bold. . . . . . . . . . . . . . . . . . . . . . . 68

Table 3.7: Comparison of the best feasible parameter and the highest energyefficiency points for the constrained optimization. Parametersthat lie on the boundary are in bold. . . . . . . . . . . . . . . . 73

Table 4.1: MRI inflow rates, MRI outflow splits and the theoretical optimalhepatic flow splits (TOHFS) at rest for the five study patients. . 86

Table 4.2: Mean SVC pressure (mmHg), power loss (mW) and mean (inspace) WSS magnitude (dynes/cm2) on the IVC graft for theFontan models. Compared to the best Y-graft design for thesame patient, increases in power loss for the T-junction and offsetdesigns are also shown. . . . . . . . . . . . . . . . . . . . . . . . 93

Table 4.3: Ranking of energy loss and hepatic flow distribution for eachpatient. The ranking of the hepatic flow distribution is based onthe differences from the theoretical optima. In patients C andD, there are two designs tied for the hepatic flow distribution. . 101

xiv

Page 16: Surgical design for the Fontan procedure using computational fluid

Table 5.1: Bounds on the design parameters for the semi-idealized model.XR and XL are measured from the SVC-PA junction to the rightand left anastomosis points, respectively. Since the PA path is aparametric spline S(t), the anastomosis location can be changedby varying the spline parameter t. In our previous study,3 patientspecific models employed a uniform 20-15-15 mm Y-graft. Tooptimize the graft size, the branch diameter was allowed to varybetween 12 and 16 mm. . . . . . . . . . . . . . . . . . . . . . . 110

Table 5.2: Mean pulsatile inflow rates, pulmonary flow splits and pressure.A respiratory model4 was superimposed to the IVC flow acquiredfrom PCMRI for each patient following our previous work. Norespiratory model was added to the SVC input. The flow ratesused for the idealized model were taken from a typical Fontanpatient.4,5 We varied the RPA/LPA flow split in the idealizedmodel for different conditions and set a Fontan pressure (centralvenous pressure) of 12 mmHg. For patients A and B, pulmonaryflow splits and pressure data were taken from MRI and catheter-ization prior to the Fontan procedure. Transpulmonary gradient(TPG) is the mean pressure difference between the SVC and theleft atrium.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

Table 5.3: Geometric parameters and power loss for patient specific models.In patient A, the optimal XR was reduced resulting in a moreproximal anastomosis for the right branch. In patient B, the rightanastomosis is more distal while the left one is more proximalin the optimized model. In both cases, a smaller branch sizeresulted in more power loss. . . . . . . . . . . . . . . . . . . . . 122

Table 6.1: Patients’ flow conditions used in simulations. The vena cava flowand pulmonary flow split were measured by PC-MRI except forpatients YF4 and YF6, who had CT imaging. We use the for-mat, RPA/LPA, to present pulmonary flow split. For patientsYF1, YF2 and YF3, “early” and “6 month” denote measure-ments taken in the early (< 1 month) and 6 month post-operativestages, respectively. For patient YF5, pre-operative and 3 monthpost-operative measurements were performed. . . . . . . . . . . 141

xv

Page 17: Surgical design for the Fontan procedure using computational fluid

Table 6.2: Mean WSS magnitude for Y-graft branches and HFD. Comparedto other patients, patient YF5 had low WSS in the left branchin the early post operative stage but the WSS in the left branchincreased in the 3 month post-operative stage in which the pul-monary flow split changed from 81/19 to 54/46. The mean WSSfor patient YF3 is low due to a lower cardiac output. The mod-ified Y-grafts for patient YF5 increased mean WSS in the leftbranch in the early post-operative stage compared to the origi-nal Y-graft. All Y-graft designs for patient YF5 skewed hepaticflow to the RPA with PAVMs in the early post-operative stagebut the HFD was improved in the 3 month post-operative stage. 145

xvi

Page 18: Surgical design for the Fontan procedure using computational fluid

ACKNOWLEDGEMENTS

When I started writing the acknowledgements, I realized that words fail me

in expressing express my gratitude towards my adviser, Professor Alison Marsden.

I would not have been able to finish my Ph.D. study without her constant support

and guidance. I am grateful to Professor Marsden for giving me an opportunity to

work in a new area with great clinical impact in which CFD and optimization could

be used to save lives. Professor Marsden’s trust and patience created a unique

environment in the lab allowing me to explore the problems to my satisfaction.

Professor Marsden provided me with a platform where I can work with a group of

fantastic people and learn from them. Doing research is not always straightforward.

Professor Marsden’s encouragement and guidance helped me overcome a lot of

difficulties and made things that were hard a lot easier. I am honored to be her

student.

Dr. Jeffrey Feinstein is one of our clinical collaborators. Our research is

driven by needs from clinicians and patients. Without his collaboration and guid-

ance, this work would not have been possible. I appreciate the medical knowledge

and advice he offered. I acknowledge Professor Shawn Shadden, Dr. Irene Vignon-

Clementel, Professor Charles Audet, Professor Sebastien Le Digabel, Professor

John Dennis, Dr. Nathan Wilson and Professor Charles Taylor for sharing their

expertise in numerical simulation, optimization and modeling. The tools devel-

oped by them have been of great help and have been the core tools of my study.

xvii

Page 19: Surgical design for the Fontan procedure using computational fluid

I would also like to thank Dr. V. Mohan Reddy and Dr. Frandics Chan. With-

out Dr. Reddy’s masterly surgical skills, the Y-graft design would not have been

implemented. Dr. Chan’s MRI data were crucial for the patient specific study.

Through my stay at UCSD I have had the chance to learn from a lot of

people. I am thankful to all of them. In particular, I greatly appreciate the

advice and help from my other committee members, Professor Juan Lasheras,

Professor Juan Carlos del Alamo, Professor Yuri Bazilevs, Professor McCulloch

and Dr. Beth Printz. I was inspired by Professor Juan Lasheras’ broad knowledge

from mechanics to biology and his attitude to research. I benefit from Professors

Bazilevs and del Alamo’s courses on finite element methods and turbulence. In

addition, I wish to thank Professor Yaosong Chen at Peking University for his

advice and help.

For the people mentioned above who taugh me, guided me, inspired me,

and encouraged me, I would like to quote an old Chinese saying to express my

gratitude and respect —“be a teacher for one day, be a father for all life”.

I appreciate all the help I have received from my labmates: Dr. Sethu-

raman Sankaran, Dibyendu Sengupta, Mahdi Esmaily Moghadam, Chris Long,

Matt Bockman, Abhay Ramachandra, Dr. Ethan Kung and Jessica Oakes. It was

always fun and helpful to learn from them.

This work was supported by the American Heart Association, a Burroughs

Wellcome Fund Career Award at the Scientific Interface, a Leducq Foundation Net-

work of Excellence grant and a UCSD Kaplan dissertation fellowship. I appreciate

xviii

Page 20: Surgical design for the Fontan procedure using computational fluid

the funding from all these agencies, which made this work possible. This disser-

tation has resulted in the following papers that have been published, accepted

or is being prepared for publication. The dissertation author was the primary

investigator and author of these publications.

Chapter 3

Yang, W., Feinstein, J. A. and Marsden, A. L. Constrained Optimization of an

Idealized Y-shaped Baffle for the Fontan Surgery at Rest and Exercise. Comput.

Meth. Appl. Mech. Engrg. 2010;199:2135-2149.

Chapter 4

Yang, W., Vignon-Clementel, I. E., Troianowski, G., Reddy, V. M., Feinstein,

J. A. and Marsden, A. L. Hepatic blood flow distribution and performance in

traditional and Y-graft Fontan Geometries: A Case Series Computational Fluid

Dynamics Study. J. Thorac. Cardiovasc. Surg. 2012;143: 1086-1097.

Chapter 5

Yang, W., Feinstein, J. A., Shadden, S. C., Vignon-Clementel, I. E. and Marsden,

A. L. Optimization of a Y-graft Design for Improved Hepatic Flow Distribution in

the Fontan Circulation. J. Biomech. Engrg., accepted.

Chapter 6

xix

Page 21: Surgical design for the Fontan procedure using computational fluid

Yang, W., Chan, F. P., Feinstein, Reddy, V. M., Marsden, A. L., and Feinstein,

J. A. Flow Simulations and Validation for the First Cohort of Y-graft Fontan

Patients., in preparation.

In addition, I would like to appreciate the support and love from my family

members. My father Zhi Yang taught me to appreciate the beauty of engineering

since I was a child. Those interesting stories he told significantly influenced my

choice of major. I am grateful to my mother Qian Wei for raising and supporting

me. I apologize for not spending more time with her during nine years of college

and graduate school study. My grandparents Yongzhong Yang, Mingqiong Dai,

Peimin Wei and Sixia Cai, uncles Hong Yang, Xin Yang, Yi Wei and Xin Wei,

aunts Zhaoxia Sun, Su Zhao, Bo Qu and Jing Xie deserve my special thanks. I

would like to particularly thank my uncle Xin Yang and aunt Su Zhao for their

tremendous support during my graduate school. The soccer games with my uncle

Xin Yang on Saturday afternoon let me take a break from work and added a good

balance to my life.

Finally, I wish to thank my good friends Rong Jiang, Peng Wang, Matt

de Stadler, Zhangli Peng and On Shun Pak for those lively conversations and for

providing both technical and personal help time and again.

xx

Page 22: Surgical design for the Fontan procedure using computational fluid

VITA

2007 Bachelor of Engineering, Peking University, China.

2009 Master of Science, University of California, San Diego, USA.

2007-2012 Research Assistant, University of California, San Diego, USA.

2012 Doctor of Philosophy, University of California, San Diego,USA.

JOURNAL PUBLICATIONS

Yang, W., Feinstein, J. A. and Marsden, A. L. Constrained Optimization of anIdealized Y-shaped Baffle for the Fontan Surgery at Rest and Exercise. Comput.Meth. Appl. Mech. Engrg. 2010;199:2135-2149.

Yang, W., Vignon-Clementel, I. E., Troianowski, G., Reddy, V. M., Feinstein,J. A. and Marsden, A. L. Hepatic blood flow distribution and performance intraditional and Y-graft Fontan Geometries: A Case Series Computational FluidDynamics Study. J. Thorac. Cardiovasc. Surg. 2012;143: 1086-1097.

Yang, W., Feinstein, J. A., Shadden, S. C., Vignon-Clementel, I. E. and Marsden,A. L. Optimization of a Y-graft Design for Improved Hepatic Flow Distribution inthe Fontan Circulation. J. Biomech. Engrg., accepted.

Yang, W., Chan, F. P., Feinstein, Reddy, V. M., Marsden, A. L., and Feinstein,J. A. Flow Simulations and Validation for the First Cohort of Y-graft FontanPatients., in preparation.

AWARDS

Kaplan Dissertation Fellowship, University of California, San Diego (2011).

xxi

Page 23: Surgical design for the Fontan procedure using computational fluid

ABSTRACT OF THE DISSERTATION

Surgical Design for the Fontan Procedure Using Computational FluidDynamics and Derivative-free Optimization

by

Weiguang Yang

Doctor of Philosophy in Engineering Sciences (Mechanical Engineering)

University of California, San Diego, 2012

Professor Alison L. Marsden, Chair

Single ventricle heart defects are among the most serious forms of congenital

heart disease. For hypoplastic left heart syndrome (HLHS), a three-staged surgical

course, consisting of the Norwood, Glenn, and Fontan surgeries is performed. In

the extracardiac Fontan procedure, the inferior vena cava (IVC) is connected to the

PAs either via a Gore-Tex tube. Serious clinical challenges remain despite post-

operative survival rates upwards of 90%. A novel Y-shaped graft has been proposed

to replace current tube-shaped grafts showing promising preliminary results.

To refine the Y-graft design and further study the hemodynamic perfor-

mance of the Y-graft, a 3D time-dependent finite element flow solver was coupled

to a derivative free optimization algorithm using surrogate management framework

xxii

Page 24: Surgical design for the Fontan procedure using computational fluid

(SMF) and mesh adaptive direct search (MADS). In the first part of this disser-

tation, an idealized Y-graft model was parameterized and optimized for energy

efficiency. Constrained optimization with a wall shear stress (WSS) constraint was

performed in order to study the risk of thrombosis.

In the second part of this dissertation, patient specific Glenn models were

virtually converted into Fontan models by implanting Y- and tube-shaped grafts for

comparison. Particular attention was paid to the hepatic flow distribution (HFD),

a clinical parameter that plays an important role in the formation of pulmonary

arteriovenous malformations (PAVMs).

In a third study, we coupled Lagrangian particle tracking to an optimal

design framework to study the effects of boundary conditions and geometry on

HFD. Two patient-specific examples showed that optimization-derived Y-grafts

effectively improved HFD, compared to initial non-optimized designs.

Based on our preliminary simulation results, the Y-graft has been translated

into use in a clinical pilot study. Post-operative flow simulations showed good

agreement with the lung perfusion data measured in the clinic. The development

of thrombosis in one patient’s Y-graft was investigated from a hydrodynamic point

of view. Results suggested that low WSS area and flow stasis should be taken into

account in the surgical design for improved HFD.

To our knowledge, this is the first study to apply formal optimization to

the Fontan surgical design. Findings in this dissertation may provide guidelines

for the future Y-graft surgeries.

xxiii

Page 25: Surgical design for the Fontan procedure using computational fluid

Chapter 1

Introduction

1.1 Single ventricle heart defects (SVHD) and

surgical palliations

The American Heart Association’s statistics show that congenital heart

defects (CHD) are the number one cause (>24%) of death from birth defects.6

In 2004, the hospital cost for CHD was $2.6 billion.6 Single ventricle heart de-

fects including hypoplastic left heart syndrome (HLHS), pulmonary atresia/intact

ventricular septum and tricuspid atresia are among the most severe CHD malfor-

mations. The prevalence of HLHS is about 2.39 per 10,000 live birth.7 In patients

with such defects, inadequate blood flow to the lungs results in caynosis after birth.

The patients uniformly die without treatment. Usually a three-staged surgery is

performed. The first stage consists of establishing stable sources of aortic and pul-

1

Page 26: Surgical design for the Fontan procedure using computational fluid

2

monary blood flow, in a Norwood procedure or variant thereof. This procedure is

typically performed in the first week of life. In the second stage, the bidirectional

Glenn procedure, the superior vena cava (SVC) is disconnected from the heart and

reimplanted into the pulmonary arteries (PAs) at about 4-6 months of age. In

the third and final stage, the Fontan procedure, the inferior vena cava (IVC) is

connected to the PAs 2-4 years after the first stage. Therefore, the Fontan proce-

dure is also called the total cavopulmonary connection (TCPC). The first Fontan

procedure was performed by Fontan and Baudet for repairing tricuspid atresia.8

In the classic Fontan procedure, the SVC was connected to the right PA and the

IVC blood flow was directed to the left PA.8 Then, an intracardiac baffle (lateral

tunnel) formed by the sinus venarum and a prosthetic patch was used to channeled

the IVC flow to the PAs. Since the 1990s, a Gore-Tex conduit has been used widely

to connect the IVC to the PAs forming an extracardiac connection.9 Figure 1.1

illustrates the surgical palliations for SVHD.

1.2 Outcomes

Despite relatively high post-Fontan survival rates upwards of 90%,10 the

5-year survival rate drops to about 80% and the long-term outlook is still unsat-

isfactory. Complications post Fontan include diminished exercise capacity, pro-

tein losing enteropathy, arteriovenous malformations, thrombosis, arrhythmias,

and heart failure.10,11 It is shown that hemodynamic performace is closely re-

Page 27: Surgical design for the Fontan procedure using computational fluid

3

stage 1 stage 2

stage 3

BT-shunt SVC

PAsaorta

LT Fontan EC Fontan

Gore-tex tube

Figure 1.1: Staged surgical palliations for SVHD. (a) Norwood procedure: a mod-ified BT-shunt is created to channel aortic flow to the pulmonary arteries (PAs)(b) Glenn procedure: the superior vena cava (SVC) is disconnected from the heartand reimplanted into the PAs. (c) Fontan procedure: the inferior vena cava (IVC)is connected to the PAs via a lateral tunnel (LT) or an extracardiac (EC) Gore-Tex tube. Figures are reproduced with permission from Gaca et al., Radiology,2008;247:617-631.

lated to Fontan patients’ outcomes. Reduced exercise performance after Fontan

completion is well documented. Exercise capacity is associated with oxygen con-

sumption and stroke volume.12 In contrast to healthy people, Fontan patients can

only respond to exercise by increasing their heart rate because the stroke volume is

limited.13 Although the mechanism of limited exercise capacity has not been fully

elucidated, it has been shown that reductions in peak oxygen uptake and stroke

volume were observed in Fontan patients.14–16 Giardini et al.17 showed that con-

Page 28: Surgical design for the Fontan procedure using computational fluid

4

version of atriopulmonary Fontan to extracardiac Fontan improved patients’ peak

oxygen uptake. Since the blood in the Fontan circulation is passively pumped to

the lung by the pressure difference between the vena cava and the left atrium, the

vascular resistance plays an important role in the regulation of cardiac output.

Animal experiments by Guyton et al.18 demonstrated that a small increase in ve-

nous resistance resulted in a significant drop in the cardiac output for a modified

circulation bypassing the right ventricle. Sundareswaran et al.19 used a computa-

tional model to correlate the Fontan geometric resistance with cardic function and

suggested that optimizing the Fontan geometry may improve exercise capacity.

Therefore, it is hypothesized that improving TCPC resistance might lower central

venous pressure and improve patients’ quality of life.13

Thrombus formation is a significant issue causing morbidity and mortality.

Thrombus can lead to chronic pulmonary embolic disease, stroke, unbalanced per-

fusion, elevation of pulmonary vascular resistance and even death.20 The incidence

of thrombosis following Fontan completion can be as high as 20% to 30%.10 For

atriopulmonary and lateral tunnel connections, there is no difference in freedom

from thrombus.20 A major adverse outcome for the extracardiac Fontan proce-

dure is the inherent risk of thrombosis in the graft. Shirai et al.21 reported a

20% incidence of thrombus formation in the conduit. The formation of throm-

bus is multifactorial including flow stagnation, hypercoagulable state and atrial

arrhythmias.10 In a study of Alexi-Meskishvili et al.,22 two out of six patients with

oversized conduits developed thrombosis, indicating a possible correlation between

Page 29: Surgical design for the Fontan procedure using computational fluid

5

conduit size and thrombosis, though the optimal conduit size is still unclear.

Pulmonary arteriovenous malformations (PAVMs), characterized by abnor-

mal communication between the pulmonary arteries and pulmonary veins, are an

uncommon but serious complication, occurring in as many as 25% of patients with

superior cavopulmonary anastomosis due to the exclusion of hepatic blood flow

from the pulmonary circulation.23 In the lungs with PAVMs, the PAs are dilated

and proliferated (Figure 1.2). Pulmonary flow enters the pulmonary veins without

being oxygenated in the pulmonary capillary bed, forming a “short circuit” in the

pulmonary circulation. Consequently, patients with PAVMs may develop cyanosis,

congestive heart failure, and respiratory failure.24 Although the cause of PAVMs

is not fully understood, clinical evidence shows that the absence of a hepatic fac-

tor carried in the IVC blood is a likely cause. While the incidence of PAVMs

decreases after Fontan completion, they may still persist or develop due to skewed

hepatic flow distribution, particularly in patients with heterotaxy and interrupted

IVC.25–27 PAVMs can be resolved by surgical correction of uneven hepatic flow

distribution (HFD).26–28

1.3 Magnetic resonance imaging (MRI) for sin-

gle ventricle heart defects

Computed tomography (CT) and magnetic resonance imaging (MRI) are

two major non-invasive imaging tools used in day-to-day clinical practice. Com-

Page 30: Surgical design for the Fontan procedure using computational fluid

6

Figure 1.2: An illustration of PAVMs in a Glenn patient. This figure is reproducedwith permission from Duncan et al., Annals of Thoracic Surgery, 2003;76:1759-1766.

pared to MRI, CT provides finer resolution with much less scanning time but

patients are exposed to ironing radiation. It has been shown that the radiation

exposure from CT scans in childhood elevates cancer risk.29 In addition, hemody-

namic information which is not available for CT can be obtained by phase con-

trast MRI (PC MRI) techniques. Therefore, MRI is the preferred modality for

diagnosing and studying congenital heart diseases, and for use in simulation stud-

ies. According to Samyn,30 cardiac MRI provides: “(1) segmental description of

cardiac anomalies, (2) evaluation of thoracic aortic anomalies, (3) non-invasive de-

tection and quantification of shunts, stenoses, and regurgitation, (4) evaluation of

conotruncal malformations and complex anatomy, (5) identification of pulmonary

and systemic venous anomalies, and importantly, (6) post-operative study and

Page 31: Surgical design for the Fontan procedure using computational fluid

7

evaluation of adult congenital heart disease.”

Pulmonary and caval flow distribution is determined by the pulmonary vas-

cular resistance and ventricular function.31 Abnormal pulmonary flow distribution

has been linked to risk of development of PAVMs.32 Fogel et al. quantified the

caval flow contribution to lungs in ten LT Fontan patients by labeling IVC or

SVC flow with a presaturation pulse.33 They found that a nearly even pulmonary

flow split was common in ten patients, and that 67%±12% of IVC flow went to

the LPA. In contrast, a lung perfusion study by Seliem et al.31 showed that only

27% of Glenn patients immediately prior to the Fontan procedure had symmetric

pulmonary flow distribution and Houlind et al.34 found that the mean RPA-LPA

flow ratio is 1.5 in seven LT Fontan patients. Uneven pulmonary flow distribution

in Fontan patients indicates a disparity in pulmonary vascular resistances on the

left and right. Hager et al.35 assessed pulmonary flow patterns in the Fontan

circulation. Compared to healthy volunteers, Fontan patients exhibited highly

variable waveforms without a typical patterns except for a slight late diastolic

peak.35 However, respiratory effects were excluded in the flow waveforms in this

study. The Fontan circulation is respiratory driven because the SVC and IVC are

disconnected from the heart. Most MRI drived flow data were obtained during

breath hold. To investigate actual vena caval flow in Fontan patients, Hjortdal et

al.36 performed real-time MRI measurements on 11 TCPC patients. Compared to

SVC flow, IVC flow is significantly influenced by respiratory effects during rest and

exercise, reaching the highest flow rate during inspiration.36 This study provides

Page 32: Surgical design for the Fontan procedure using computational fluid

8

an important basis for simulating actual pulsatile flow in Fontan patients.

Compared to CFD, a unique advantage for MRI is the capability of measur-

ing velocity fields in vivo. Be’eri et al.37 managed to obtain a velocity vector field

on a plane that includes the Fontan pathway. Sundareswaran et al.38 extended a

single plane acquisition to multiple plane acquisitions and obtained a 3D velocity

field by using divergence-free interpolation techniques. Recently, Markl et al.39 ap-

plied state-of-the-art time-resolved 3D magnetic resonance velocity mapping (4D

MRI) to four Fontan patients. Complex in vivo flow fields were revealed. In all

patients, IVC flow was evenly perfused to two lungs with RPA-skewed pulmonary

flow splits.39 Since 4D MRI is still a new technique in the development stage,

quantifying small structures and highly dynamic flow patterns is limited.39 In

addition, quantification of wall shear stress by MRI is still challenging due to near-

wall resolution issues. However, these promising imaging techniques will facilitate

the study of Fontan hemodynamics together with computational fluid dynamics

(CFD).

1.4 Computational fluid dynamics for single ven-

tricle heart defects

With the advent of medical imaging processing, CFD for blood flow entered

an era of image-based simulation in the 1990s.40, 41 X-ray, magnetic resonance

and ultrasound imaging derived geometries allow one to better characterize local

Page 33: Surgical design for the Fontan procedure using computational fluid

9

hemodynamics. With PC-MRI, patients’ flow information can be incorporated into

simulations achieving a patient-specific modeling.42 Although most CFD studies

for blood flow focus on adult arterial vessels, there has been a growing interest in

modeling blood flow in children with congenital heart diseases. In the pioneering

work of Dubini et al.,43,44 blood flow in Fontan models with T-junction and offset

connections was simulated. Their studies revealed that the offset design reduced

energy loss compared to the T-junction design. Subsequently, surgeons adopted

their designs connecting the graft with an offset relative to the SVC in order to

reduce energy loss due to the caval flow collision. Following the work of Dubini et

al.,43 a series of numerical studies were carried out showing that the Fontan geom-

etry plays an important role in Fontan patients’ hemodynamic performance.2, 45–47

The impacts of geometric factors including the offset value, baffle size, PA diame-

ter and connection flaring on energy loss were emphasized in most studies.45, 47, 48

The PA size was found to be the strongest correlate for energy dissipation.48

The limitations of early work include the use of idealized models, steady

inflow conditions and focus on energy loss as the sole parameter of interest.43,47

Migliavacca and colleagues were the first to achieve a flow simulation with image

derived realistic models.45 Whitehead et al.46 studied power loss in ten patient spe-

cific models, but steady inflow conditions is not physiologically realistic and may

have resulted in inaccurate estimates of quantities of interest compared to un-

steady conditions.4,49 The pulsatility of caval flow in Fontan patients is reduced as

the IVC and SVC are disconnected from the right atrium. But respiratory effects

Page 34: Surgical design for the Fontan procedure using computational fluid

10

on the IVC flow are pronounced.36 Since the MRI acquisition is often cardiac-

gated and resulting measurements do not account for the effect of respiration,4,50

Marsden et al. introduced a respiratory model for pulsatile Fontan hemodynamic

simulations and showed nonnegligible differences in energy efficiency and pressure

drop. Later on, multiple physiologically relevant parameters were used to com-

prehensively evaluate patients’ hemodynamic performance.2, 4, 51 Results showed

that rankings of competing designs were sensitive to the choice of parameter under

consideration. For example, models with high energy efficiency can result in highly

uneven hepatic flow distribution and patients with low Fontan pressure.

Recently increasing attention has been paid to the quantification of HFD

due to a close connection with PAVMs. Bove and associates showed that the in-

tracardiac Fontan models resulted in more even HFD than extracardiac Fontan

models due to a better mixing of caval flow.52 Dasi et al.53 used patient specific

models with steady inflow conditions and showed that the HFD correlated with the

IVC offset for extracardiac models. Yang et al.3 compared the hemodynamic per-

formance of the T-junction, offset and Y-graft designs in multiple patient specific

models with an emphasis on HFD. It has been shown that overall the Y-graft de-

signs resulted in better HFD than traditional tube-shaped grafts but the geometry

and pulmonary flow splits significantly influence HFD.

Outflow boundary conditions are as important as inflow boundary condi-

tions in determining blood flow patterns. Applying proper boundary conditions

is essential to obtaining physiologic results. Vignon-Clementel et al. illustrated

Page 35: Surgical design for the Fontan procedure using computational fluid

11

the difference in pulmonary flow distribution caused by using zero pressure and

resistance boundary conditions.50 For pediatric hemodynamic simulations, it is

difficult to obtain outflow rates at multiple branches due to resolution issues and

difficulties in synchronizing the measured outflow data.50 Since constant pressure

at the outlets may alter the flow field and pressure distribution, time-dependent

and physiological boundary conditions are preferred.50 Recently, residence, RCR

(resistor-capacitor-resistor) and other lumped parameter boundary conditions that

use a circuit analogy to model the circulation have been incorporated into flow

simulations for single ventricle heart diseases.2, 4, 5, 54 Migliavacca and colleagues

coupled CFD into a closed-loop network such that a 3D flow simulation is per-

formed in the TCPC and 0D circuit simulations are performed for the rest of the

vascular system.54 This multiscale lumped parameter network provides a tool to

model the influences of local geometry and physiologic changes on global systemic

and cardiac parameters.

In most CFD simulations for SVHD, a rigid wall assumption is used.13

Orlando et al.55,56 first considered the effect of compliant walls on energy loss,

reporting a 10% increase. Bazilevs et al.57 applied the state of art fluid structure

interaction (FSI) to a complex patient specific Fontan model and showed that pres-

sure and wall shear stress were overpredicted in the simulation with rigid walls.

Recently, Long et al.58 performed Fontan FSI simulations with variable wall prop-

erties and showed that the differences in energy loss and hepatic flow distribution

were small. Thus, anatomically realistic models with a rigid wall assumption are

Page 36: Surgical design for the Fontan procedure using computational fluid

12

still adequate to evaluate Fontan patients’ hemodynamic performance.

1.5 In-vitro flow experiments for single ventricle

heart defects

In-vitro flow experiments for modeling the Fontan circulation began ear-

lier than numerical flow simulations. de Laval and colleagues’ pioneering in-vitro

hydrodynamic study dates back to 1988.59 They found that a valveless chamber

caused turbulence and increased resistance to steady flow. Their in-vitro exper-

iments supported the replacement of the atriopulmonary connection by the new

TCPC surgical methods. To our knowledge, this was the first study that used

fluid mechanics to improve the Fontan surgical procedure. Later on, in-vitro ex-

periments of Low et al.60 showed that the lateral tunnel TCPC reduced energy loss

compared to the atriopulmonary connection. Since the lateral tunnel and extrac-

ardiac connections gained popularity in the 1990s, the atriopulmonary connection

was less well studied in in-vitro experimental studies for the Fontan procedure.

Sharma et al.61 studied the effect of caval offset and varying pulmonary flow split

on energy loss using idealized glass models. They concluded that 1 to 1.5 diameter

offset with an even pulmonary flow split resulted in lower energy loss.61 Based

on the study of Sharma et al.,61 idealized models with different vessel sizes and

anastomoses were tested in a similar manner.62, 63 Lardo et al.64 performed the

Fontan procedure with three variants (intracardiac lateral tunnel, extracardiac lat-

Page 37: Surgical design for the Fontan procedure using computational fluid

13

eral tunnel and extracardiac conduit) on fresh explanted sheep heart preparations

and compared energy efficiency in an in-vitro flow loop. The extracardiac conduit

was shown to have higher energy efficiency than two tunnel configurations and a

further reduction of 36% in energy loss was observed in the offset configuration.64

Compared to idealized models, a patient specific model can result in more complex

and unsteady flow structures even with steady inflow conditions.65

In addition, in-vitro flow experiments are valuable to validate numerical

simulations. Ryu et al.47 compared CFD solutions to in-vitro experiments show-

ing CFD gave similar flow field and power loss values in an idealized model. Khu-

natorn et al.66 further compared velocity fields in idealized TCPC models with

steady inflow conditions. The differences between PIV and CFD for the mean

axial velocity were within 20% but significant differences were shown in the sec-

ondary flow patterns. Pekkan and coworkers compared CFD derived power loss

to PIV data for a patient specific model showing close agreement under steady

inflow conditions.65,67 The discrepancy increased with increasing cardiac output

indicating a challenge for the flow solver to model turbulence transition. However,

the Reynolds number in the Fontan circulation is usually lower than 2000 in most

cases.67

While most in-vitro flow studies focused on energy loss, Walker et al.68,69

investigated the effects of connection geometry on hepatic flow distribution by

injecting dye into the IVC flow. It has been shown that the straight connection

resulted in relatively less sensitive hepatic flow distribution to the pulmonary flow

Page 38: Surgical design for the Fontan procedure using computational fluid

14

split because of a better mixing though it dissipated more energy compared to

the offset, flared and curved designs. However the actual hepatic flow distribution

in the T-junction design can vary significantly from patient to patient due to the

effects of non-idealized geometries and uneven pulmonary flow splits.3

Steady inflow conditions were used in the in-vitro studies mentioned above.

Although useful conclusions may still be drawn from steady flow experiments, the

flow field is altered. Therefore, besides realistic anatomic geometry, physiologic

flow conditions are needed for in-vitro flow experiments to study Fontan hemo-

dynamics. Figliola and colleagues have developed mock flow circuits that allow

one to simulate the Fontan circulation with physiologic realism.70 In mock flow

circuits, resistance, compliance and inertance are implemented by capillary tubing

or honeycomb matrix, variable air volume chambers, and lengths of flow tubing,

respectively.70 Similar to multiscale numerical modeling for the Fontan circula-

tion,54 upper body, hepatic and lower body compartments and the two pulmonary

branches were modeled by mock flow circuits, and the TCPC test section consists

of a MRI-derived Fontan model.70 This system is particularly useful to duplicate

clinical reports, perform patient specific in-vitro experiments and validate numer-

ical simulations.

Page 39: Surgical design for the Fontan procedure using computational fluid

15

1.6 A novel Y-graft and optimal design

In the previous sections, we have shown that the Fontan geometry plays

an important role in the hemodynamic performance, and that fluid mechanics

promoted the surgical community’s shift from the atriopulmonary connection to

TCPC and from the extracardiac T-junction to the offset design. Although sur-

gical modifications with other advances in pediatric cardiology and intensive care

medicine greatly improved early postoperative survival rates, clinicians and engi-

neers have not stopped pursuing better surgical designs for Fontan patients.10,13

In recent years, a Y-shaped graft has been proposed by two research groups. So-

erensen et al.1 introduced a dual-bifurcation design that bifurcates the SVC and

IVC flow (Figure 1.3a). Steady simulations with an idealized model showed that

energy dissipation was improved by reducing direct flow collision.1 However, this

design not only introduces extra artificial materials but also increases the surgical

technical difficulty. Therefore, no further study on the dual bifurcation design has

been made. Marsden et al.2 used a Y-shaped graft to replace the tube-shaped

graft for the IVC flow only (Figure 1.32). A patient specific model shows promis-

ing results , demonstrating that the Y-graft design improved the energy efficiency,

SVC pressure and HFD, compared to the tube-shaped graft.

Although the preliminary results for the Y-graft design support the re-

placement of the T-junction with the Y-graft design, several problems were still

unclear and further studies were needed to refine the design before translating the

Page 40: Surgical design for the Fontan procedure using computational fluid

16

SVC

IVC

RPA LPA

SVC

IVC

RPA LPA

a)

b)

Figure 1.3: A sketch for two novel designs. (a) A dual-bifurcation design proposedby Soerensen et al.1 bifurcates the IVC and SVC flow. (b) A Y-shaped graftproposed by Marsden et al.2 bifurcates the IVC flow only.

Y-graft into clinical use. Does the Y-graft outperform the traditional designs in

all patients with a variety of anatomic structures and flow conditions? What is

the optimal design for the Y-graft? Does a one-size-fits-all designs exist? Which

metrics should we use to evaluate the hemodynamic performance of the Y-graft

design? The studies in this thesis are motivated by these questions.

Optimal design has been used widely in a series of engineering problems.

In structural mechanics, one may want to know an optimal material distribution

to maximize the stiffness of a structure.71 In a thermal insulation system, one may

Page 41: Surgical design for the Fontan procedure using computational fluid

17

want to design a heat intercepts with minimum heat flow flux.72 In operations

research, a famous optimization example is the traveling salesman problem that

finds the shortest route to visit all cities once and return to the original. In

fluid mechanics, shape optimization is employed in harbor design, microfluidic

technology, combustion and other applications.73, 74 Aircraft design is one of the

most important research activities in shape optimization for fluid mechanics, which

is driven by considerable benefits brought by optimization.75

Optimization techniques can be categorized into two groups: gradient-based

and derivative-free algorithms. In gradient-based methods, derivatives are usually

obtained by adjoint or finite difference methods.76 Since the cost of obtaining gradi-

ent information by using adjoint methods is independent of the number of design

variables, it is widely used in aerodynamic design, as pioneered by Jameson.77

However, several difficulties limit the application of adjoint methods. First, ad-

joint methods are difficult to implement and not portable to different solvers. Sec-

ond it is challenging to apply adjoint methods to time-accurate problems. Third,

noisy or unavailable derivatives make adjoint methods inapplicable.76 In contrast,

derivative-free methods including pattern search, response surfaces and evolution-

ary algorithms are non-intrusive and suitable for non-differentiable cost functions

at a price of increased cost. The surrogate management framework (SMF) has

been successfully applied to an expensive trailing-edge noise reduction problem by

Marsden et al..78–80 Despite large applications of optimization in engineering prob-

lems, shape optimal design is still relatively new to bioengineering, particularly to

Page 42: Surgical design for the Fontan procedure using computational fluid

18

flow simulation based surgical design in which trial and error methods are used.

To our knowledge, SMF had not been used in bioengineering problems previously

before Marsden and colleagues applied it to a few idealized biofluid problems.76

Since there has been considerable work in the development of shape optimization

tools for engineering fluid mechanics, we believe that formal optimization tools

which systematically sort out optimal solutions from numerous candidates can fa-

cilitate discovery of new surgical designs and patients will eventually benefit from

the applications of optimization.

1.7 Outline of the thesis

The work presented in this thesis evaluated the hemodynamic performance

of the Y-gaft Fontan design and identified a series of optimal designs under different

conditions. Chapter 2 briefly introduces the methods used in this work. Chapter

3 presents a constrained optimization study of an idealized Fontan model. Energy

efficiency is chosen as the objective function. The optimal shapes for steady and

unsteady flow conditions were identified. A trade-off relationship between the en-

ergy efficiency and low WSS area was revealed. Chapter 4 extends idealized models

to patient specific models. Multiple patient specific models with T-junction, offset

and Y-graft designs were studied with particular attention paid to the HFD. In

Chapter 5, we applied optimization techniques to study the necessity of unequal-

sized branches for the Y-graft design and the influences of geometric characteristics

Page 43: Surgical design for the Fontan procedure using computational fluid

19

and boundary conditions on HFD using an idealized model. Two underperform-

ing patient specific Y-graft Fontan models were optimized to achieve even HFD.

Chapter 6 presents post-operative hemodynamics evaluations for the first cohort

of the Y-graft Fontan patients and in-vivo validation results. One adverse event,

in which a patient developed thrombosis post Fontan, was studied providing data

and principles for future Fontan surgical design. Finally, the major results of this

work are summarized and related future work is discussed in Chapter 7.

Page 44: Surgical design for the Fontan procedure using computational fluid

Chapter 2

Methods

In this chapter, we briefly introduce the numerical methods used in this

work. Finite element methods (FEM) were used to simulate blood flow and the

derivative-free surrogate management framework (SMF) together with mesh adap-

tive direct search (MADS) was used to identify optimal solutions. Although the

literature for FEM and optimization is vast, fundamental concepts are introduced

here for the sake of completeness. The contents of this chapter are primarily drawn

from the textbooks and studies of Hughes et al.,81, 82 Zienkiewicz et al.,83 Donea et

al.,84 Vignon-Clementel et al.85 and Audet et al.86

20

Page 45: Surgical design for the Fontan procedure using computational fluid

21

2.1 Finite element methods (FEM) for blood flow

problems

2.1.1 FEM for convection-dominated flow

The concept of finite elements originated from the work by Hrennikoff (1941)

and Courant (1942).87 Argyris, Clough and Zienkiewicz were the pioneers who de-

veloped FEM for structural analysis in the 1950s-1960s.88 The rigorous mathemat-

ical foundation was laid by Strang and Fix in the 1970s.89 It is worth mentioning

that Kang Feng, a Chinese mathematician, independently developed a FEM the-

ory so-called “Finite difference method based on variation principle” in 1965 in

parallel with the developments in the West90 and made original contributions to

the natural integral operator for the natural boundary reduction, which is also

known as the Dirichlet-to-Neumann (DtN) map.91–93

Unlike the finite difference method, FEM starts with a weak or variational

form that is converted from the strong form or classical form of the governing

equations using the principles of virtual work. After obtaining an equivalent weak

form of the problem, the weak form can be converted to a system of equations

and then discretized into a matrix form which can be solved numerically. Al-

though FEM achieved great success in solid mechanics, it historically encountered

some difficulties in convection-dominated flow problems. Let us use a 1D steady

convection-diffusion equation to illustrate the difficulties caused by using standard

Page 46: Surgical design for the Fontan procedure using computational fluid

22

Galerkin methods.

The strong form of the steady convection-diffusion problem, (S), is stated

as follows:

(S)

⎧⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎩

Given f : Ω → R, find u : Ω → R, such that

a · ∇u− κ∇2u = f on Ω

u(0) = g0 on Γ

u(L) = gL on Γ

(2.1)

where a is the convection velocity, κ is the diffusion coefficient and f is the force

term. Then, we define a collection of trial solutions, S, and a collection of weighting

functions, V , as follows:

S ={u|u ∈ H1, u(0) = g0 and u(L) = gL

}, (2.2)

V ={w|w ∈ H1, w(0) = 0 and w(L) = 0

}, (2.3)

where H1 denotes a class of Sobolev spaces, in which functions and their first

derivative are square integrable.

We now multiply equation (2.1) by w and integrate the diffusion term by

parts. Let u,x = du/dx and (w, u) =∫ΩwudΩ. We obtain the weak form (W ) as

follows:

Page 47: Surgical design for the Fontan procedure using computational fluid

23

(W )

⎧⎪⎪⎪⎨⎪⎪⎪⎩Given f : Ω → R, find u ∈ S such that for all w ∈ V

(w, au,x) + (κu,x, w,x) = (w, f) ,

(2.4)

Equation (2.4) can now be discretized by the Galerkin method. We approximate

u and w by uh and wh such that

u ≈ uh =

n∑A=1

dANA + g0N0 + gLNn+1, (2.5)

w ≈ wh =

n∑A=1

cANA, (2.6)

where NA’s are shape functions that satisfy

NA(0) = 0, A = 1, 2, ..., n+ 1, (2.7a)

NA(L) = 0, A = 0, 2, ..., n (2.7b)

N00 = 1 (2.7c)

Nn+1(L) = 1. (2.7d)

(2.7e)

We can derive a system of linear algebraic equations by substituting equations

Page 48: Surgical design for the Fontan procedure using computational fluid

24

(2.5) into equation (2.4).

n∑B=1

[(NA, aNB,x

)+

(NA, κNB,x

)]dB

= (NA, f)−(NA, aNn+1,x

)−

(κNn+1, NA,x

), for A = 1, 2, 3...n (2.8)

Given a mesh size h, the mesh Peclet number Pe is defined as ah2κ. For

a large Pe number, which indicates the flow is convection dominated, the stan-

dard Galerkin method is not stable. Considering a 1D steady convection-diffusion

problem with a zero force term, f = 0 and Dirichlet boundary conditions, g(0) =

0, g(L) = 1 in a dimensionless domain, L = 10, it can be proven that the Galerkin

scheme is equivalent to the central difference method which introduces negative

numerical diffusion for the diffusive term and results in unstable solutions.84 An

upwind scheme was proposed to eliminate the oscillation and achieve exact nodal

solutions. The optimal upwind scheme is equivalent to solving a modified equation

with artificial diffusion using a Galerkin method:

a · ∇u− (κ+ κ)∇2u = 0, (2.9)

κ = βah

2,

β = cothPe− 1/Pe,

Page 49: Surgical design for the Fontan procedure using computational fluid

25

where β is a parameter to control the magnitude of artificial diffusion.

0 1 2 3 4 5 6 7 8 9 100

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

GalerkinEADSUPGEaxct Sol.

a=1 κ=1

0 1 2 3 4 5 6 7 8 9 10−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

0.6

0.8

1

x

u

a=10 κ=1

x

u

0 1 2 3 4 5 6 7 8 9 10−5

−4

−3

−2

−1

0

1

x

u

a=100 κ=1

Figure 2.1: Solutions for the 1D steady convection-diffusion problem (2.1) withf = 0, g(0) = 0, g(10) = 1 using Galerkin, exact artificial diffusion (EAD) andstreamline-upwind-Petrov-Galerkin (SUPG) schemes.

Figure 2.1 shows that the standard Galerkin method is unstable for con-

vection dominated flow. However, the upwind scheme is not consistent, resulting

Page 50: Surgical design for the Fontan procedure using computational fluid

26

in incorrect solutions when the forcing term is present (Figure 2.2). Brooks and

Hughes proposed a stabilized consistent scheme called streamline-upwind-Petrov-

Galerkin (SUPG) to overcome these difficulties.81 The idea of SUPG is to apply a

modified weighting function w = w+p to all terms in order to achieve a consistent

formulation. For equation (2.4), the corresponding SUPG scheme is

(wh, auh

,x

)+

(wh

,x, κuh,x

)+

nel∑e=1

∫Ωe

p(auh

,x − kuh,xx − f

)dx =

(wh, f

), (2.10)

p = aw,xτ,

τ = κ/ ‖a‖2 ,

where∑nel

e=1 and Ωe denote a summation from the 1st element to the nelth element

and the domain of the eth element. Figure 2.2 shows that the SUPG overcomes

the shortcoming of the upwind scheme.

2.1.2 Stabilized FEM for Navier-Stokes equations

In the previous section, we have shown that the standard Galerkin method

results in unstable solutions for convection-dominated flow problems. The SUPG

scheme suppresses the oscillatory phenomena in a consistent manner. In this sec-

tion, the SUPG formulation is applied to incompressible viscous fluid flows. The

incompressible Navier-Stokes equations for a Newtonian fluid can be written as

Page 51: Surgical design for the Fontan procedure using computational fluid

27

0 1 2 3 4 5 6 7 8 9 10−1

0

1

2

3

4

5

6

7

8

x

u

GalerkinEADSUPGExact Sol.

a=10 κ=1

Figure 2.2: Solutions for the 1D steady convection-diffusion problem (2.1) withf = −16 a

10

(−2 + 4 x

10

), g(0) = 0, g(10) = 1 using Galerkin, exact artificial diffusion

(EAD) and streamline-upwind-Petrov-Galerkin (SUPG) schemes.

(S)

⎧⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎩

Given f : Ω× (0, T ) → R3, find u(x, t) and p(x, t) : Ω → R3, such that

ρu,t + ρu · ∇u = −∇p +∇ ·T+ f in Ω× (0, T )

T = μ(∇u+∇uT

)∇ · u = 0 in Ω× (0, T )

u(x, 0) = u0(x) in Ω

u = g in Γg × (0, T )

n · (−pI+T) = h in Γh × (0, T )

(2.11)

where ρ, μ, u, p, T, and f are density, dynamic viscosity, velocity, pressure, devi-

Page 52: Surgical design for the Fontan procedure using computational fluid

28

atoric stress tensor and body force, respectively. The boundary is divided into Γg

and Γh (Γg ∪Γh = Γ and Γg ∩Γh = ∅ ), in which the velocity (Dirichlet condition

) and traction (Neumann condition) are prescribed, respectively.

The trial solution and weighting function spaces for the semi-discrete for-

mulation are defined as

S ={u|u (x, t) ∈ H1 (Ω)3 , t ∈ [0, T ] ,u (x, t) = g on Γg

}, (2.12)

W ={w|w (x, t) ∈ H1 (Ω)3 , t ∈ [0, T ] ,w (x, t) = 0 on Γg

}, (2.13)

P ={p|p (x, t) ∈ H1 (Ω) , t ∈ [0, T ]

}. (2.14)

The weak form for equation (2.11) using the stabilized finite element method is :

Page 53: Surgical design for the Fontan procedure using computational fluid

29

(W )

⎧⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎪⎩

Given f : Ω× (0, T ) → R3 and g : Γg × (0, T ) → R�

find u(x, t) ∈ S and p(x, t) ∈ W such that ∀ w ∈ W and q ∈ P

BS (w, q;u, p) = 0

BS (w, q;u, p) = BG (w, q;u, p)

+∑nel

e=1

∫Ωe

(u · ∇w · τmR+∇ ·wτc∇ · u) dx

+∑nel

e=1

∫Ωe

[w · (−τmR · ∇u) + (R · ∇w) · (τR · ∇u)] dx

+∑nel

e=1

∫Ωe

∇q · τmρRdx

BG (w, q;u, p) =∫Ω[w · (ρu,t + ρu · ∇u− f) +∇w : (−pI+T)] dx

−∫Ω∇q · udx−

∫Γh

w · (−pI+T) · nds+∫Γqu · nds

where R is the residual vector of the momentum equation in Equation (2.11)

and τm, τc and τ are stabilization parameters that are constructed to achieve exact

solutions in the case of 1D model problems .81, 85, 94 The details of these parameters

can be found in .81,85, 94 A second-order accurate generalized α-method is used to

integrate the above semi-discrete equations with respect to time.94, 95

2.1.3 Boundary conditions

Dirichlet and Neumann boundary conditions were used in the problem

stated above. A no-slip boundary condition is imposed on the walls. Since the

Page 54: Surgical design for the Fontan procedure using computational fluid

30

inlet flowrates usually can be obtained by PC-MRI, Dirichlet boundary condi-

tions were used at the inlets by mapping parabolic or Womersley profiles. Outlet

boundary conditions are critical to obtain physiologically relevant solutions. In

most cases, the pressure or velocity measurements are not available at the outflow

boundaries. In addition, it is infeasible to model the entire vascular system in 3D

flow simulations. The 3D model is truncated at the boundaries of a domain of

interest. Thus, it is challenging to prescribe correct Dirichlet or Neumann outlet

boundary conditions directly. Alternatively, a model that prescribes a pressure-

flow relationship accounting for the downstream vasculature is preferred for the

outflow boundary conditions.96

Figure 2.3: A spatial domain is divided into a 3D domain Ω modeled by Navier-Stokes equations and a downstream Ω′ modeled by lump parameter models. TheDtN outflow boundary conditions are prescribed on the boundary ΓB that separatesΩ and Ω′.

Lumped parameter models including resistance, impedance and three el-

ement Windkessel models (RCR) are standard choices for cardiovascular simu-

lation boundary conditions. Lumped parameter models quantitatively describe

the pressure-flow relationship but lack detailed local flow information. Therefore,

lumped parameter models are suitable for modeling downstream vascular beds.

Page 55: Surgical design for the Fontan procedure using computational fluid

31

Vignon-Clementel et al.85 incorporated resistance, impedance and RCR models

into 3D finite element flow simulations, using the DtN map to achieve physiologic

flow conditions.85,97 In the work of Vignon-Clemental et al.,85 a spatial domain Ω

is discretized into a 3D numerical domain Ω and a downstream lumped parameter

domain Ω′ such that Ω ∩ Ω′ = ∅ and Ω ∪ Ω′ = Ω as shown in Figure 2.3. Then, u

is decomposed as

u = u+ u′ with u|Ω′ = 0,u′|Ω = 0 and u|ΓB= u′|ΓB

(2.15)

Other variables and weighting functions can be decomposed in a similar way. Since

variables and weighting functions for the domain Ω vanish on the domain Ω′ and

vice versa, the variational form for domain Ω becomes:

BG(w, q;u, p) =

∫Ω

w · (ρu,t + ρu · ∇u− f) +∇w :(−pI+ T

)dx

−∫Γh

w ·(−pI+ T

)· nds

+

∫Ω′w′ · (ρu′

,t + ρu′ · ∇u′ − f) +∇w′ : (−p′I+T′) dx

−∫Γ′

h

w′ · (−p′I+T) · n′ds−∫Ω

∇q · udx

+

∫Γ

qu · nds−∫Ω′∇q′ · udx+

∫Γ′q′u′ · n′ds (2.16)

Since Ω∩Ω′ = ∅, the third and seventh terms in the right hand side (RHS) have no

contribution to domain Ω. Therefore, the variational form for the 3D subdomain

Page 56: Surgical design for the Fontan procedure using computational fluid

32

Ω is shown as follows:

BG(w, q;u, p) =

∫Ω

w · (ρu,t + ρu · ∇u− f) +∇w :(−pI+T

)dx

−∫Γh

w ·(−pI+ T

)· nds−

∫ΓB

w′ · (−p′I+T) · n′ds

−∫Ω

∇q · udx+

∫Γ

qu · nds+∫ΓB

q′u′ · n′ds (2.17)

In Equation (2.17), the third and sixth integrals are the boundary terms that

connect to the downstream domain Γ′. Since variables and weighting functions

for domains Ω and Ω′ are equal, and n′ = −n on the boundary ΓB, the only

unknown term is the pressure p′ in the third integral. Lumped parameter models

approximate p′ on the boundary by specifying a pressure-flow relationship.

Resistance model

Resistance is the simplest pressure-flow relationship which defines a con-

stant ratio of pressure to flowrate R = P/Q. Assuming the pressure is constant

over the cross sectional area of the boundary ΓB, we have:

p′ − n′ ·T′ · n = −R

∫ΓB

u′ · n′ds = R

∫ΓB

u · nds (2.18)

Therefore, the unknown Neumann boundary condition can be approximated

as

Page 57: Surgical design for the Fontan procedure using computational fluid

33

(−p′I+T′) ≈ −R

∫ΓB

u · ndsI− n · T · nI+T (2.19)

Windkessel RCR model

Instead of using a single resistance, a Winkessel RCR model uses a capacitor

and two resistors to model the downstream domain (Figure 2.4). The proximal

resistance R and capacitance C represent the major downstream arteries. The

distal resistance R represents the distal vascular bed. The RCR circuit ordinary

differential equation is:

P +RdCdP

dt= (R +Rd)Q+RRdC

dQ

dt+ Pd +RdC

dPd

dt(2.20)

RdR

C

P, Q

Pd(t)

Pd(t)

Figure 2.4: A three element Winkessel model.

Similarly, we use the pressure-flow relationship in Equation (2.20) to ap-

proximate the unknown Neumann boundary conditions in Equation (2.17).

Page 58: Surgical design for the Fontan procedure using computational fluid

34

p′ − n′ ·T′ · n′ =[P ′(0) +R

∫ΓB

u′(0) · n′ds− P′d(0)

]e−t/(RdC) + P

′d(t)

− R

∫ΓB

u′(t1) · n′ds−∫ t

0

e−(t−t1)/(RdC)

C

∫ΓB

u′(t1) · n′dsdt1 (2.21)

Therefore, the Neumann boundary conditions can be approximated as

(−p′I+T′) ≈(−R

∫ΓB

u(t1) · nds)I

−(∫ t

0

e−(t−t1)/(RdC)

C

∫ΓB

u(t1) · ndsdt1)I− n · I · nI

−[(

P ′(0)−R

∫ΓB

u · nds− P′d(0)

)e−t/RdC + P

′d(t)

]I+ T (2.22)

2.2 Surrogate management framework (SMF)

For many simulation-based optimization problems with multiple design pa-

rameters, it is prohibitively expensive to perform a large number of simulations to

identify an optimal solution. Despite inevitable errors, surrogate models, which

approximate or interpolate the true cost functions based on a limited set of known

data, may predict trends in the true cost function, reducing unsuccessful trials

when gradient information is not available. Generally, surrogate models can be

polynomials such as Lagrangian interpolation or statistics-based models such as

Kriging, which is also known as Gaussian process regression. Compared to the true

Page 59: Surgical design for the Fontan procedure using computational fluid

35

cost function, surrogate models are usually cheap to evaluate. Thus, the methods

used to search the minimum of surrogate models are less restricted.

SMF, proposed by Booker et al.,98 incorporates a surrogate model into

pattern search algorithms to improve search efficiency with a convergence proof

provided by pattern search algorithms. The SMF consists of the following steps.

First, we employ latin hypercube sampling (LHS) to construct a well distributed

initial set in the parameter space.99 An initial surrogate is built by evaluating

the cost function at each design point in the initial set. We define a mesh in

the parameter space, and all points subsequently evaluated by the algorithm are

restricted to lie on this mesh. The search and poll step are executed alternately,

depending on whether a design point that improves the current best cost function

is found. Figure 2.5 shows a flowchart of the SMF algorithm using mesh adaptive

direct search (MADS) which is a type of pattern search algorithms.86 In the

search step, the minimum of the current surrogate function, the incumbent point,

is evaluated to calibrate the model prediction. Search steps are performed until

they fail to find an improved point, at which time a poll step is performed. In

the poll step, points neighboring the current best point are evaluated in a set of

positive spanning directions.86 Letting Dk be the positive spanning set in Rn at

iteration k, the polling points Pk are given by

Pk = {xk +Δmk d : d ∈ Dk} , (2.23)

Page 60: Surgical design for the Fontan procedure using computational fluid

36

where xk is the current best point and Δmk is the mesh size in parameter space at

iteration k, defined to be equal to 4−l, l = 0, 1, 2, 3....

If the poll step is successful, the algorithm returns to the search step. If the

poll step fails to find an improved point, the mesh is refined and we return to the

search step. When the parameter space mesh has been refined to a size smaller than

the minimum size set by the user, the optimization algorithm will stop and output

the optimal solution. The poll step guarantees convergence to a local minimum

of the function, following previously published convergence proofs.98, 100–102 This

indicates that SMF will converge regardless of the search strategy under certain

conditions .98 In addition, the rule to update the mesh size is flexible. The mesh

size can be returned to the previous size or kept at the current size if the search

or poll step is successful.103

Ini�aliza�onLHS Search

PollA set of posi�ve

spanning direc�on

Stop

Improved?

Improved?

Yes

No

Yes

Yes

NoRefine mesh

Figure 2.5: Flowchart of SMF using MADS. Search and poll steps are executedalternately according to whether a design point that improves the current best costfunction is found.

Page 61: Surgical design for the Fontan procedure using computational fluid

37

2.2.1 Surrogate models

Kriging is part of a class of statistical interpolation techniques for a random

field, pioneered by geostatisticians Krige and Matheron .104–106 Later on, Kriging

was applied to deterministic and random simulation models with multiple inputs

.107–111 Generally speaking, a Kriging model is a linear combination of n observa-

tions with n weights such that the variance of prediction is minimized. Since the

actual function is unknown, the error at each estimated point is modeled based on

probability theory. Therefore, the key to creating a Kriging model is to determine

the combination of weights for a given data set. Following Lophaven et al. ,112 we

briefly introduce the construction of Kriging.

Assuming we have m pairs of design sites S = [s1 · · · sm]T with si ∈ Rn

and responses Y = [y1 · · · ym]T, the Kriging predictor y(x) for x ∈ Rn is a linear

combination of known responses

y(x) = cTY, (2.24)

where c = c(x) ∈ Rm.

Then, we assume the stochastic process Y can be modeled by a linear

combination of p regression models and a random function:

Y = Fβ + Z, (2.25)

Page 62: Surgical design for the Fontan procedure using computational fluid

38

where F = [Fij ] = [fj(si)] ∈ Rmp is the basis function matrix, β = [β1 · · ·βp]T

is the coefficient vector, and Z = [z1 · · · zm]T is a random function vector. For

x = [x1 · · ·xn]T, the regression models with polynomials of orders 0 to 2 are defined

as

Constant: f1(x) = 1, p = 1, (2.26a)

Linear: f1(x) = 1, f2(x) = x1, . . . , fp(x) = xn, p = n + 1 (2.26b)

Quadratic: f1(x) = 1, . . . , fn+1(x) = xn, . . . , fn+2(x) = x21, . . . ,

f2n+1(x) = x1xn, . . . fp(x) = x2n, p =

1

2(n + 1)(n+ 2). (2.26c)

Therefore, y(x) is modeled by

y(x) = f(x)Tβ + z. (2.27)

Using Equations (2.24),(2.25) and (2.27), we obtain an expression for the error:

y(x)− y(x) = cTY − y(x)

= cT(Fβ + Z)− (f(x)Tβ + z)

= cTZ − z + (FTc− f(x))Tβ (2.28)

Page 63: Surgical design for the Fontan procedure using computational fluid

39

Imposing the unbiased condition, we require FTc(x) = f(x). Thus, the mean

squared error (MSE) is

E[(y(x)− y(x))2]

=E[(cTZ − z)2]

=E[z2 + cTZZTc− 2cTZz] (2.29)

The correlation matrix for the random function z between two design sites is define

as follows,

Rij = R(θ, si, sj), i, j = 1, . . . , m, (2.30)

where θ is a correlation parameter. A vector of correlations between an unknown

point x ∈ Rn and known design sites can be defined as follows,

rx = [R (θ, s1, x) . . .R (θ, sm, x)]T . (2.31)

Thus we have E[z2] = σ2, E[Zz] = σ2r, E[ZZT] = σ2R, and Equation (2.29)

becomes

E[(y(x)− y(x))2] = σ2(1 + cTRc− 2cTr). (2.32)

Correlation models are user defined. A Guassian correlation function is chosen in

Page 64: Surgical design for the Fontan procedure using computational fluid

40

this work. Applying Lagrange multipliers to Equation (2.32) to minimize the MSE

with the unbiased condition, the Lagrangian and its gradient with respect to c are

L(c, λ) = σ2(1 + cTRc− 2cTr)− λT(FTc− f), (2.33)

L′(c, λ) = 2σ2(Rc− r)− Fλ. (2.34)

When the MSE is minimized, (2.34) is zero. Thus we obtain a linear system

Rc− Fλ

2σ2= r (2.35a)

FTc = f (2.35b)

Solving Equation (2.35), we get an expression for c in terms of F , R and r

c = R−1[r − F (FTR−1F )−1(FTR−1r − f)]. (2.36)

Substitute (2.36) into (2.24), we have

y(x) = rTR−1Y − (FTR−1r − f)T(FTR−1F )−1FTR−1Y. (2.37)

Page 65: Surgical design for the Fontan procedure using computational fluid

41

2.2.2 Mesh adaptive direct search (MADS)

MADS was developed by Audet and Dennis to improve on the convergence

theory of previous GPS polling methods, particularly in the presence of nonlinear

constraints. First, we define a mesh

Mk = x+Δmk Dz : x ∈ Vk, z ∈ NnD , (2.38)

where k denotes the number of iterations, Vk ⊂ Rn is the set of all points with

known cost function values, Δmk ∈ R+ is the mesh size parameter, and D ∈ Rn×nD

is a set of positive spanning directions. At least n + 1 and at most 2n vectors

are required to form a positive basis. The polling points for a polling center xk at

iteration k can be defined as

Pk = {xk +Δmk d : d ∈ Dk ⊂ D} . (2.39)

Pattern search algorithms are differentiated by the choice of positive spanning

direction Dk. The original variant of SMF used generalized pattern search (GPS)

for polling. The drawback of the GPS algorithm is that the polling directions are

chosen from a fixed finite set D whose columns contains all possible combinations

of -1, 0, 1 except [0 0]T. For example, D =

⎡⎢⎢⎣ 1 0 −1 0 1 1 −1 −1

0 1 0 −1 1 −1 −1 1

⎤⎥⎥⎦,

when n = 2. Since the polling distance is given by Δmk ‖d‖∞, the polling parameter

Δpk, which determines the magnitude of the polling distance, is equal to the mesh

Page 66: Surgical design for the Fontan procedure using computational fluid

42

size parameter Δmk . If the polling frame shrinks more slowly than the mesh size, the

number of polling candidates will be increased. For MADS, Dk is constructed such

that Δmk ‖d‖ ≤ Δp

k max {‖d′‖ : d′ ∈ D} and limk→+∞ inf Δmk = limk→+∞ inf Δp

k =

0.

Audet et al.86 proposed an instance of MADS called LTMADS that uses

a lower triangluar matrix (LT) to construct positive spanning directions Dk. The

procedure is summarized below.

1. Construct a LT matrix B in which the diagonal terms are −1√Δm

k

or 1√Δm

k

and

lower entries are integers randomly chosen with equal probability in the open

interval ] −1√Δm

k

, 1√Δm

k

[.

2. Permute the lines and rows of B randomly. Letdq =[Bi1,jq , Bi2,jq , . . . , Bin,jq

]Tfor q = 1, . . . , n, where i1, i2, . . . , in and j1, j2, . . . , jn are random permutations

of the set 1,2,. . . ,n.

3. Construct a positive basis:

(a) n+ 1 directions: Dk = [d1, d2, . . . , dn+1], where dn+1 = −∑n

i=1 di.

(b) 2n direction: Dk = [d1, . . . , dn, dn+1, . . . , d2n] = [d1, . . . , dn,−d1, . . . ,−dn]

Since the entries in Dk are chosen from [ −1√Δm

k

1√Δm

k

], we have ‖d‖∞ = 1√Δm

k

and

‖d‖∞ ≤ n√Δm

k

and for 2n and n+1 positive spanning directions respectively. Thus

the polling frame defined by ‖Δmk d‖∞ = Δm

k ‖d‖∞ is bounded by n√

Δmk = Δp

k.

It has been shown the set of polling directions generated by LTMADS over all

iterations is dense.86

Page 67: Surgical design for the Fontan procedure using computational fluid

Chapter 3

Constrained Optimization of an

Idealized Y-graft Model

To extend the previous framework,76 we now propose a model problem in

which we optimize a new Y-graft design for the Fontan procedure, a surgery used

to treat single ventricle heart defects. We perform constrained optimization of an

idealized Y-graft model with multiple design parameters under pulsatile rest and

exercise flow conditions.

In this work, formal shape optimization with established convergence the-

ory was performed on an idealized Y-graft geometry that was parameterized using

six design parameters. To our knowledge, this work represents the first use of

formal optimization algorithms for Fontan surgery design to date. The goal of this

study is to demonstrate the feasibility of applying optimization to Y-graft design,

to compare the optimal shapes with those tested previously, to identify optimal

43

Page 68: Surgical design for the Fontan procedure using computational fluid

44

parameters for an idealized Y-graft Fontan model, and to assess the impact of

different flow conditions and constraints on the resulting optimal design. Results

from this work demonstrate that the SMF framework is efficient and robust for

pulsatile 3D problems with unsteady flow and constraints. While it is unlikely

the idealized model used in this work will be sufficient to make a conclusive sur-

gical recommendation, this work will lay the foundation for future patient specific

optimization that may assist in improving patient outcomes.

3.1 Methods

3.1.1 Model construction and parameterization

Idealized 3-D solid models of the Fontan geometry were generated using a

customized version of the open-sourced Simvascular software environment.113,114

The geometries of the IVC, Y-graft, PAs and SVC were defined analytically, and

constructed by lofting together a series of circles and ellipses with prescribed radii.

The model was parameterized with six parameters (Figure 3.1), each of which

could be systematically varied to generate a range of Y-graft designs. Models

were generated automatically using a computer script which takes the six design

parameters as inputs and produces as output a three dimensional solid model. The

design parameters Dbranch, LIV C , XLPA and XRPA are defined as the diameter of

branches, length of the IVC trunk, and positions where left and right branches

connect to the LPA and RPA, respectively. Parameters ΔR and ΔL, which control

Page 69: Surgical design for the Fontan procedure using computational fluid

45

the curvature of the two branches, are the perpendicular distances to the dashed

lines C0C1 and C0C2. The centerline paths of the branches C0C3C1 and C0C4C2 are

defined by interpolating two cubic Hermite spline functions passing through control

points C0, C1, C2, C3 and C4, marked with circles shown in Figure 3.1a. Given

LIV C , XLPA, XRPA, the end points of the two graft branches, the positions of C0,

C1 and C2, are fixed. The branch curvatures are controlled by manipulating the

control points C3 and C4, which are the midpoints of curves C0C3C1 and C0C4C2,

respectively. After the centerlines of the idealized Y-graft model are determined, a

series of circles perpendicular to the centerlines are lofted together to form a solid

model. Since the sizes of the IVC, graft, PAs and SVC were different, interpolation

was used to smoothly connect the geometry at each junction.

In addition to the six design parameters described above, four other con-

stant parameters define the model geometry. The diameters of the SVC, IVC, PAs

and the distance between the PAs and bottom of the graft, which are denoted

by DSV C ,DIV C , DPA and LPA−IV C , respectively, were measured from MRI image

data of a typical 4-year-old male Fontan patient with a traditional T-junction graft.

Tables 3.1 and 3.2 list the bounds of the design parameters and constants. The

maxima of XLPA and XRPA were based on the distances between the SVC and

the first branches of LPA and RPA as measured by the image data. Other bounds

for parameters were chosen to give reasonable flexibility in the design space while

maintaining surgical feasibility.

Page 70: Surgical design for the Fontan procedure using computational fluid

46

−10 −8 −6 −4 −2 0 2 4 6 8 100

2

4

6

8

10

12

XLPAXRPA

C2

ΔL

C4

C0

LIVC

ΔR C3

C1

(a) Centerline paths of the model. Circles C0, C1, C2, C3, C4

are the control points of the two branches. Parameters ΔR andΔL, which control the curvatures of the two branches, denotethe perpendicular distance to the dashed line C0C1 and lineC0C2. The other four design parameters are XLPA, XRPA,LIV C and Dbranch. Dbranch is marked in Figure 3.1b.

Outflow RPA

0 0.5 1 1.5 2 2.5 3−10

0

10

20

30

40

50

60

time(s)

IVC

flo

wra

te(c

c/s

)

0 0.5 1 1.5 2 2.5 3−10

0

10

20

30

40

50

60

time(s)

SV

C fl

ow

rate

(cc

/s)

Inflow IVC

Inflow SVC

Outflow LPA

Dbranch

(b) Solid model and pulsatile flow waveform applied at theinlets of IVC and SVC. Flow directions are denoted witharrows. Resistance boundary conditions are used at theoutlets of the PAs.

Figure 3.1: Model parametrization showing the six design parameters used forshape optimization (a), and the resting pulsatile IVC and SVC flow waveformsused for inflow boundary conditions (b).

Page 71: Surgical design for the Fontan procedure using computational fluid

47

Table 3.1: Bounds on the design parameters for the idealized model. Negativevalues for ΔR and ΔL indicate inward convex branches and positive values denoteoutward concave branches. Bounds were chosen to be consistent with MRI datafrom a typical patient.

Design parameters Max (cm) Min (cm)

LIV C 4.0 2.0Dbranch 2.4 1.2XRPA 3.0 1.0ΔR 0.1 -0.3

XLPA 6.0 1.0ΔL 0.1 -1.0

Table 3.2: Values of the four constant geometric parameters used in model con-struction, taken from MRI data of a typical Fontan patient.

Constant parameters cm

DSV C 1.2DIV C 2.0DPA 1.2

LPA−IV C 7.0

Page 72: Surgical design for the Fontan procedure using computational fluid

48

3.1.2 Flow simulation and boundary conditions

To simulate blood flow, a stabilized finite element solver94, 115 from the Sim-

vascular software package was used to solve the time-dependent 3D Navier-Stokes

equations. The simvascular flow solver has been developed from PHASTA(Parallel,

Hierarchical, Adaptive, Stabilized, Transient Analysis),115 which uses the SUPG

(Streamline-Upwind Petrov-Galerkin) formulation81 and a generalized α-method95

with second order accuracy in time. To tailor PHASTA for blood flow simulations,

developments in the areas of boundary conditions85 and fluid-solid interaction116

have been made. The solver has been previously validated using comparisons with

analytical solutions and pulsatile flow experiments.94, 117, 118 We employ a rigid-

wall and Newtonian approximation in this work. The viscosity was set to 0.04

g/(cm · s) and the density of blood was assumed to be 1.06 g/cm3. Resistance

boundary conditions85 were applied at the outlets of the PAs to model the resis-

tance of the downstream pulmonary beds. Resistance values for the resting flow

conditions were chosen such that the mean IVC pressure would be 12 mmHg, a

standard clinical value, and flow split of the RPA/LPA would be approximately

55%/45%.

While many previous Fontan simulations have employed steady inflow con-

ditions at rest, recent work has demonstrated the importance of including pul-

satile flow, respiration, and exercise effects.4, 46, 51 In this study, steady and pul-

satile inflow profiles were applied at the IVC and SVC inflow faces, and mapped to

Page 73: Surgical design for the Fontan procedure using computational fluid

49

Table 3.3: Mean flow rates, Re in the IVC and SVC and resistance drops at restand two levels of simulated exercise.

Exercise level IVC (cc/s) SVC (cc/s) Resistance drop ReIV C(mean) ReSV C(mean)

rest 17.87 14.07 0% 302 3872X exercise 35.74 14.07 5% 603 3873X exercise 53.61 14.07 10% 905 387

parabolic velocity profiles. Pulsatile flow values were obtained from phase-contrast

MRI data of a typical Fontan patient.36 Following our previous work,4 a respira-

tory model was superimposed on the IVC inflow waveform. To simulate exercise,

the IVC flow rate was increased by 2 and 3 times (referred to as 2X and 3X in

the text) the resting flow value and the SVC rate was kept unchanged.36 Resis-

tance values were decreased by 5% and 10% to simulate vasodilation for the 2X

and 3X exercise levels, respectively.4, 119 The SVC and IVC pulsatile waveforms

with respiration for the three cases are shown in Figure 3.2, while the mean flow

rates, resistance drops and Reynolds number (Re) are listed in Table 3.3. Mesh-

Sim (Simmetrix Inc., Clifton Park, NY) is integrated into Simvascular to generate

tetrahedral meshes using an iterative mesh optimization algorithm to improve the

mesh quality.113,120In this work, meshes were generated automatically with approx-

imately 300,000 elements. The number of elements was increased by about 30-70%

for the pulsatile exercise cases to ensure that quantities of interest did not vary

significantly. The time step was chosen according to the CFL (Courant, Friedrichs,

and Lewy) condition, and it varied from 0.01s to 3.3× 10−4s. Typical 3X exercise

simulations under steady and pulsatile inflow conditions required approximately

2.9× 102 and 4.1× 103 CPU seconds with 52 processors, respectively.

Page 74: Surgical design for the Fontan procedure using computational fluid

50

0 0.5 1 1.5 2 2.5 3

0

20

40

60

80

100

t (s)

flow

rate

(cc/

s)

IVC restSVC rest

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8

0

20

40

60

80

100

t (s)

flow

rate

(cc/

s)

IVC 2XSVC 2X

0 0.2 0.4 0.6 0.8 1 1.2 1.4

0

20

40

60

80

100

t (s)

flow

rate

(cc/

s)

IVC 3XSVC 2X

Figure 3.2: Pulsatile waveforms for the rest, 2X and 3X exercise cases. To simulatetwo exercise levels, the IVC flow rate was increased by 2 and 3 times; SVC flowwas unchanged.

3.1.3 Unconstrained optimization

Following the method outlined in Marsden et al.,76 we consider the opti-

mization problem,

minimize J(x),

subject to x ∈ Ω, (3.1)

where J : Rn → R is the cost function, and x is the vector of parameters.

SMF introduced in Chapter 2 was used to minimize the cost function. To

fully automate the shape optimization procedure, the following sub steps are linked

in our framework: model generation, meshing, cost function evaluation (flow sim-

ulation and post processing) and data transfers. A series of custom scripts were

created and linked with Simvascular to implement and link the sub steps above

so that the optimization procedure did not require any user intervention. These

scripts call specific sub-functions inside Simvascular to automatically generate the

Page 75: Surgical design for the Fontan procedure using computational fluid

51

Parameters

Op�miza�on Post processing Flow simula�on

Model construc�on1. Create the centerline of vessels2. Taking circles perpendicularly3. Lo� and generate solid model 4. FE mesh genera�on

Figure 3.3: The shape optimization procedure is made up of a series of automatedsub-steps from model construction to the input of the cost function value into theoptimization algorithm.

prescribed model, and run the required flow simulation (including meshing, bound-

ary condition prescription, and post-processing), and return a cost function value.

The optimization code was written in Matlab, compiled with the Matlab compiler,

and was called automatically from within the Simvascular scripts. Figure 3.3 shows

the sub steps of the optimization procedure, where each box represents an auto-

mated sub step. The boxes and arrows in Figure 3.3 form a loop that repeats until

the convergence criteria are satisfied.

3.1.4 Polling strategies

The polling directions used in SMF must form a positive spanning set.

In this work we evaluate two competing polling strategies in order to determine

the sensitivity of the optimal solution to the polling strategy used and to com-

pare computational efficiency. In our previous work and most optimization results

presented in this paper, n + 1 positive spanning directions were employed in the

Page 76: Surgical design for the Fontan procedure using computational fluid

52

polling step using the LTMADS method, which has been introduced in Chapter

2.86 LTMADS generally performs well and requires fewer function evaluations than

traditional generalized pattern search (needing 2n poll directions). However, due

to the random permutation required, the results of optimization using LTMADS

are not deterministic and large polling angles may arise.

Recently Abramson et al.121 proposed a new polling strategy called Or-

thoMADS to generate deterministic and orthogonal polling directions that make

results repeatable using 2n poll points. In practice, the only difference between

LTMADS and OrthoMADS is the method used to generate the positive spanning

basis, D. The procedure used to construct the 2n orthogonal basis in Abramson et

al.121 is summarized below, and we refer the reader to this work for further details.

1. Choose a Halton direction ut, where t is the tth direction in the Halton se-

quence.

2. Find a scalar αt,l such that ‖qt(αt,l)‖ is as close as possible to 2|l|/2, where

qt(αt,l) = round(αt,l

2ut−e‖2ut−e‖

), l is the mesh parameter defined in equation

(2.23) and e is the vector whose components are all equal to 1.

3. Apply Householder transformation.

(a) T = ‖qt(αt,l)‖2(In − 2vvT ), where v =qt(αt,l)

‖qt(αt,l)‖ and In is the identity

matrix of dimension n.

(b) D = [T ,−T ].

While the bulk of results presented here use the LTMADS method, we also present

Page 77: Surgical design for the Fontan procedure using computational fluid

53

a thorough comparison of optimization results and efficiency using LTMADS and

OrthoMADS. Comparisons were made by performing multiple runs of LTMADS

and comparing to a single deterministic run of OrthoMADS.

3.1.5 Constrained optimization

The SMF method can be extended to the case of constrained optimiza-

tion in a straightforward manner using a filter method. The filter method was

introduced by Fletcher and Leyffer,122 and has been successfully applied to pat-

tern search algorithms86 and aeronautic constrained optimization problems in our

previous work.79,80 One of the advantages of the filter method is that it not only

identifies the optimal solutions which satisfy constraints, but it also reveals trade-

offs between the cost function and the constraint violation. The filter method can

thus provide a range of choices to help users such as surgeons and clinicians make

comprehensive decisions.

We consider the general constrained optimization problem

minimize J(x),

subject to x ∈ Ω, C(x) ≤ 0, (3.2)

where J : Rn → R is the cost function, and x is the vector of parameters. The

constraints are given by m functions ci : Rn → R, i = 1,2,...,m such that C(x)

Page 78: Surgical design for the Fontan procedure using computational fluid

54

= (c1(x),...,cm)T . To evaluate how closely the design satisfies the constraints, a

constraint violation function H is defined that can be formed by a single constraint

function or the weighted sum of multiple constraint functions. The bounds on the

parameter space are defined by a polyhedron in Rn denoted by Ω (Table 3.1).

In reviewing the algorithm for the constrained optimization problem, we

first define some basic concepts for completeness.103, 122

Definition 1. A point x is feasible if H(x) = 0, where H(x) is the constraint

violation function.

Definition 2. A point x is infeasible if H(x) > 0.

Definition 3. An infeasible point x′ is dominated or filtered by a point x if the

following conditions hold: H(x) ≤ H(x′) and J(x) ≤ J(x′).

Definition 4. An infeasible point x′ is undominated or unfiltered if there is no

dominating infeasible point.

Definition 5. A point x is called the least infeasible point if

H(x) = min {H(x′),x′ ∈ U}, where U is the set of undominated points.

Figure 3.4 is an example of a filter. Each point in Figure 3.4, plotted in the form of

cost function value J vs. constraint violation value H , represents a combination of

design parameters. In Figure 3.4a, points marked with either a triangle or a circle

are undominated points that form the filter. The filter excludes the dominated

points marked with stars because for each star point, there is at least one point with

Page 79: Surgical design for the Fontan procedure using computational fluid

55

a lower cost function and a smaller constraint violation in the filter. During the

optimization, the filter will be improved if a new undominated point is identified.

For example, a point with J = 0.4 and H = 0.15 will improve the current filter in

Figure 3.4a such that the improved filter is the one shown in Figure 3.4b. The filter

is continuously updated in this manner during the optimization until convergence

criteria are satisfied.

The basic algorithm for constrained optimization is the same as that of

the unconstrained optimization. The difference lies in the criteria that make the

search or poll step successful. In the constrained optimization, the search or poll

step is considered successful if a new undominated point is identified. The rest of

the SMF implementation remains the same.

3.1.6 Choice of cost function and constraints for Fontan

optimization

Low exercise capacity is a typical outcome after the Fontan procedure, and

Fontan patients’ exercise capacity typically decreases progressively over time.16

It has been previously hypothesized that the hemodynamic energy loss may be

related to diminished exercise capacity because a higher energy loss will impose a

larger work load on the heart.2, 46, 123 Based on this and our previous work, we have

chosen energy loss as the primary cost function for the Y-graft optimization, with

the acknowledgement that there are several other candidates for clinically relevant

Page 80: Surgical design for the Fontan procedure using computational fluid

56

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.80

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

H

J

dominated pointsfeasible pointsbest feasible pointleast infeasible pointundominated points

(a) Original filter example. Points are shown in a plot of cost function value J vs.constraint violation value H .

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.80

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

H

J

dominated pointsfeasible pointsbest feasible pointleast infeasible pointundominated points

(b) Improved filter example. A non-dominated point withH = 0.15 and J = 0.4 improvedthe filter forming a new filter.

Figure 3.4: Example of a filter for the constrained optimization problem. The filtershown in (a) is improved when a dominating point is found, producing the filtershown in (b).

Page 81: Surgical design for the Fontan procedure using computational fluid

57

cost functions that should also be considered in future work. Using the Reynolds

transport theorem, energy efficiency can be calculated by integrating the energy

flux over all inlets and outlets using the Equations (3.3) shown below. The energy

dissipation (neglecting gravitational effects) is given by

Ediss =

Ein︷ ︸︸ ︷−

Nin∑i=1

∫Ai

(p+1

2ρu2)u · dA−

Eout︷ ︸︸ ︷Nout∑i=1

∫Ai

(p+1

2ρu2)u · dA, (3.3)

where u is the velocity, p is the pressure, ρ is the density, Nin and Nout are the

number of model inlets and outlets, respectively, and Ai is the area of the ith inlet

or outlet. The energy efficiency is then

Eeffic = Eout/Ein. (3.4)

The efficiencies are the mean values over one respiratory cycle. Thus, cost function

J can be defined by

J = 1− Eeffic. (3.5)

Thromboembolic complications, which occur in 20-30% of cases, are another

common morbidity of the Fontan surgery.10, 124 For the extracadiac Fontan, the

incidence of conduit thrombosis is about 3-10%.20, 22, 125 Using a patient specific

model, Marsden et al.2 showed that although the 12mm Y-graft design has higher

energy efficiency, it has slightly larger regions of low WSS. Although the etiology

Page 82: Surgical design for the Fontan procedure using computational fluid

58

Figure 3.5: Time-averaged shear stress magnitude (dynes/cm2) of the optimalshape over one respiratory cycle during the rest condition using pulsatile waveform.

of thromboembolic complications is not well understood, some studies suggest

that shear stress plays a role in the formation of thrombosis.126 In the regions of

low WSS, it is hypothesized that higher particle residence time leads to increased

platelet aggregation and more adherence to the wall, increasing the probability of

clot formation.127 Previous studies have also shown that prothrombotic substances

may accumulate due to the flow or flow stasis.128–130 Here we hypothesize that the

formation of thrombosis is related to areas of low WSS. In order to limit the risk

of thrombosis in the proposed Y-graft design, we optimize to reduce energy losses

with a constraint on the time-averaged WSS. The constraint violation function is

defined by integrating the total surface area on the model for which the WSS falls

below a critical value, τcrit. The constraint violation is defined by

H =

∫S

fdS, f =

⎧⎪⎪⎪⎨⎪⎪⎪⎩1 if τ ≤ τcrit,

0 otherwise.

(3.6)

Figure 3.5 shows the time-averaged WSS contours of the optimal design

Page 83: Surgical design for the Fontan procedure using computational fluid

59

using the pulsatile waveform for the rest case. We observe that most of the low

WSS areas are located on the outer wall of the bifurcation near the IVC trunk and

higher WSS appears in the PAs, that have a relatively small diameter. Based on

the WSS contours of large graft designs, we chose 0.5 dynes/cm2 as the critical

value in order to minimize the deep blue regions in the WSS contours.

3.2 Results

3.2.1 Unconstrained optimization

Idealized Y-graft models were optimized for six different cases. Initial op-

timization was performed with no constraints, using energy efficiency as the cost

function. In the first group, steady inflow boundary conditions were used to sim-

ulate rest and two levels of exercise using 2 times and 3 times the IVC flow rate.

In the second group, pulsatile waveforms were applied using the respiration model

in the IVC. Figure 3.6 shows the surface pressure and velocity magnitude of the

optimal shapes for the three steady inflow cases. The resulting optimal param-

eters are summarized in Table 3.4. To reduce the total number of cost function

evaluations, optimal parameters for the rest case were added to the initial set at

the start of the 2X exercise optimization case. Similarly, optimal parameters from

the two previous cases were added to the initial set in the 3X exercise optimization

case.

Not surprisingly, the optimal parameters from the rest case are not optimal

Page 84: Surgical design for the Fontan procedure using computational fluid

60

under exercise flow conditions. Examining the parameters, the optimal shapes

for both the rest and 2X exercise cases have larger diameter branches than the

IVC trunk. The optimal size of the Y-graft branches is shown to decrease with

increasing exercise level, becoming smaller than the IVC trunk for the 3X exercise

case. One possible reason is that the large graft results in a smaller pressure drop

at low flow rates, but that flow separation leads to more energy loss at higher flow

rates.

(a) Pressure contours for the optimal shapes under the rest, 2X and 3X exercise levels.

(b) velocity magnitude on the centerline cut plane for the optimal shapes under the rest,2X and 3X exercise levels.

Figure 3.6: Unconstrained optimization results using steady inflow conditions.

Little difference was found between the optimal branch diameters with

steady and pulsatile (23.4 mm vs. 22.8 mm) inflow conditions at rest. How-

ever for the 2X exercise case the pusatile waveform produced a different optimal

solution than the corresponding steady case. This is likely due to higher energy

Page 85: Surgical design for the Fontan procedure using computational fluid

61

10 20 30 40 50 60 70 80 90 1000.011

0.0115

0.012

0.0125

0.013

0.0135

0.014

0.0145

number of evaluations

J

rest

2X

3X

Figure 3.7: Convergence history for the unconstrained optimization under steadyinflow conditions.

dissipation and a higher peak flow rate in the pulsatile case. The optimal branch

diameter is 18 mm in the 2X pulsatile case, which is smaller than the IVC trunk

(20 mm). The optimal parameters for the 3X exercise case are identical to the 2X

exercise case. This suggests that the optimal parameter set for the 2X exercise

case is also a local minimum in the 3X exercise case.

Figure 3.7 shows the convergence history for the three steady inflow cases.

The cost function was improved by 12%, 3% and 6% for the rest, 2X and 3X

exercise cases, respectively, compared with the best cost function in the initial set.

Since the optimal parameters for the previous cases were added to the initial sets

of the exercise cases, the cost function improvements for 2X and 3X exercise cases

are not as large as those for rest case. The total number of function evaluations,

48 and 65 for the 2X and 3X exercise cases, respectively, was less than that of the

Page 86: Surgical design for the Fontan procedure using computational fluid

62

rest case, which required 97 evaluations. In the second group using pulsatile inflow

conditions, the rest, 2X and 3X exercise cases required 74, 35 and 35 function

evaluations, respectively. The optimal designs for the 2X and 3X pulsatile cases

are identical owing to the fact that the 2X optimal solution was included in the

initial set for the 3X case, and no further improvement was found during the

optimization.

Figures 3.8 and 3.9 show the mean pressure and instantaneous velocity

magnitude of the optimal shapes using pulsatile inflow waveforms. In Figure 3.9

we observe that jets are formed in the IVC and bifurcated by the inner wall of the

two Y-shaped branches. The velocity is lower near the outer walls of the branches,

and flow separation and recirculation are observed, especially in the large graft.

Figure 3.10 shows more details of the velocity fields in the large and small

grafts. To illustrate why the optimal shape from the rest case is not optimal

at exercise, simulations were run using the optimal shape from the rest case with

both the 2X and 3X pulsatile inflow conditions. Compared with the optimal design

(small graft) for the 2X and 3X exercise cases, the large graft design has larger

separation regions near the outer walls of the branches. Making the assumption

that larger flow separation will result in higher energy dissipation compared to the

rest case explains why the optimal shapes for the 2X and 3X exercise cases have

smaller branches.

Although the optimal graft size differs with inflow conditions, there are

some common characteristics shared by all optimal shapes. First, XRPA and XLPA

Page 87: Surgical design for the Fontan procedure using computational fluid

63

Figure 3.8: Mean pressure for the optimal shapes under the rest, 2X and 3Xexercise levels using pulsatile waveforms.

almost reach their maximum bound (XRPA ≈ 3 cm and XLPA ≈ 6 cm) producing

a large angle between the branches. Second, ΔR and ΔL are close to their negative

maximum bound of -0.3 and -1.0, respectively, corresponding to convex curvatures.

A negative maximum value of ΔR and ΔL makes the divider that bifurcates the

IVC flow as sharp as possible for given values of XRPA and XLPA. It is likely that

more kinetic energy will be lost if the jet from the IVC impinges on an obtuse

divider and reflects. This agrees with our findings that designs with a wide span

and inward convex branches have higher energy efficiency than those with straight

branches and an obtuse divider.

If XRPA and XLPA are close to the lower bound of 1 cm, the two branches

will be merged into one common branch forming a T-junction or offset model.

Computational results demonstrate that the optimization algorithm leads to a Y-

graft design rather than a T-junction or offset in order to minimize energy loss.

This agrees well with our previous findings that demonstrated lower energy loss in

a patient specific Y-graft model.

Page 88: Surgical design for the Fontan procedure using computational fluid

64

(a) Instantaneous velocity magnitude of the optimal shape for the rest case. A respiratorycycle for the rest case is 2.86s.

(b) Instantaneous velocity magnitude of the optimal shape for the 2X case. A respiratorycycle for the 2X exercise case is 1.71s.

(c) Instantaneous velocity magnitude of the optimal shape for the 3X case. A respiratorycycle for the 3X exercise case is 1.33s.

Figure 3.9: Instantaneous velocity magnitude on the centerline cut plane of theoptimal shapes using pulsatile waveforms for the rest, 2X, and 3X cases withunconstrained optimization.

Table 3.4: Optimal parameters, cost function values and number of evaluations forthe unconstrained optimization using different inflow conditions. Parameters thatlie on the boundary are in bold.

inflow conditions J number of evaluations LIV C (cm) Dbranch(cm) XRPA (cm) ΔR (cm) XLPA (cm) ΔL (cm)

steady rest 0.0114 97 3.04 2.34 3.0 -0.3 6.0 -1.0steady 2X exercise 0.0119 48 2.95 2.02 3.0 -0.237 6.0 -1.0steady 3X exercise 0.0134 65 2.58 1.90 2.89 -0.29 6.0 -0.99

pulsatile rest 0.0122 74 3.58 2.28 2.2 -0.060 4.0 -0.34pulsatile 2X exercise 0.0137 35 3.5 1.8 3.0 -0.3 6.0 -1.0pulsatile 3X exercise 0.0156 35 3.5 1.8 3.0 -0.3 6.0 -1.0

Page 89: Surgical design for the Fontan procedure using computational fluid

65

(a) Comparison of velocity vector fields at peak IVC inflow for the rest (left)and 2X exercise (right) optimal shapes under the 2X pulsatile inflow condition,colored by velocity magnitude (cm/s).

(b) Comparison of velocity vector fields at peak IVC inflow for the rest (left) andthe 3X exercise (right) optimal shapes under the 3X pulsatile inflow condition,colored by velocity magnitude (cm/s).

Figure 3.10: Velocity vectors at peak IVC inflow for the exercise and rest optimalshapes at exercise conditions. Compared with the exercise optimal design (left),the large graft of the rest optimal design (right) results in more flow separationand causes more energy loss.

Page 90: Surgical design for the Fontan procedure using computational fluid

66

3.2.2 Polling comparison

To assess the robustness of the solution to the choice of polling method,

and assess the efficiency of the algorithm, two polling strategies were compared.

Unconstrained optimization using both LTMADS and OrthoMADS was performed

on the idealized Y-graft model. We used the same cost function J as above and

ran simulations under the 3X steady inflow condition. Taking into account the

computational expense, we performed five runs of LTMADS to compare with an

OrthoMADS instance.

First, only the poll step was employed and the search step was turned off.

Results are shown in Figure 3.11 and Table 3.5. We followed the same standard

as in Abramson et al.121 to evaluate the result given by OrthoMADS. The score

m indicates that the result given by OrthoMADS is as good or better than m

out of five LTMADS instances, with a relative precision of 0.1%. Thus the score

for OrthoMADS will be 5 if OrthoMADS gives the best result compared with all

five runs of LTMADS. According to table 3.5, OrthoMADS scored 3 and required

12% more function evaluations than the average number of evaluations required by

LTMADS. Considering that OrthoMADS employed 2n (n = 6) polling directions

and LTMADS here used n + 1 directions, OrthoMADS behaved remarkably well

in this case. In addition, Figure 3.11 shows that OrthoMADS found its best point

more quickly than all five LTMADS runs but took more function evaluations to

reach the convergence criteria because of the 2n polling directions.

Page 91: Surgical design for the Fontan procedure using computational fluid

67

10 20 30 40 50 60 70 80 90 100

0.0134

0.0136

0.0138

0.014

0.0142

0.0144

0.0146

0.0148

0.015

number of evaluations

J

LT1

LT2

LT3

LT4

LT5

ortho

Figure 3.11: Convergence history for 5 LTMADS and 1 OrthoMADS instanceswith poll only under the 3X steady inflow condition.

Table 3.5: Comparison results for 5 LTMADS and 1 OrthoMADS instances withpoll only under the 3X steady inflow condition. Parameters that lie on the bound-ary are in bold.

case J number of evaluations score LIV C (cm) Dbranch(cm) XRPA (cm) ΔR (cm) XLPA (cm) ΔL (cm)

LTMADS1 0.01341680 80 2.0 1.875 3.0 -0.3 6.0 -1.0LTMADS2 0.01340101 76 2.0 1.894 3.0 -0.3 5.9219 -1.0LTMADS3 0.01349008 77 2.0 1.95 3.0 -0.3 6.0 -0.9313LTMADS4 0.01344364 93 2.0 1.931 2.875 -0.3 6.0 -1.0LTMADS5 0.01344225 87 2.0 1.8 3.0 -0.3 6.0 -1.0average 0.01343936 83

OrthoMADS 0.01344225 93 3 2.0 1.95 3.0 -0.3 6.0 -1.0

In the second comparison, we added the search step back into the opti-

mization. Due to the search step, the number of function evaluations was reduced

by approximately 27% and 24% for LTMADS and OrthoMADS, respectively, as

shown in Figure 3.12 and Table 3.6. In this comparison, OrthoMADS and 3 in-

stances of LTMADS found identical solutions and the other 2 LTMADS instances

were worse than OrthoMADS. Thus, OrthoMADS scored 5 when the search step

was added to the optimization.

Page 92: Surgical design for the Fontan procedure using computational fluid

68

10 20 30 40 50 60 70 80

0.0134

0.0136

0.0138

0.014

0.0142

0.0144

0.0146

0.0148

0.015

number of evaluations

J

LT1

LT2

LT3

LT4

LT5

ortho

Figure 3.12: Convergence history for 5 LTMADS and 1 OrthoMADS instanceswith search and poll together under the 3X steady inflow condition.

Table 3.6: Comparison results for 5 LTMADS and 1 OrthoMADS instances withsearch and poll together under the 3X steady inflow condition. OrthoMADS foundthe best solution among 5 instances of LTMADS, with relative precision 0.1%.Parameters that lie on the boundary are in bold.

J number of evaluations score LIV C (cm) Dbranch(cm) XRPA (cm) ΔR (cm) XLPA (cm) ΔL (cm)

LTMADS1 0.01336113 55 3.0 1.8 3.0 -0.3 6.0 -1.0LTMADS2 0.01380732 57 3.375 1.8 2.875 -0.3 6.0 -0.5875LTMADS3 0.01336574 66 3.0 1.8 3.0 -0.3 6.0 -1.0LTMADS4 0.01337445 50 3.0 1.8 3.0 -0.3 6.0 -1.0LTMADS5 0.01338340 76 2.875 1.8 3.0 -0.3 6.0 -1.0average 0.01345840 61

OrthoMADS 0.01336049 71 5 3.0 1.8 3.0 -0.3 6.0 -1.0

3.2.3 Constrained optimization

Constrained optimization using the filter method was performed for the

rest, 2X and 3X cases using pulsatile inflow conditions. The mesh size Δmk was

not allowed to increase and n + 1 polling directions were employed. To reach

convergence, 65, 42 and 62 function evaluations were required for the rest, 2X

and 3X cases, respectively. Figure 3.13 shows the final results plotted as cost

function J vs. constraint violation function H for the constrained optimization at

Page 93: Surgical design for the Fontan procedure using computational fluid

69

rest. The models in Figure 3.13 correspond to the best feasible point and three

undominated points that form the filter. Three feasible points were identified for

the rest case. The best feasible point (marked by a large square) satisfies the

constraint and has the lowest cost function value. The lowest undominated point

is the design with the highest energy efficiency during the optimization. Figure

3.13 shows a clear trend that the best feasible point has smaller sized branches

merging together compared with the unconstrained results. As the conduit size

and bifurcation angle increase, the cost function is decreased but the constraint

violation function is increased. The model shown at the bottom right corner of

Figure 3.13 has the highest energy efficiency, and, not surprisingly, is similar to the

optimal shape from the unconstrained optimization for the rest case. The model

has large radius, wide span branches with a slightly smaller branch size than the

unconstrained optimization result. These results showed that the WSS constraint

has a strong effect in the constrained optimization for the rest case. In order to

minimize the areas of low WSS, the Y-graft became narrow with close branches.

We observe that a large conduit size and wide span branches cause considerable

areas of low WSS located on the outer wall of the bifurcation, which corresponds

to decreased energy dissipation.

With increasing flow rate, it was expected that the effect of the WSS con-

straint would diminish and more feasible points would be identified. Results for

the 2X exercise case, shown in Figure 3.14, are consistent with our expectations.

Eleven feasible points were identified. In the 2X exercise case, there is no substan-

Page 94: Surgical design for the Fontan procedure using computational fluid

70

0 2 4 6 8 10 120.014

0.016

0.018

0.02

0.022

0.024

0.026

H

J

dominated pointsthe best feasible pointfeasible pointsthe least infeasible pointundominated points

Figure 3.13: Final results of the constrained optimization plotted as cost functionJ vs. constraint function H for the rest case. The model in the upper left corneris the best feasible design and the model in right bottom corner is the design withhighest energy efficiency. Differences in shape among these models show a strongeffect of the WSS constraint for the rest case.

tial difference in geometry among these three points. Table 3.7 lists the parameters

of the best feasible and highest energy efficiency models for all constrained opti-

mization cases.

When the IVC flow rate increased to 3X, the WSS constraint became even

easier to satisfy and 32 feasible points were identified. As shown in Figure 3.15a,

points are tightly clustered near the axis corresponding to zero constraint violation.

We observe that the difference between the best feasible point and highest energy

efficiency point for the 3X exercise case becomes even less pronounced than the 2X

exercise case, but the size of branches is decreased, as shown in Table 3.7. All the

models in Figure 3.15b have wide-span and inward branches that are characteristic

of unconstrained optimization.

Page 95: Surgical design for the Fontan procedure using computational fluid

71

0 0.1 0.2 0.3 0.4 0.5 0.6

0.016

0.018

0.02

0.022

0.024

0.026

0.028

H

J

dominated pointsthe best feasible pointfeasible pointsthe least infeasible pointundominated points

Figure 3.14: Final results of the constrained optimization plotted as cost functionJ vs. constraint function H for the 2X exercise case. The number of feasiblepoints is increased to 11. The best feasible, least infeasible and highest energyefficiency models are listed from left to right. Different points in the filter plothave similar geometry. Results for the 2X exercise case show that the effect of theWSS constraint is weakened as inflow rates increase.

3.3 Discussion

In this work, we have coupled an efficient derivative-free optimization al-

gorithm with a 3D Navier-Stokes cardiovascular flow solver to optimize the shape

of a newly developed surgical design for the Fontan procedure. The application

of the constrained optimization to the Y-graft has demonstrated that the SMF

method can be efficiently coupled to pulsatile, 3D, cardiovascular simulations with

a reasonable number of design parameters. Compared with our previous work, this

work increased geometric complexity and number of design parameters, and added

constraints to the cardiovascular optimization framework. While computational

cost is still an issue, Y-graft optimization was performed efficiently, requiring at

Page 96: Surgical design for the Fontan procedure using computational fluid

72

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.80.016

0.018

0.02

0.022

0.024

0.026

0.028

0.03

0.032

H

J

dominated pointsthe best feasible pointfeasible pointsthe least infeasible pointundominated points

(a) Final filter for optimization with the 3X exercise condition.Points are tightly clustered near the axis corresponding to zeroconstraint violation. There are 32 feasible points.

0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05 0.055

0.0164

0.0165

0.0166

0.0167

0.0168

H

J

dominated pointsthe best feasible pointfeasible pointsthe least infeasible point undominated points

(b) A close-up of the filter for the 3X exercise case. Compared to the 2X exercisecase, there is even less difference between the best feasible point and undominatedinfeasible points.

Figure 3.15: Final results of the constrained optimization plotted as cost functionJ vs. constraint function H for the 3X exercise case. The number of feasible pointsis increased to 32.

Page 97: Surgical design for the Fontan procedure using computational fluid

73

Table 3.7: Comparison of the best feasible parameter and the highest energy effi-ciency points for the constrained optimization. Parameters that lie on the bound-ary are in bold.

inflow conditions LIV C (cm) Dbranch(cm) XRPA (cm) ΔR (cm) XLPA (cm) ΔL (cm) J H

rest best feasible 4.0 1.5 1.0 0.0 1.0 -0.725 0.0229 0rest highest energy efficiency 3.0 2.1 3.0 -0.3 4.75 -1.0 0.0140 8.5030

2X best feasible 3.5 1.8 3.0 -0.2 3.5 -0.175 0.0161 02X highest energy efficiency 3.5 1.8 3.0 -0.2 4.75 -0.175 0.0154 0.0140

3X best feasible 4.0 1.5 2.6 -0.3 5.7 -0.86 0.0164 03X highest energy efficiency 3.9 1.5 2.6 -0.3 6.0 -0.90 0.0163 0.0193

most 97 function evaluations for the unconstrained case (using LTMADS) and 65

evaluations for the constrained optimization case.

The unconstrained SMF method was tested using two polling strategies,

LTMADS and OrthoMADS, under the 3X steady inflow condition. Despite an

increased cost for each individual poll step, OrthoMADS behaved well overall.

OrthoMADS gave the best or one of the best repeatable results among the 5 runs

of LTMADS with n + 1 polling directions, with only a 12-16% increase in the

number of function evaluations using polling alone. Addition of the search step

saved 24% of the function evaluations for OrthoMADS compared to polling alone,

illustrating that the surrogate function is essential in improving efficiency of SMF.

OrthoMADS as an alternative polling strategy will be most suitable when the

computational cost can be reduced efficiently using a good quality surrogate, and

the number of parameters is not to large. The polling comparison confirmed that

the SMF method is robust and efficient, producing similar results with two different

polling strategies in both cost function and number of function evaluations.

Shape optimization with and without constraints has identified several in-

Page 98: Surgical design for the Fontan procedure using computational fluid

74

teresting trends in the resulting optimal Y-graft under various flow conditions.

Results from the unconstrained optimization cases revealed that optimal Y-graft

designs differed between rest and exercise, as well as steady and pulsatile flow.

This further illustrates that steady flow simulations alone are not sufficient to test

new surgical designs, and that pulsatility will likely produce different conclusions.

Under low flow rates, optimal shapes tend to have graft branches that are larger

than the IVC. However, under exercise conditions the maximum graft size does not

necessarily have the highest energy efficiency and larger areas of flow separation are

observed. The observed trend is that optimal designs for the exercise conditions

tend to have smaller graft branches than those for rest conditions. Additionally,

the optimal shape for the 2X exercise case is also a local optimum for the 3X cases.

We employed a filter method for constrained optimization that illustrated

the trade-off between the energy cost function and the WSS constraint. A large

graft branch size and wide span branches resulted in high energy efficiency but

larger areas of low WSS. Results from the three constrained optimization cases

showed that the impact of the WSS constraint decreases with increasing exercise

level. While the WSS constraint forced the branches close together in the rest

case, the Y-graft was clearly the design of choice at exercise. Because patients

(particularly children) spend a large percentage of time in a mild exercise state, the

choice of design should ultimately combine both rest and exercise considerations.

The critical WSS value we used in this study has not yet been validated due

to lack of necessary biological and clinical data related to thrombosis in the Fontan

Page 99: Surgical design for the Fontan procedure using computational fluid

75

graft. However, results of the constrained optimization have clearly demonstrated

that the WSS constraint and exercise conditions can influence the Fontan geometry.

The examples with the WSS constraint demonstrate that constrained optimization

allows us to consider more than one factor in the surgical design. Further studies

are needed to validate the link between the formation of thrombosis and areas of

low WSS. The critical WSS value which may cause thrombosis is important for

Fontan patients, but is still unknown and may differ from one patient to another. In

addition, high shear can also increase platelet aggregation.131 As more clinical data

is obtained and correlated with simulation results, more accurate knowledge of the

mechanical factors that cause thrombosis will be determined, and this information

should be incorporated into future optimization work.

The optimization results presented here agree with previous findings that

the Y-graft is likely to offer superior performance compared to traditional designs

that are currently used. Compared with a trial-and-error design method, opti-

mization offers improved confidence that the Y-graft is a promising design for the

Fontan surgery, due to the known convergence theory and global search nature

of the SMF optimization algorithm. Despite this, clinical implementation of the

Y-graft design will need to take into account several additional clinical and surgi-

cal considerations. Our previous work has confirmed that there is sufficient space

anatomically for the Y-graft by embedding the Y-graft model in the image data

to ensure that it did not overlap with anatomic features.2 However, a large Y-

graft branch size may result in technical difficulties in the actual operation. For

Page 100: Surgical design for the Fontan procedure using computational fluid

76

example, oversized grafts may be more challenging to connect smoothly with the

PAs in surgical practice. The results presented in this work demonstrate that the

choice of Y-graft design should balance energy loss, areas of low WSS and technical

feasibility. Thus, a design with slightly smaller branches than the IVC can achieve

a balance between energy efficiency at different exercise levels and the potential

thrombosis risks. However, weighting of these factors should be investigated and

related to clinical data in future work. Future work should also consider additional

clinical factors as cost functions or constraints in multiple objective optimization,

including flow distribution to the right and left PAs, oxygen saturation levels, and

technical difficulty.

3.4 Limitations

A major limitation of this study is the use of idealized cylindrical models

that we believe results in unrealistically high energy efficiency, and small differences

in efficiency between models. Since this study is a first step towards applying

a formal optimization method to the Fontan procedure, and parameterization of

patient specific models presents significant challenges, optimization on an idealized

model was a necessity in the present work. Moreover, our previous patient specific

simulations indicate that variations of the graft geometry (9 mm vs. 12 mm)

and inflow conditions (rest vs. exercise) can cause far more significant efficiency

differences that those observed in this work. We therefore expect that differences

Page 101: Surgical design for the Fontan procedure using computational fluid

77

between models will be significantly accentuated in the patient specific cases, and

this will be a focus of future study.

Another limitation of this work was that the constant parameters used to

define the model geometry represent a single patient. Future work should optimize

multiple patient specific models. The influence of compliant walls and growth of

the patient also merit consideration in future optimization work. In addition, the

use of a non-Newtonian assumption for the fluid, has been shown in some previous

work132,133 to affect WSS and flow profiles.

Future work should also consider the effect of uncertainties in simulation

parameters, inflow rates, and implementation of the surgical design. Robust opti-

mization that accounts for uncertainties could identify designs that are less sensi-

tive to small changes in design parameters, thus allowing for a “fudge-factor” in

surgical implementation.134,135 Efficiency of the optimization algorithm could also

be improved by performing function evaluations in parallel for each search and poll

step, which would lead to significant savings in computational time.

3.5 Acknowledgments

This work was supported by the American Heart Association, a Burroughs

Wellcome Fund Career Award at the Scientific Interface and a Leducq Founda-

tion grant. The authors wish to thank Prof. John Dennis, Prof. Charles Audet,

Dr. Sebastien Le Digabel, cardiac surgeons Dr. Mohan Reddy and Dr. John

Page 102: Surgical design for the Fontan procedure using computational fluid

78

Lamberti for sharing their expertise on optimization and Fontan surgical proce-

dures. We gratefully acknowledge the use of software from the Simvascular open

source project(http://simtk.org),114 Cardiovascular Simulation, Inc., as well as the

expertise of Dr. Nathan Wilson and Prof. Charles Taylor.

Chapter 3, in full, is a reprint of the material as it appears in Numerical

Grid Generational in Yang, W., Feinstein, J. A. and Marsden, A. L. Constrained

Optimization of an Idealized Y-shaped Baffle for the Fontan Surgery at Rest and

Exercise. Comput. Meth. Appl. Mech. Engrg. 2010;199:2135-2149. The disserta-

tion author was the primary investigator and author of this paper.

Page 103: Surgical design for the Fontan procedure using computational fluid

Chapter 4

Hemodynamic Evaluations for

traditional and Y-graft Fontan

Geometries

The impact of geometry on Fontan hemodynamics is now widely accepted

in the engineering and clinical communities.13, 45, 50, 51 Recent advances in compu-

tational fluid dynamics (CFD), computer aided design (CAD), magnetic resonance

imaging (MRI), and fluid structure interaction methods (FSI) for blood flow mod-

eling have led to studies of Fontan hemodynamics and surgical design progressing

from idealized to patient specific models, steady to unsteady flow, and trial and

error to optimal design.2,4, 44, 57, 136

Although the preliminary studies showed the Y-graft design to improve

hemodynamics overall, it has not yet been confirmed that the superiority of the

79

Page 104: Surgical design for the Fontan procedure using computational fluid

80

Y-graft is universal.2 The purpose of the study in Chapter 4 is to evaluate the

potential performance of the Y-graft Fontan procedure in simulation as a step

towards clinical implementation. Multiple virtual patient models are used, and

multiple parameters51 are evaluated for each, with particular attention paid to the

hepatic flow distribution.

4.1 Methods

4.1.1 Geometrical model construction

Virtual surgery was performed on five models by implanting a Y or tube-

shaped graft into patient specific Glenn models5 constructed from MRI images. All

patients were consented as part of an institutional review board approved protocol

at Lucile Packard Children’s Hospital (Stanford University). Model construction

was performed using a custom version of the open source Simvascular software

package,114 as in our previous work.2

Geometric parameters were chosen based on our previous optimization

study.136 A Y-graft with a 20 mm diameter trunk and 15 mm diameter branches

was chosen for all patients. The T-junction and offset models were constructed

with a 20 mm diameter tube-shaped graft following common clinical practice, and

resulting models are shown in Figure 4.1.

In the virtual implantation, the centerline paths of the SVC and PA anatomy

of the original Glenn models were left unchanged. The space created by the defla-

Page 105: Surgical design for the Fontan procedure using computational fluid

81

tion of the right atrium (RA) during the Fontan procedure was accounted for in IVC

placement during virtual surgery. In addition, the PA segmentations are enlarged

to match the baffle size in order to avoid creating unrealistic stenoses. Although

this is purely a simulation study, all models were constructed under guidance of

a surgeon to replicate as closely as possible a realistic surgical implementation of

the intended design.

Patient B has a left PA (LPA) stenosis which is commonly observed in

Fontan patients due to aortic arch override. Three potential Y-graft designs were

proposed for this patient model. In Y-graft I, the stenosis is relieved by placing

the anastomosis of the left branch at the stenosis. In Y-graft II, the left branch

is anastomosed distal to the stenosis, without relieving it. In Y-graft III, the

left branch is also anastomosed distal to the stenosis, but the stenosis is relieved.

Patient E has heterotaxy and a right PA (RPA) stenosis. Instead of an LPA offset

connection, the baffle is anastomosed to the RPA in a mirror image of the other

patients, denoted as Offset I. The stenosis is relieved in the Y-graft I, T-junction

and Offset I designs. An LPA-offset connection without relieving the stenosis (II)

is also constructed. We redesigned the Y-graft for patient E, denoted as Y-graft

II, by bringing the left branch closer to the SVC-PA junction and less anteriorly

convex, based on the simulation results of Y-graft I.

Page 106: Surgical design for the Fontan procedure using computational fluid

82

A

B

C

D

E

Y-graft

Y-graft I Y-graft II Y-graft III

Y-graft

Y-graft I

Y-graft

Y-graft II

T-junction

T-junction

T-junction

T-junction

offset

offset

offset

offset

offset I offset II

T-junction

Glenn Fontan

Figure 4.1: Original Glenn models and variations of Fontan geometries for fivepatients. The Y-graft includes a 20 mm trunk and two 15 mm branches. The sizeof the tube-shaped graft is 20 mm. Patients B and E have a stenosis in the LPAand RPA, respectively, denoted by arrows.

Page 107: Surgical design for the Fontan procedure using computational fluid

83

4.1.2 Flow simulation and boundary conditions

Flow simulations were performed with a stabilized finite element Navier-

Stokes solver,94 assuming rigid walls and Newtonian flow with a density of 1.06g/cm3

and viscosity of 0.04 g/(cm s). MeshSim (Simmetrix, Inc., Clifton Park, NY)

was employed to generate tetrahedral meshes automatically and anisotropic mesh

adaptation was performed to ensure mesh convergence. Final meshes consisted of

approximately 1 to 1.5 million elements.

During MR imaging, phase contrast magnetic resonance imaging (PCMRI)

slices were acquired in the SVC, IVC, LPA and RPA for each Glenn patient.

The SVC waveform from PCMRI was applied directly to the SVC inflow face by

mapping it to a parabolic profile. Because the IVC flow waveforms were acquired

in the Glenn patients prior to surgery, when the IVC was still connected to the RA,

these waveforms exhibited higher cardiac pulsatility and less respiratory pulsatility

compared to typical Fontan patients.137 To account for this, the amplitude of the

IVC waveform of each Glenn patient was scaled to match typical Fontan data using

four patients from a previous study, while keeping the mean the same. Additionally,

a respiratory model4 was superimposed on the scaled IVC waveform following our

previous work.

A three element Windkessel circuit model (RCR, resistor-capacitor-resistor)85,96, 138

was applied at each outlet. Predicting changes in pulmonary resistance following

Fontan surgery remains an open question. However, since this study aims to model

Page 108: Surgical design for the Fontan procedure using computational fluid

84

immediate post-operative flow conditions, we have assumed that the downstream

resistances do not change significantly in the short period after the surgery. There-

fore, the same parameters for the Windkessel model were employed for the Glenn

and Fontan simulations in all patients, so that the comparison of post-operative

states among patients remains consistent.

Exercise flow conditions were simulated by increasing IVC flow and decreas-

ing outflow resistances, following our previous work.4 The mean IVC flow rate was

increased by 2 and 3 times (referred to as 2X and 3X in the text) to simulate

exercise conditions and the total resistances of each branch was decreased by 5%

and 10%, respectively.

4.1.3 Determination of performance parameters

Mean SVC pressure, wall shear stress and power loss were computed with

standard methods.2 To quantify the hepatic flow distribution in different surgical

designs, Lagrangian particle tracking139 is performed (Figure 4.3).

It is well known that the ratio of IVC/SVC flow and LPA/RPA resistances

vary widely among patients. While ideally one would like to achieve a 50/50

hepatic flow split for all patients, this will not be possible in every case. A simple

conservation of mass analysis allows us to determine the theoretical optimum for

hepatic flow distribution in a given situation.

We assume that hepatic flow is well mixed in the IVC such that the hepatic

and IVC flow distributions are the same, and that the theoretical optimum for the

Page 109: Surgical design for the Fontan procedure using computational fluid

85

hepatic flow distribution is the split closest to 50/50 that satisfies conservation of

mass.

Let us derive the dependence of the hepatic flow distribution on the vena

caval and pulmonary distributions (see Figure 4.2 for a schematic representation

of the Fontan configuration). Basic conservation of mass dictates that the ratio

between the RPA flow and the LPA flow can be defined as

QRPA

QLPA=

QIV C · x+QSV C · yQIV C · (1− x) +QSV C · (1− y)

=fs

1− fs, (4.1)

where Q is the flow rate, fs is the fraction of total inflow going to the RPA, x is

the fraction of hepatic flow going to the RPA, and y is the fraction of SVC flow

going to the RPA. Then, the hepatic flow distribution can be expressed by

x = fs +QSV C

QIV C

· (fs − y). (4.2)

Equation 4.2 shows that the hepatic flow distribution is a function of the

inflow rates, outflow split and percentage of SVC flow going to the RPA.

For a given hepatic flow split x, the bounds of the RPA flow are

QIV C · x ≤ QRPA ≤ QIV C · x+QSV C . (4.3)

Thus, a 50/50 hepatic flow split can not exist when QRPA < QIV C · 0.5 or

QRPA > QIV C · 0.5 + QSV C . Since QRPA is given as boundary condition, when a

Page 110: Surgical design for the Fontan procedure using computational fluid

86

Figure 4.2: Based on conservation of mass, we have QRPA = QIV C · x + QSV C · yand QLPA = QIV C · (1−x)+QSV C · (1− y), where x is the fraction of hepatic flowgoing to the RPA, and y is the fraction of SVC flow going to the RPA.

Table 4.1: MRI inflow rates, MRI outflow splits and the theoretical optimal hepaticflow splits (TOHFS) at rest for the five study patients.

Patient Age (yr.) BSA (m2) IVC (ml/s) SVC (ml/s) RPA/LPA flow split TOHFS (RPA/LPA)

A 4.8 0.63 13.4 14.4 81/19 61/39B 3.9 0.66 12.6 15.3 63/37 50/50C 2.8 0.56 6.3 15.2 54/46 50/50D 3.0 0.61 14.8 27.8 55/45 50/50E 3.0 0.74 12.7 19.4 70/30 50/50

50/50 split is infeasible, the theoretical optimum for the hepatic flow distribution

is defined as the value closest to 50/50 achieved by taking y = 0 or y = 1, i.e.

QIV C · x = QRPA or QIV C · x+QSV C = QRPA

Table 4.1 lists the theoretical best hepatic flow distribution for the five

patients in our study. We observe that a perfect 50/50 split can theoretically exist

for four out of the five patients, with patient A as the exception, for whom the

RPA receives more than 80% of the total inflow.

The main focus of this study is the evaluation of Fontan designs in the im-

mediate post-operative period. However, it is unlikely that pulmonary resistances

Page 111: Surgical design for the Fontan procedure using computational fluid

87

t=0s t=0.24s t=0.6s

t=2.4s t=4.8s t=12s

RPA LPA

Figure 4.3: Visualization of the particle tracking in the model Y-graft II for patientB. Particle tracking is terminated when particles are washed from the model.

remain constant over time for most Fontan patients due to age, growth, and remod-

eling. We therefore asses the robustness of the hepatic flow distribution to changes

in the pulmonary flow split for different surgical designs. For each patient, the

Y-graft, offset and T-junction designs were analyzed under rest conditions for a

range of pulmonary resistances to compare the robustness of different designs.

4.2 Results

4.2.1 Hepatic flow distribution

The percentages of hepatic flow to the RPA and LPA, and differences from

the theoretical optima at rest, 2X and 3X exercise conditions are shown in Figure

4.4. In patient A, the Y-graft design is closest to the theoretical optimum, even

though patient A’s right lung receives more than 80% of the total venous return.

Page 112: Surgical design for the Fontan procedure using computational fluid

88

To illustrate how the geometry influences the hemodynamics of hepatic flow dis-

tribution, the velocity fields for three patients are shown in Figure 4.5. In the

T-junction design of patient A, the SVC jet clearly blocks the hepatic flow enter-

ing the LPA and skews it to the RPA. In contrast, the Y-graft design effectively

mitigates this effect.

In patient B, Y-graft I hemodynamics shows that relieving the stenosis with

a proximal anastomosis allows the SVC jet to enter the baffle and hinder hepatic

flow. However, Y-graft II, which leaves the stenosis intact, achieves a nearly perfect

50/50 distribution at rest. Overall for patient B, Y-graft II has the best hepatic

distribution at rest, while the offset design performs best at exercise.

In patient C, the T-junction and offset designs skew hepatic flow to the

RPA and LPA, respectively. In both patients C and D, when the pulmonary flow

split is close to 50/50, the offset design has poor performance with a highly skewed

hepatic flow split. Overall for patient C, the Y-graft design has the best hepatic

performance though none of the designs achieve the theoretical optimum.

In patient D, although the SVC anastomosis is inclined to channel flow to

the LPA, Figure 4.5 shows that the lower section bc of the SVC is ineffective in

changing the direction of flow. Improvements are observed in both the Y-graft and

T-junction during exercise, and the Y-graft achieves a nearly 50/50 split in the 3X

exercise case. Overall for patient D, the T-junction does best at rest, while the

Y-graft does best at exercise.

In patient E, the T-junction, Offset I and Y-graft I designs skew hepatic

Page 113: Surgical design for the Fontan procedure using computational fluid

89

flow strongly to the RPA with a 70/30 (RPA/LPA) pulmonary flow split because

of the smooth graft and the low distal RPA resistance. For the T-junction design,

hepatic flow distribution remains constant with exercise, whereas Y-graft I pro-

gressively improves the distribution from rest to exercise. The skewed hepatic flow

is corrected in the Y-graft because the SVC jet suppresses hepatic flow entering

the RPA from the right branch (Figure 4.5). Overall for patient E, the offset II

design results in the best hepatic flow distribution at rest and exercise.

Robustness test

The hepatic flow distribution in the Y-graft, offset and T-junction designs

for three different pulmonary flow splits, and their averaged deviations for a 25%

change in flow split, are shown in Figure 4.6. The optimal design depends heavily

on the pulmonary flow split for individual patients. The Y-graft and T-junction

designs are relatively more robust in two and three patients, respectively. Although

the T-junction design underperforms for most patients with the original flow split,

deviations in hepatic flow distribution are smaller than those of the offset design.

4.2.2 SVC pressure

Pressure levels increase in all models under exercise conditions, compared

to the values at rest. The SVC pressure levels in the Y-graft are the same or

slightly lower than those in the T-junction and offset designs at rest and exercise

conditions in five patients (Table 4.2). This finding agrees with our previous work

Page 114: Surgical design for the Fontan procedure using computational fluid

90

0%5%

10%15%20%25%30%35%40%45%50%

Y-gra� T-junc�on Offset

rest2X3X

0%

5%

10%

15%

20%

25%

30%

35%

40%

45%

50%

Y-gra� I Y-gra� II Y-gra� III T-junc�on Offset

rest2X3X

0%5%

10%15%20%25%30%35%40%45%50%

Y-gra� T-junc�on Offset

rest2X3X

0%5%

10%15%20%25%30%35%40%45%50%

Y-gra� T-junc�on Offset

rest2X3X

% IV

C flo

w

Diff

. fro

m th

eor.

opt.

Patient A

Patient E

Patient D

Patient C

Patient B

0%10%20%30%40%50%60%70%80%90%

100%

Y-gra� T-junc�on Offset Theore�cal op�mum

IVC-RPAIVC-LPA

0%10%20%30%40%50%60%70%80%90%

100%

Y-gra� I Y-gra� II Y-gra� III T-junc�on Offset Theore�cal op�mum

IVC-RPAIVC-LPA

0%10%20%30%40%50%60%70%80%90%

100%

Y-gra� T-junc�on Offset Theore�cal op�mum

IVC-RPAIVC-LPA

0%10%20%30%40%50%60%70%80%90%

100%

Y-gra� T-junc�on Offset Theore�cal op�mum

IVC-RPAIVC-LPA

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

Y-gra� I Y-gra� II T-junc�on Offset I Offset II Theore�cal op�mum

IVC-RPAIVC-LPA

0%

5%

10%

15%

20%

25%

30%

35%

40%

45%

50%

Y-gra� I Y-gra� II T-junc�on Offset I Offset II

rest2X3X

Figure 4.4: Left: Hepatic flow distribution at rest. Right: Differences (percentageof the IVC flow) from the theoretical optima for each design at rest and exercise.Note that the theoretical optima for patient A at rest, 2X and 3X are 61/39, 70/30and 72/28, respectively, and that a 50/50 split can not be achieved in theory.

Page 115: Surgical design for the Fontan procedure using computational fluid

91

a

bc

T-junction Y-graft

T-junction Y-graft

Y-graft IIY-graft I

A

E

D

Right Left Right Left

Figure 4.5: Time-averaged velocity vectors in the Y-graft and T-junction modelsfor patients A, D and E. In the T-junction design for patient A, the SVC jetblocks the hepatic flow entering the LPA. In patient D, most SVC flow is directedto the RPA due to a curved SVC. In patient E, Y-graft II improves the hepaticflow distribution by having a straight proximal branch for the RPA, in which theSVC jet blocks hepatic flow going to the RPA from the right branch, compared toY-graft I.

Page 116: Surgical design for the Fontan procedure using computational fluid

92

Patient EPatient D

Patient CPatient BPatient A

0%10%20%30%40%50%60%70%80%90%

100%

0% 20% 40% 60% 80% 100%

Y-gra�Offset

% inflow to RPA

% IV

C flo

w to

RPA

0%10%20%30%40%50%60%70%80%90%

100%

0% 20% 40% 60% 80% 100%

Y -gra� IY-gra� IIY-gra� IIIOffset

% inflow to RPA

% IV

C flo

w to

RPA

0%10%20%30%40%50%60%70%80%90%

100%

0% 20% 40% 60% 80% 100%

Y-gra�Offset

% inflow to RPA

% IV

C flo

w to

RPA

0%10%20%30%40%50%60%70%80%90%

100%

0% 20% 40% 60% 80%

Y-gra�Offset

% inflow to RPA

% IV

C flo

w to

RPA

0%10%20%30%40%50%60%70%80%90%

100%

0% 20% 40% 60% 80% 100%

Y-gra� IIOffset II

% inflow to RPA%

IVC

flow

to R

PA

Figure 4.6: Hepatic flow distribution changes with variations in pulmonary flowsplit. Patients’ original pulmonary flow splits are marked by the arrows at the xaxis. The table shows the averaged deviations with respect to the original hepaticflow distribution for a 25% change in pulmonary flow split.

which identified a pressure shielding effect causing lower SVC pressure with the

Y-graft design due to the lack of a head-on flow collision in the junction.2

4.2.3 Power loss

Table 4.2 gives the power loss values for the Y-graft, T-junction and offset

designs for each patient. Energy loss increases with increasing exercise levels. The

T-junction designs result in the highest energy losses in most patients due to a

direct flow collision, agreeing with previous studies.

4.2.4 Wall Shear Stress

Mean WSS values on the IVC graft generally increase with increasing inflow

rates (Table 4.2). Exceptions to this are the T-junction and offset designs for

Page 117: Surgical design for the Fontan procedure using computational fluid

93

Table 4.2: Mean SVC pressure (mmHg), power loss (mW) and mean (in space)WSS magnitude (dynes/cm2) on the IVC graft for the Fontan models. Comparedto the best Y-graft design for the same patient, increases in power loss for theT-junction and offset designs are also shown.

patient modelSVC pressure power loss mean WSS

rest 2X 3X rest 2X 3X rest 2X 3X

AY-graft 15.3 19.4 22.9 5.1 14.5 24.9 2.3 4.6 7.2

T-junction 15.5 19.6 23.4 16% 14% 14% 4.5 5.9 7.9Offset 15.4 19.7 23.4 10% 15% 17% 1.9 4.5 7.4

B

Y-graft I 12.9 16.7 19.9 2.4 5.4 8.6 6.5 6.5 8.7Y-graft II 12.8 16.6 19.8 2.0 4.6 7.5 2.2 5.2 7.6Y-graft III 12.8 16.6 19.8 2.0 4.7 7.6 2.3 4.6 7.0T-junction 13.1 17.1 20.5 60% 70% 76% 6.1 6.8 9.0

Offset 13.0 17.0 20.4 40% 57% 64% 3.1 5.1 7.4

CY-graft 10.4 11.9 13.0 1.8 2.9 4.1 3.0 3.3 4.2

T-junction 10.4 11.9 13.0 0% 7% 5% 5.5 5.4 5.3Offset 10.4 11.9 13.1 0% 10% 15% 3.0 2.2 2.4

DY-graft 10.8 13.5 15.8 5.8 12.8 19.9 3.8 6.9 9.6

T-junction 10.9 13.7 16.1 10% 15% 17% 4.0 4.5 5.9Offset 10.8 13.6 16.0 5% 11% 16% 2.5 6.0 9.8

E

Y-graft I 22.9 27.7 31.7 5.14 10.8 17.1 3.9 5.7 8.0Y-graft II 22.8 27.6 31.7 4.7 9.7 15.6 2.9 4.7 7.15T-junction 22.9 27.8 31.8 14% 15% 15% 5.4 7.1 10.2Offset I 22.8 27.6 31.6 -1% -2% -5% 3.1 6.6 6.8Offset II 23.4 28.6 33.1 65% 86% 90% 2.7 4.6 8.1

Page 118: Surgical design for the Fontan procedure using computational fluid

94

A

T-junction OffsetY-graft

C

Figure 4.7: Contours of time-averaged WSS (dynes/cm2) at rest for patients Aand C.

patient C, in which the SVC jet enters the tube-shaped graft and increases the

WSS at a low IVC flow rate. The Y-graft and offset designs result in lower mean

WSS values on the graft than the T-junction design. This is consistent with the

fact that the T-junction design usually causes more energy dissipation. Time-

averaged WSS at rest for two representative patients is shown in Figure 4.7. The

other three patients exhibited similar behavior.

4.2.5 Averaged results

By averaging the hepatic distributions and power losses in the best per-

forming Y-graft, T-junction and offset designs over the five patients in the study

Page 119: Surgical design for the Fontan procedure using computational fluid

95

02468

101214161820

rest 2X 3X

Y-gra�T-junc�onOffset

Ave.

Pow

er lo

ss (m

W)

*

*

*

0%5%

10%15%20%25%30%35%40%45%50%

rest 2X 3X

Y-gra�T-junc�onOffset

Ave.

diff

. fro

m th

eor.

opt.

** *

Figure 4.8: Averaged differences from the theoretical optima and power losses overfive patients. The best performing of the Y-graft and offset designs for patients Band E are used. The differences between the Y-graft and T-junction designs arestatistically significant (∗P < 0.05).

(see Figure 4.8), we found that the Y-graft design has the lowest average difference

from the theoretical optima under rest and exercise conditions (i.e., the Y-graft

design results in more even hepatic flow distribution than other designs overall),

and that the distribution approaches the theoretical optima with increasing exer-

cise level. Significant differences (P < 0.05) are found in the mean hepatic flow

distribution and power loss between the Y-graft and T-junction designs.

4.3 Discussion

In this study, a multi-parameter approach was employed to evaluate three

Fontan designs in five patients, with particular focus on hepatic flow distribu-

tion. This multiple-patient series has demonstrated that the Y-graft design can

significantly improve hepatic flow distribution and moderately improve energy loss

and SVC pressure. However our results emphasize that no one-size-fits-all design

Page 120: Surgical design for the Fontan procedure using computational fluid

96

achieves satisfactory hepatic distributions in all patients. The hemodynamics af-

fecting flow distribution are non-intuitive in many cases, and small differences in

geometry can dramatically influence results. Because these subtle changes are not

easily elucidated with standard imaging modalities, simulations should be used to

determine the best candidates for a Y-graft preoperatively, and to refine the graft

design for each patient.

4.3.1 Hepatic flow distribution

Hepatic flow distribution in the Fontan is driven by multiple factors. The

SVC jet can prevent hepatic flow from reaching the LPA, and this was the leading

cause of uneven hepatic flow distribution in our study, agreeing with previous re-

sults of Dasi et al.53 Although previous studies in idealized geometries showed that

the T-junction design effectively mixes the IVC and SVC flows and distributes the

hepatic flow evenly,69 these phenomena were not observed uniformly in the patient

specific geometries in our study. The Y-graft design distributes hepatic flow more

evenly than the T-junction in most patients by avoiding a straight flow collision,

but the wrong choice of Y-graft may lead to unfavorable hemodynamics. The

LPA-offset design generally achieves satisfactory hepatic distribution in patients

with high LPA resistance, but an unfavorable distribution in patients with an equal

pulmonary flow split. This result is consistent with previous work of Bove et al.52

which shows that the IVC flow in the total cavopulmonary connection and tradi-

tional extracardiac Fontan models is skewed to the LPA with a pulmonary flow

Page 121: Surgical design for the Fontan procedure using computational fluid

97

split close to 50/50.

The non-intuitive result for patient D demonstrates that the geometry of

the SVC and relevant boundary conditions play an important role in distributing

the IVC and SVC flows. The hepatic flow distribution depends on the SVC flow

distribution (Equation 4.2) with an inverse relationship between the percentages of

hepatic and SVC flow going to the RPA. Thus the effect of the SVC-PA anastomosis

to the hepatic flow distribution should be carefully considered in surgical planning

for the Glenn procedure.

Although the maximum change in hepatic flow distribution from rest to

exercise was less than 20%, values generally approached their theoretical optima

during exercise. When the IVC flow rate is increased during exercise, the inter-

action between the two caval flows in the T-junction design generally enhances

mixing and distributes hepatic flow more evenly. Salim et al.140 showed that the

contribution of IVC flow to the cardiac output increases from 45% at 2.5-3 years

old to 65% (the adult value) at 6.6 years old. In our study, 45-48% of systemic

venous return is contributed by the IVC in two patients (mean 4.3 years) while

the IVC contribution is between 29-35% in three patients (mean 2.9 years). Thus,

simulations under exercise conditions may reveal some trends in the hepatic flow

distribution with increasing age, though changes in distribution of caval flow to

the PAs with age are still unknown.33

Our robustness test confirms that hepatic flow distribution depends strongly

on the pulmonary flow split. For some patients, the Y-graft design is relatively

Page 122: Surgical design for the Fontan procedure using computational fluid

98

more robust than the offset design for a wide range of flow conditions with less

chance to skew all IVC flow to one lung but this characteristic is not universal

for all Y-graft designs. Although the T-junction design skews hepatic flow with

patients’ original pulmonary flow splits, it performs better under the conditions

with a high RPA resistance due to a slight RPA-offset for the anastomosis. Thus,

the optimal working condition for the LPA-offset design is opposite. No design

emerged as the clear winner over a wide range of flow splits.

4.3.2 Power loss

The mechanism of energy dissipation in the Fontan has been well dis-

cussed.2,44 Compared to traditional designs, the Y-graft design reduces energy

dissipation by bifurcating the IVC to decrease flow competition. In our previous

work,2 the Y-graft design demonstrated reduced energy loss in a single patient

specific model. This study further compared the energy loss between the Y-graft

and traditional designs by examining multiple patients. The Y-graft design re-

duces energy loss at rest by 5-27% in four out of five patients, compared to the

offset design. The differences in energy loss are more pronounced during exercise,

agreeing with previous work.

4.3.3 SVC pressure

It is clinically well accepted that lower Fontan pressure generally correlates

with better outcomes. Overall, the Y-graft design offers moderate reductions in the

Page 123: Surgical design for the Fontan procedure using computational fluid

99

SVC pressure, compared to the T-junction and offset designs. In the T-junction

design, a higher SVC pressure is usually required to overcome competing flow from

the IVC. However the average SVC pressure differences are not as pronounced as

those reported in our previous work,2 likely because IVC flow rates were lower in

the younger patients in this study. This trend suggests that the significance of the

pressure shielding effect observed with the Y-graft may increase as patient’s age

and their relative IVC flow increases.

4.3.4 Wall shear stress

The Y-shaped grafts have lower mean WSS values than the tube-shaped

grafts. These results are qualitatively consistent with our previous work on ide-

alized Y-graft shape optimization,136 which shows a trade-off between energy effi-

ciency and areas of low WSS. While the WSS values in the Y-graft are generally

comparable to values in offset designs, further investigation into this issue is war-

ranted, and patients with known thrombotic tendency should likely be excluded as

candidates for the Y-graft procedure. The impingement of SVC flow results in a

relatively high WSS region in the intervening segment of the PA (Figure 4.7), which

does not suggest an increased likelihood of flow stasis and thrombus formation in

that region.

Page 124: Surgical design for the Fontan procedure using computational fluid

100

4.3.5 Ranking

Based on power loss and hepatic flow distribution, the proposed surgical

designs are ranked for each patient (Table 4.3) at rest and exercise. For energy loss,

the Y-graft design is superior to the others in four out of five patients regardless

of rest or exercise conditions, and Offset I wins by a narrow margin in patient

E. However, more variations emerge in the ranking based on the hepatic flow

distribution.

For patients A and C, the Y-graft design is clearly the final winner, as

it was uniformly ranked first in both energy loss and hepatic flow distribution.

For patient B, the offset design provides 10% improvement in the hepatic flow

distribution at exercise, with better robustness to the pulmonary flow split, but

produces over 40% more power loss than Y-graft II. We therefore select Y-graft

II as the final candidate for patient B. For this patient, we note that Y-graft III

overcomes the disadvantage of Y-graft II in robustness, while keeping the power

loss almost unchanged. If we were to weight the robustness more heavily, then

Y-graft III would likely be the recommended choice. For patient D, the Y-graft

design is chosen because of lower energy losses and progressive improvements in the

hepatic flow distribution during exercise, but the T-junction would be preferred

if the robustness were weighted more heavily than the power loss. For patient

E, Y-graft II is chosen because it is well balanced between energy loss and the

hepatic flow distribution, and it is more robust to changes in the pulmonary flow

Page 125: Surgical design for the Fontan procedure using computational fluid

101

Table 4.3: Ranking of energy loss and hepatic flow distribution for each patient.The ranking of the hepatic flow distribution is based on the differences from thetheoretical optima. In patients C and D, there are two designs tied for the hepaticflow distribution.

patient modelpower loss hepatic flow

recommended designsrest 2X 3X rest 2X 3X

AY-graft 1 1 1 1 1 1 �

T-junction 3 2 2 3 3 3Offset 2 3 3 2 2 2

B

Y-graft I 3 3 3 5 5 5Y-graft II 1 1 1 1 2 2 �Y-graft III 2 2 2 3 3 3T-junction 5 5 5 5 5 5

Offset 4 4 4 2 1 1

CY-graft 1 1 1 1 1 1 �

T-junction 3 2 2 2 2 3Offset 2 3 3 3 2 2

DY-graft 1 1 1 2 1 1 �

T-junction 3 3 3 1 1 2Offset 2 2 2 3 3 3

E

Y-graft I 3 3 3 3 3 3Y-graft II 2 2 2 2 2 2 �T-junction 4 4 4 4 4 5Offset I 1 1 1 5 5 4Offset II 5 5 5 1 1 1

split compared to other designs. These surgical recommendations are limited to

the patients we studied. Generalization of these choices should be made carefully,

as metrics used in this study could change as additional clinical data is obtained

in future work. The reader should be cautioned that not all variants of the Y-graft

design that we tested in this study performed well. However, the potential of the

Y-graft design to improve hemodynamic performance is promising, and a variant

of the Y-graft design was our final recommendation for all patients.

Page 126: Surgical design for the Fontan procedure using computational fluid

102

4.4 Limitations

A main limitation in this study is the lack of postoperative data on patients’

resistances and caval flow rates. Although this information would not typically be

available in a clinical pre-surgical design study, the development of sophisticated

models that can predict changes and remodeling in inflows and outlet boundary

conditions will be crucial for future surgical design and management of patients.

The evolution of this dynamic process is still an open question, and beyond the

scope of this study. Validation studies of pre-surgical design, long term hemody-

namics and uncertainty analysis should be incorporated into future work.

In this study we assumed an optimal hepatic flow distribution of 50/50,

and evaluated designs according to how closely they met this criterion. The tar-

get values chosen in this study could be adjusted on an individual basis as our

understanding of the relationship between hepatic flow concentration and lung de-

velopment is improved. Because the theoretical optimum was not achieved for all

patients in the study, it is possible that further design optimization may improve

some underperforming Y-grafts. Future work using patient specific optimization

would likely lead to further design refinement in some cases, as this would allow for

more systematic exploration of the design space. In addition, Bove et al.52 show

lateral tunnel (LT) Fontan models constructed from hemi-Fontan models result in

even hepatic flow distribution and lower power loss because of better mixing in the

right atrium, so future studies on the Y-graft should include comparisons with the

Page 127: Surgical design for the Fontan procedure using computational fluid

103

LT Fontan as well. Finally, inevitable discrepancies will occur between the virtual

design and its actual surgical implementation. This introduces certain geomet-

rical uncertainties that could result in differences between computer simulations

and the actual conditions. In addition, the use of rigid walls and the Newtonian

assumption for blood may affect the results presented in this study.

4.5 Conclusions

Using five Glenn patient specific models, we performed virtual Fontan surg-

eries and compared the hemodynamic performance of the Y-graft, T-junction and

offset configurations. We have demonstrated that the geometry considerably in-

fluences the hepatic flow distribution, and the hepatic flow split is not necessarily

equal to the pulmonary flow split. Theoretical analysis showed that a 50/50 hepatic

flow split is not attainable for some patients. Overall, the Y-graft design results

in more even hepatic flow distribution and moderate improvements in energy loss

and SVC pressure. The offset design is able to achieve an even hepatic flow dis-

tribution for patients with highly unequal pulmonary flow splits, but is sensitive

to variations in pulmonary flow split. It is important to note that, while a Y-graft

design was the best choice for all patients in the study, not all Y-graft configura-

tions performed well. The results of this study indicate that graft designs should

be optimized for individual patients prior to surgery. In conclusion, the Y-graft

is a promising new design that warrants testing in clinical application and long

Page 128: Surgical design for the Fontan procedure using computational fluid

104

term clinical trials. With further validation, simulations should be used to identify

the best candidates for the Y-graft procedure, and to rule out those patients who

should continue to receive conventional treatments.

4.6 Acknowledgments

This work was supported by the American Heart Association, a Burroughs

Wellcome Fund Career Award at the Scientific Interface, a Leducq Foundation

Network of Excellence grant and INRIA associated team grant. We are grateful

to Shawn Shadden for sharing his expertise and codes in particle tracking as well

as Sethuraman Sankaran, Frandics Chan, Heidi Terwey, and Mary Hunt Martin

for their helpful discussions and expertise. We also wish to acknowledge the use

of Simvascular (simtk.org), as well as the numerical modeling expertise of Nathan

Wilson and Charles Taylor.

Chapter 4, in full, is a reprint of the material as it appears in Yang, W.,

Vignon-Clementel, I. E., Troianowski, G., Reddy, V. M., Feinstein, J. A. and

Marsden, A. L. Hepatic blood flow distribution and performance in traditional

and Y-graft Fontan Geometries: A Case Series Computational Fluid Dynamics

Study. J. Thorac. Cardiovasc. Surg. 2012;143: 1086-1097.

Page 129: Surgical design for the Fontan procedure using computational fluid

Chapter 5

Y-graft optimal design for

improved hepatic flow

distribution

In Chapter 4, we demonstrated improved hemodynamic performance of the

Y-graft in multiple patient models. However, the use of a non-optimized design

resulted in underperforming Y-graft designs in two out of five patient-specific mod-

els.3 Automated shape optimization was applied to the Y-graft design using an

idealized model to reduce energy loss in a study of Yang et al.136 However, formal

shape optimization has not previously been applied to improve HFD. Preliminary

results have shown that the Fontan connection geometry is a critical determinant

of HFD.3 Thus, the impact of Fontan geometry and flow conditions on HFD, and

methods to improve underperforming designs merit further exploration.

105

Page 130: Surgical design for the Fontan procedure using computational fluid

106

In this chapter, we couple shape optimization to HFD quantification to

systematically improve Y-graft performance. The goals of this study are: 1) to

evaluate a new Y-graft design with unequal branch sizes, 2) understand the influ-

ence of flow splits on choice of optimal Y-graft, 3) examine the effect of SVC flaring,

and 4) improve HFD in previously underperforming Y-grafts. To achieve this, we

couple Lagrangian particle tracking to a derivative-free optimization framework

for cardiovascular geometries introduced by Marsden and colleagues76, 136 to opti-

mize HFD in Y-graft models. Because patients often present clinically with uneven

pulmonary flow splits, due to differing pulmonary resistances and lung sizes, we

identify optimal designs for a range of flow conditions. While a 55/45 (RPA/LPA)

ratio is commonly accepted,2, 141 Seliem et al.31 observed that 35% of patients

immediately prior to Fontan have a moderate to severe uneven pulmonary flow

distribution ranging from 28% to 2% flow to one lung. In agreement with this,

moderately uneven pulmonary flow distribution was found in two out of five Glenn

patients in our previous studies.3, 5 To achieve optimal HFD in these patients, we

hypothesize that unequal Y-graft branch sizes may be needed. Simple conservation

of mass analysis3 demonstrates that a 50/50 hepatic flow split is not possible in pa-

tients with a highly uneven overall pulmonary flow split. We therefore identify the

best theoretical distribution for each pulmonary flow split scenario, and set that as

our target for optimization. In addition, the theoretical analysis shows that HFD

also depends on the IVC/SVC flow ratio. Thus, the impacts of IVC/SVC flow

ratio on the optimal shapes are studied in idealized and patient-specific models.

Page 131: Surgical design for the Fontan procedure using computational fluid

107

Since the SVC-to-PA anastomosis may play an important role in flow interaction,3

we also investigate effects of the SVC anastomosis on the choice of optimal graft

design. Finally, two patient-specific models with underperforming Y-grafts are

optimized to improve skewed HFD.

5.1 Methods

This study is divided into two parts. In the first part, we optimize an ide-

alized Fontan model to identify optimal shapes for even HFD with variations in

the pulmonary flow split, IVC/SVC flow ratio and SVC-to-PA anastomosis geom-

etry. In the second part, we present two patient-specific examples to demonstrate

that optimization-derived designs effectively improve the HFD in cases that were

previously underperforming.

5.1.1 Geometrical model construction

Idealized cases

The construction of the idealized models follows our previous work (Chapter

3) in which an automated script is used to generate the model given a set of

geometric input parameters.136 Since we have hypothesized that unequal-sized

branches may be necessary to improve HFD, we also allow branch diameters to vary

independently in the model during optimization (Figure 5.1 upper right), resulting

in 7 design parameters for this case. Design parameters used to define the Y-graft

Page 132: Surgical design for the Fontan procedure using computational fluid

108

design space include the branch diameters DR and DL, the distances between the

branch anastomosis and the SVC XR and XL, the branch curvatures ΔR and ΔL,

and the trunk length LIV C . In addition, there are six constant parameters used to

define the dimensions of the IVC trunk, PA and SVC.136

In the Glenn procedure, the SVC is connected to the PA in an end to side

fashion. The anastomosis may be flared to direct flow preferentially to one lung

or the other or, in preparation for the subsequent Fontan procedure.5, 142 To study

the influence of the SVC geometry on the choice of optimal graft design, we replace

the regular SVC-PA junction in selected models with two representative cases: 1.

LPA-flared junction (Figure 5.1 bottom left) and 2. curved junction (Figure 5.1

bottom right).

Patient-specific cases

Since previous patient-specific studies3, 51 have shown large variations in PA

topology, the idealized model (Figure 5.1) constructed for general analysis cannot

sufficiently represent a specific patient’s geometric and hemodynamic characteris-

tics. We therefore optimized a semi-idealized model that incorporates key features

of the patient specific model, and then use a patient-specific model for final verifi-

cation.

For each patient, we create a semi-idealized model (Figure 5.2) as follows.

The patient-specific centerline path of the PA with circular segmentations and the

patient’s SVC are incorporated into the semi-idealized Glenn model to preserve

Page 133: Surgical design for the Fontan procedure using computational fluid

109

−10 −8 −6 −4 −2 0 2 4 6 8 100

2

4

6

8

10

12

DLPADRPA C2

XL

C4

C0

LIVC

XR C3

C1

RR RL

Figure 5.1: Model parameterization and flared SVC anastomosis. Upper left: De-sign parameters and centerlines of an idealized Y-graft Fontan model. Upper right:A representative Y-graft model. Parameters DL and DR allow two branches tovary independently. Bottom left: An LPA-flared SVC anastomosis with a straightjunction for the RPA side. Bottom right: A curved-to-LPA SVC anastomosis.

each patient’s main geometric characteristics (Figure 5.2). The PA diameter is

set to an average measured from the patient-specific Glenn model. Then, a Y-

graft is implanted into the Glenn model to form a semi-idealized Fontan model for

optimization.

The optimization for these cases only included designs with equal-sized

branches, which reduced the number of design parameters to four (Figure 5.2) :

XL, XR, LIV C and Dbranch. For the PA path, we can define a parametric spline

Page 134: Surgical design for the Fontan procedure using computational fluid

110

Table 5.1: Bounds on the design parameters for the semi-idealized model. XR

and XL are measured from the SVC-PA junction to the right and left anastomosispoints, respectively. Since the PA path is a parametric spline S(t), the anastomosislocation can be changed by varying the spline parameter t. In our previous study,3

patient specific models employed a uniform 20-15-15 mm Y-graft. To optimize thegraft size, the branch diameter was allowed to vary between 12 and 16 mm.

Patient Dbranch mm LIV C mm XR mm XL mm

Patient A 12-16 10-20 0.7-15 0.5-15.2Patient B 12-16 10-25 0.5-20.5 0.6-22

S(t) =

⎧⎪⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎪⎩

x = s1(t)

y = s2(t)

z = s3(t)

, (5.1)

such that the center of each PA segmentation lies on the curve. Thus, design pa-

rameters XL and XR are a function of parameter t. To determine the bounds of

XL and XR, the aortic arch and previous non-optimized patient specific models

constructed under the guidance of a surgeon were used as a reference. The param-

eter bounds are listed in Table 5.1. In this study, we use the terms “proximal” and

“distal” to describe the anastomosis location relative to the SVC . The anastomosis

points (marked with solid squares in Figure 5.2) are allowed to slide along the PA

path within the bounds. Similarly, the path of the Y-graft is defined by a Hermite

spline.

Based on surgical practice at our institution , the PA at the anastomosis

will be enlarged to match the graft size when a large graft is anastomosed to the

Page 135: Surgical design for the Fontan procedure using computational fluid

111

1 2 3

. .XLPA

Dbranch

LIVC

XRPA

Figure 5.2: 1. A patient-specific Glenn model. 2. In the semi-idealized Glennmodel, the PA is approximated by uniform circular segmentations and the pul-monary artery branches are neglected. The PA diameter is equal to the averageddiameter of the patient-specific PA. 3. A Y-graft is implanted forming a semi-idealized Fontan model for the same patient. The design parameters for the Y-graft are XL, XR, LIV C and Dbranch. When large branches are anastomosed, thesegmentation at the anastomosis is enlarged to the graft size. Then the rest ofthe PA segmentations are enlarged linearly according to the distance to the closestanastomosis.

PA and pressurized. For example, in measurements taken from images of a post-

operative Fontan patient who had a 20 mm extracardiac conduit placed, the size

of the PA at the anastomosis was about 19mm. To account for this change, the

circular segmentations along the PA path are adjusted automatically when the

branch size is larger than the PA size. First, the segmentation at the anastomosis

point is enlarged to the graft size. Then the rest of the PA segmentations are

enlarged linearly according to the distance to the closest anastomosis. The two

outlets of the PA are kept unchanged.

To improve initially underperforming Y-grafts, we first optimize the Y-graft

in the semi-idealized model. After an optimal Y-graft is identified, we implant it

into the patient-specific model to verify the hemodynamic performance.

Page 136: Surgical design for the Fontan procedure using computational fluid

112

Table 5.2: Mean pulsatile inflow rates, pulmonary flow splits and pressure. Arespiratory model4 was superimposed to the IVC flow acquired from PCMRI foreach patient following our previous work. No respiratory model was added tothe SVC input. The flow rates used for the idealized model were taken from atypical Fontan patient.4,5 We varied the RPA/LPA flow split in the idealized modelfor different conditions and set a Fontan pressure (central venous pressure) of 12mmHg. For patients A and B, pulmonary flow splits and pressure data were takenfrom MRI and catheterization prior to the Fontan procedure. Transpulmonarygradient (TPG) is the mean pressure difference between the SVC and the leftatrium.5

Model IVC (cc/s) SVC (cc/s) RPA/LPA flow split pressure (mmHg)

Idealized 17.9 14.1 varied 12Patient A 6.3 15.2 54/46 SVC:9, TPG:4Patient B 14.8 27.8 55/45 SVC:8, TPG:4

5.1.2 Flow simulation and boundary conditions

A custom stabilized finite element Navier-Stokes solver94 was employed to

simulate blood flow, assuming rigid walls and Newtonian flow with a density of

1.06g/cm3 and viscosity of 0.04 g/(cm s). Anisotropic mesh adaptation based on

the Hessian of the velocity field was performed for the patient-specific models, to

ensure mesh convergence of the solution.143

Pulsatile inflow conditions were employed for the IVC and SVC inlets with

a parabolic profile. A respiratory model was superimposed on the IVC flow wave-

form acquired from PCMRI for each patient following our previous work.4 No

respiratory model was added to the SVC input because little respiratory variation

is typically found in the SVC flow.4, 36 The mean flow rates for inlets are listed in

Table 5.2.

Resistance boundary conditions were employed for the idealized and semi-

Page 137: Surgical design for the Fontan procedure using computational fluid

113

idealized models. For a given pulmonary flow split, resistance values (R = P/Q)

were determined by setting 12 mmHg as the mean Fontan pressure, a typical clin-

ical value. For patient-specific simulations, a three element RCR circuit model96

was applied at each outlet. Total resistance was tuned to match the MRI-derived

pulmonary flow split and catheterization-derived transpulmonary gradient (TPG)

(Table 5.2), assuming that the pre-Fontan outlet boundary conditions are still valid

for immediate post-operative flow conditions.3, 5 The proximal and distal resis-

tances and capacitance for each outlet were determined based on a morphometry-

based pulmonary arterial tree and outlet areas.5 Resistances for each patient’s

semi-idealized model matched the total LPA and RPA resistances of the corre-

sponding patient-specific model.

Quantification of HFD

We assume that hepatic flow is well mixed in the IVC such that the hepatic

and IVC flow distributions are the same, and that the theoretical optimal HFD

is the closest value to 50/50. Based on Equation 4.2, the theoretical optimum for

the HFD can be obtained for any combination of the pulmonary flow split and

IVC-to-SVC flow ratio, as shown in Figure 5.3. The fraction of SVC flow going to

the RPA, SRPA, is allowed to change between 0 and 1 to achieve the best HFD.

A few conclusions can be drawn from Equation 4.2 and Figure 5.3: 1) A perfect

50/50 split is not always possible for some combinations of the pulmonary flow split

and IVC/SVC flow ratio without violating conservation of mass. For example, if

Page 138: Surgical design for the Fontan procedure using computational fluid

114

a patient has an overall flow split to LPA and RPA of 19/81, and an IVC/SVC

flow split ratio of 0.93 then the best hepatic distribution one can achieve is 0.63,

in the case that all the SVC flow goes to the RPA. 2) FRPA = 0.5 is the only value

for which the theoretical optimum is always 50/50, regardless of QIV C

QSVC. 3) For an

IVC/SVC ratio of 1, the theoretical optimum is 50/50 for a percentage inflow to

the RPA between 25% and 75%.

00.1

0.20.3

0.40.5

0.60.7

0.80.9

1

0

0.2

0.4

0.6

0.8

10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

QIVC

/(QIVC

+QSVC

)% inflow to RPA

best

% h

epat

ic fl

ow to

RPA

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

Figure 5.3: Optimal values for the HFD. Based on Equation (4.2), the theoreticaloptimum for the HFD, defined as the value closest to 50/50, is determined givenan inflow ratio QIV C

QSVCand a pulmonary flow split FRPA (% inflow to RPA).

To quantify the HFD, we used the same particle tracking introduced in

Chapter 4.

5.1.3 Optimization algorithm

The surrogate management framework (SMF)98 together with mesh adap-

tive direct search (MADS)86 is employed for optimizing HFD, following the work

of Dennis, Audet and Marsden.76, 102, 136 We consider the optimization problem,

Page 139: Surgical design for the Fontan procedure using computational fluid

115

minimize J(x),

subject to x ∈ Ω, (5.2)

where J : Rn → R is the cost function, Ω ⊂ Rn denotes the feasible region and x

is the vector of parameters.

The cost function J in this study is defined as

J = |HRPA − 0.5| , (5.3)

where HRPA is the fraction of IVC flow going to the RPA. For a given x, J is

obtained by performing a 3D simulation and particle tracking. We optimize the

HFD to achieve a target flow split of 50/50. For cases in which the theoretical

optimum is not 0.5, the best cost function value is therefore larger than zero.

5.2 Results

5.2.1 Idealized cases

We first examine the question of whether unequal sized branches may be

advantageous to improve HFD for cases of highly uneven pulmonary flow split.

To mimic an uneven pulmonary flow split, the LPA/RPA resistance ratio was set

Page 140: Surgical design for the Fontan procedure using computational fluid

116

to 4. In the first case, we allowed branch diameters to vary independently during

optimization to determine if unequal branches are needed to achieve optimal per-

formance. In subsequent tests, we restricted the number of design variables to use

equal-sized branches only and kept boundary conditions unchanged to determine

whether unequal-sized branches are the only way to achieve optimal HFD.

Figure 5.4 shows that shape optimization for a case with a highly uneven

pulmonary flow split 79/21 (RPA/LPA) caused a large difference in the optimal

branch diameters. The branch size for the RPA reduced to the lower bound,

increasing resistance for IVC flow streaming to the RPA. In contrast, optimization

results for the case of equal-sized branches achieved almost the same optimal HFD,

with a 9% smaller energy loss. In addition, outflow rates in both cases were almost

unchanged, confirming that the overall pulmonary flow split is largely determined

by the outlet boundary conditions, and not the local geometry.

These initial results suggested that unequal-sized branch diameters were

unnecessary for achieving optimal HFD. To further confirm this hypothesis, we

optimized the idealized model with equal-sized branches over a large range of pul-

monary flow splits. Figure 5.5 shows that the theoretical optima for the pulmonary

flow splits we tested are achieved, and the optimal geometry depends on the pul-

monary flow split. We observe that anastomosis locations were more distal on the

side of higher pulmonary resistance.

Shape optimization of the Y-graft was also performed for two different SVC

flaring configurations. Figure 5.6 shows the optimal Y-grafts for these two flared

Page 141: Surgical design for the Fontan procedure using computational fluid

117

1.00 1.10 1.20 1.30 1.40 1.50 1.60 1.70 1.80 1.90 2.00

0% 20% 40% 60% 80% 100%

unequal branchesequal branches

% IVC flow to RPA

Pow

er lo

ss (m

W)

Figure 5.4: A comparison of HFD and energy loss for optimal unequal and equal-sized branches. HFDs for the unequal and equal-sized branches are 63/37 and65/35 (IVC-RPA/IVC-LPA), respectively, but equal-sized branches perform betterin reducing energy loss. The pulmonary flow split is 79/21 (RPA/LPA).

0%10%20%30%40%50%60%70%80%90%

100%

0% 20% 40% 60% 80% 100%

resultsTheor. Opt.

% inflow to RPA

% IV

C flo

w to

RPA

1.6%

0%1%

0%

Figure 5.5: Optimal Y-grafts with equal-sized branches for a large range of pul-monary flow splits. Theoretical optima given by Equation 4.2 are achieved by usingoptimization. The difference from the theoretical value is shown at each point.

Page 142: Surgical design for the Fontan procedure using computational fluid

118

SVC anastomoses and a straight anastomosis under the same flow split conditions.

While the optimal geometry for the straight SVC connection is almost symmetric,

the optimal geometry for both flared configurations is asymmetric with the RPA

connection more proximal, and the LPA connection more distal. These differences

in geometry result from SVC streaming to the LPA side in case of a flared SVC

anastomosis.

straight LPA flared curved

Figure 5.6: Time-averaged flow fields of optimal Y-grafts for a straight SVC-PAjunction and two types of flared SVC anastomoses. The pulmonary flow split is55/45 (RPA/LPA). Compared to the model with a straight SVC-PA junction, theoptimal Y-grafts for two flared SVC anastomoses have a more distal anastomosisfor the LPA.

Since the HFD is also a function of the IVC/SVC flow ratio, we altered

the ratio without changing the cardiac output to examine the impact on different

models. Figure 5.7a shows that the idealized Y-graft model is insensitive to changes

in this ratio. However, the changes in the IVC/SVC flow ratio can significantly

influence the HFD in a patient-specific model due to a more complex flow field.

Page 143: Surgical design for the Fontan procedure using computational fluid

119

0%10%20%30%40%50%60%70%80%90%

100%

0% 20% 40% 60% 80% 100%

idealizedpa�ent-specific

QIVC/Qinflow

% IV

C flo

w to

RPA

Figure 5.7: HFD vs. QIV C

Qinflowfor an idealized model and a patient-specific model

(patient B). Patient B’s original inflow ratio QIV C

Qinflowis marked by an arrow. Total

inflow is kept constant in this comparison. The idealized Y-graft is optimized foran IVC inflow-to-total inflow ratio of 45%. There is only 1% change in the Y-graft model when the ratio is altered. However, the patient specific model is moresensitive to the change of IVC inflow-to-total inflow ratio.

5.2.2 Patient-specific cases

To test the ability of optimization to improve hepatic flow in a patient

specific model, we now present the results for patient specific optimization using

a semi-idealized model. Figure 5.8 shows that the optimal Y-graft identified via

a semi-idealized model for patient A resulted in a consistent HFD after it was

implanted into the patient-specific Glenn model. Compared to the original non-

optimized design, the HFD is improved by 79% and the left branch anastomosis in

the optimal design is more proximal. With a more proximal anastomosis of the Y-

graft to the RPA, the SVC flow suppressed hepatic flow, reducing its concentration

Page 144: Surgical design for the Fontan procedure using computational fluid

120

in the RPA to a balanced level. An additional optimization case for patient A was

run in which a uniform PA diameter was used. When the resulting optimal Y-graft

was implanted into the patient specific model, there was no improvement in HFD.

In patient B, the right upper lobe (RUL) is adjacent to the SVC. In our

first optimization run, the RUL was not included in the semi-idealized model for

optimization and the target HFD was set to 50/50 as before. Figure 5.9 shows that

the target optimal Y-graft made no improvement in the patient-specific model even

though it achieved the target value in the semi-idealized model. By examining each

outflow rate, we found that the flow to the RUL accounted for 15% of the RPA

outflow. Since the RUL mainly received flow from the SVC, the concentration of

SVC flow in the RPA flow was sensitive to the presence of the RUL. To compensate

for the presence of the RUL in the semi-idealized model, the target hepatic flow

split value was changed to 0.65 in the second optimization run. Figure 5.10 shows

that a Y-graft that achieved a hepatic flow split of 65/35 in the idealized model

successfully distributed hepatic flow evenly when it was implanted to the patient-

specific model. In the resulting model, the right anastomosis is more distal, while

the left one is more proximal in the optimized model and the branches have less

curvature in the optimized design compared to the previous one. Table 5.3 lists

geometric parameters and power loss for patient specific models. A smaller branch

size resulted in more power loss in both patients.

Page 145: Surgical design for the Fontan procedure using computational fluid

121

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

Model 1 Model 2 Model 3

IVC-RPAIVC-LPA

% IV

C flo

w

Model 1: optimal Y-graftwith semi-idealized model

Model 2: optimal Y-graftwith patient-specific model

Model 3: non-optimized Model 2

RPALPA

a)

b)

Figure 5.8: a) Time-averaged velocity vector fields in the semi-idealized andpatient-specific models for patient A. b) Particle snapshots taken at T=3s for thenon-optimized and optimal models. The bar chart shows the semi-idealized model(upper left) has a similar hepatic flow split to the patient-specific model (upperright) for the same optimal Y-graft, and that the optimized Y-graft improves theHFD by 79%, compared to the original non-optimized design (lower left). Theoptimal and non-optimized branch sizes are 12.9 and 15 mm, respectively.

Page 146: Surgical design for the Fontan procedure using computational fluid

122

Model 1: optimized for 50/50 Model 2: patient specific

RUL

0%10%20%30%40%50%60%70%80%90%

100%

Model 1 Model 2 Model 2 w/o RUL

IVC-RPAIVC-LPA

% IV

C flo

w

Figure 5.9: Time-averaged velocity vector fields in the semi-idealized and patient-specific models with and without the RUL for patient B. The Y-graft is optimizedfor a HFD of 50/50. Due to the effect of the RUL, the optimized Y-graft skewedthe hepatic flow by around 15% after it was implanted into the patient-specificmodel. When the RUL is excluded from the patient-specific model, the HFD isconsistent with the idealized model prediction.

Table 5.3: Geometric parameters and power loss for patient specific models. Inpatient A, the optimal XR was reduced resulting in a more proximal anastomosisfor the right branch. In patient B, the right anastomosis is more distal while theleft one is more proximal in the optimized model. In both cases, a smaller branchsize resulted in more power loss.

Patient Dbranch (mm) LIV C (mm) XR (mm) XL (mm) Power loss(mW)

A optimized 12.9 15.25 7.3 15.2 1.9A non-optimized 15.0 18.5 15.0 19.0 1.8B optimized 16.0 20.8 20.2 21.9 5.3

B non-optimized 15.0 13.5 17.0 23.0 5.8

Page 147: Surgical design for the Fontan procedure using computational fluid

123

0%

10%

20%

30%

40%

50%

60%

70%

80%

90%

100%

Model 1 Model 2 Model 3

IVC-RPAIVC-LPA

% IV

C flo

w

Model 1: optimized for 65/35 Model 2: optimal Y-graftwith patient-specific model

Model 3: non-optimized Model 2

RPA LPA

a)

b)

Figure 5.10: a) Time-averaged velocity vector fields and HFD for patient B. TheY-graft in the semi-idealized model (upper left) is optimized for a hepatic flowsplit of 65/35 (RPA/LPA) to account for the overestimation of the RPA hepaticflow in the semi-idealized model. b) Particle snapshots taken at T=3s for the non-optimized and optimal models. The bar chart shows the optimal Y-graft improvesthe performance by 94% achieving an even HFD in the patient-specific model(upper right), compared to the non-optimized design (lower left). The optimaland non-optimized branch sizes are 16 and 15 mm, respectively.

Page 148: Surgical design for the Fontan procedure using computational fluid

124

5.3 Discussion

By coupling particle tracking to an optimization framework, we optimized a

series of Y-graft designs to achieve optimal HFD in different scenarios. In idealized

cases, the unequal branch design was optimized and compared to the equal branch

design. Results showed that unequal branches are not necessary to achieve an even

HFD in all cases that we evaluated. We studied the impacts of the pulmonary

flow split, IVC/SVC flow ratio and SVC flaring on choice of optimal Y-graft.

Again, the Fontan connection geometry played an important role in distributing

the hepatic flow. Two previously underperforming patient-specific Y-graft designs

were improved by optimizing semi-idealized models with the patients’ SVC and

PA paths and then implanting the optimal design into the corresponding patient

specific model.

5.3.1 Idealized cases

We have previously observed that Y-graft designs with equal-sized branches

resulted in skewed hepatic flow in some patient-specific models. To have greater

control over hepatic distribution, we hypothesized that unequal branches might

improve distribution in select cases, particularly for patients with highly uneven

right/left pulmonary resistances. In this study, we tested this hypothesis using

an idealized model with a right/left pulmonary flow split close to 80/20 and com-

pared to results for equal-sized branches. Optimization with and without unequal

Page 149: Surgical design for the Fontan procedure using computational fluid

125

branches were both shown to achieve the target optimal hepatic distribution. How-

ever, designs with unequal branches dissipated more energy due to a higher pressure

drop in the smaller branch. Although reducing the branch size on one side can

increase the resistance to that side, our results showed that changes in the branch

geometry had little-to-no influence on the overall pulmonary flow split. More im-

portantly, the use of unequal-sized branches had limited impact on the HFD, while

the anastomosis locations had a significant impact. Thus, for models with a normal

pulmonary flow split around 45/55, equal-sized branches can easily achieve the tar-

get optimal flow distribution through a proper choice of anastomosis locations and

branch curvature. In addition, graft designs with unequal branch diameters may

increase difficulties in manufacturing compared to equal-sized branches. Thus, the

equal-sized branches should be suitable for most cases regardless of the pulmonary

flow split.

Since patients often come in with a range of relative blood flow distribution,

our intent was to explore this range of starting conditions for Y-graft implantation.

For patients with widely unequal pulmonary flow distribution, we identified a series

of optimal Y-graft designs in which the theoretical optimal HFD is achieved over

a large range of pulmonary flow splits. Results show that the optimal anastomosis

location depends on the pulmonary resistances for the corresponding side. For a

fixed IVC/SVC ratio, if the inflow received by the RPA is reduced, one needs to

reduce the SVC flow that goes to the RPA to keep an even HFD (Equation 4.2).

This can be achieved by modifying the Y-graft such that its implantation facilitates

Page 150: Surgical design for the Fontan procedure using computational fluid

126

flow from the IVC to the RPA. This thus leads to a more distal anastomosis

of the graft to the RPA, which streams IVC flow towards the RPA, the branch

with higher resistance. Therefore, the PA with lower resistance requires a more

proximal anastomosis, while the PA with higher resistance requires a more distal

anastomosis. In addition, HFD is more sensitive to the anastomosis location than

the branch size.

During the Glenn procedure, some surgeons flare the SVC towards the

LPA intending to channel more SVC flow to the LPA which usually has a higher

resistance than the RPA.5, 142 However the pulmonary flow split is dictated by

the downstream resistance, and thus the distribution of SVC flow is equal to the

pulmonary flow split in the Glenn. After Fontan completion, SVC geometry may

play an important role in determining HFD. We performed Y-graft optimization

for two types of flared SVC anastomoses. For both flared geometries, the resulting

optimal designs had a more distal branch anastomosis on the LPA side compared

to the straight SVC case. This can be explained by the fact that flaring to the LPA

increases SVC flow to the LPA. Hence, according to Equation 4.2, to maintain the

optimum HFD for given flow split and IVC/SVC flow ratio, SVC flow to the RPA

needs to be facilitated. This can be achieved either by impeding IVC flow to the

RPA (leading to a more proximal anastomosis on that side) or by facilitating IVC

flow to the LPA (leading to a more distal anastomosis on that side). The latter

case lead here to the optimum solution. Compared to the flared SVC case, the

curved SVC anastomosis channeled more SVC flow to the LPA for most Y-grafts

Page 151: Surgical design for the Fontan procedure using computational fluid

127

evaluated during the optimization and required about 200% more evaluations to

achieve the theoretical optimum. The relative pulmonary blood flow in the Glenn

patient plays a signficant role in subsequent Fontan blood flow distribution and,

as such, consideration should be given to measuring this routinely in pre-operative

Fontan patients. In addition, surgeons should carefully consider the specifics of the

anastomosis for the SVC-PA junction in the Glenn procedure with an eye towards

the impact on the HFD in the Fontan procedure. For example, if a patient has an

RPA/LPA flow split of 80/20, it would be preferential to have most of the SVC

flow going to the RPA when the IVC is connected to the PA in order to optimize

hepatic flow split. If the Glenn, however, had the SVC-PA anastomosis flared

towards the LPA, a Y-graft that achieves the target HFD would be difficult.

We also studied the effect of changing the IVC/SVC flow ratio on HFD

in both idealized and patient-specific models. In the idealized case, for a Y-graft

optimized for a certain IVC/SVC flow ratio, the HFD remains almost unchanged

when the IVC/SVC ratio is altered. However, a patient specific model is more

sensitive to changes in IVC/SVC flow. A possible reason is that the Y-graft in

this patient-specific model has proximal anastomoses. The trend is qualitatively

consistent with the change from rest to exercise which was observed in our previous

study,3 in which increasing the IVC flow momentum improved the HFD. Thus a

significant change in IVC flow momentum directly influences the SVC-IVC flow

interaction. These findings indicate that the idealized model is not sufficient for

capturing this sensitivity.

Page 152: Surgical design for the Fontan procedure using computational fluid

128

5.3.2 Patient-specific cases

In patient A, we demonstrated that the patient’s PA path and size should

be included in the semi-idealized model to achieve consistent predictions when

the graft design is implanted in the patient specific model. Since optimization

involves systematically evaluating a series of designs, the use of idealized models

can significantly reduce overall computational cost. Simulations for a patient spe-

cific model are 15 times more expensive than for a semi-idealized model. However,

without proper consideration, an oversimplified model may be lacking important

local information and result in a failed prediction. Direct patient-specific shape

optimization requires a methodology solution to parameterize and automatically

manipulate the graft anastomosis. At the present time, it remains a challenge to

parameterize patient-specific models, due to the need for manipulating a complex

surface while maintaining its integrity using relatively few design variables. In the

future, a fully parameterized patient specific model could serve as a high fidelity

model and a multi-fidelity optimization could be performed with a low-fidelity

semi-idealized model.144

In patient B, the importance of including the RUL when computing the

HFD was revealed. In the model for patient A, the RUL flow was lumped together

with the RPA when constructing the semi-idealized model. However, the same

approach did not prove adequate for patient B because the RUL attachment point

was on the SVC. Results showed flow streaming from the SVC to the RUL with very

Page 153: Surgical design for the Fontan procedure using computational fluid

129

little mixing and less than 1% of hepatic flow delivered to the RUL. Therefore the

presence of the RUL directly influenced the concentration of SVC flow in the RPA

outflow. In contrast, including or excluding the RUL will not produce significant

differences in the HFD if the RUL is distal to the SVC. Thus, for patients with

a direct RUL-SVC connection, the semi-idealized model must compensate for the

RUL flow contribution, as done in patient B. The extra hepatic flow going to

the RPA in the semi-idealized model would be offset when the same Y-graft was

implanted into the patient-specific model with the RUL.

In patient A, without optimization, HFD is too skewed towards the RPA.

According to the semi-idealized simulations, the Y graft can achieve better HFD

either with a more proximal anastomosis on the RPA side (to reduce IVC flow to the

RPA) or a more distal anastomosis on the LPA side (to favor IVC flow towards the

LPA). Indeed, the optimized geometry lead to a more proximal anastomosis on the

RPA side. On the other side, due to geometric constraints, the anastomosis could

not be placed more distally. To compensate, the diameter was decreased to favor

hepatic flow to the LPA. By contrast for patient B, optimization needed to increase

IVC flow to the RPA. Here, the optimum was found by using a more proximal

anastomosis on the LPA side (impeding IVC flow to that side) and increasing

the graft diameter (facilitating flow to the RPA). The complex interplay of the 3D

geometry and flow is highlighted in these patient specific cases, where identification

of optimum configurations requires computer simulations using the full model.

No consistent choice for optimal branch size was observed. Patient A’s

Page 154: Surgical design for the Fontan procedure using computational fluid

130

optimal branch size (12.9 mm) was smaller than the non-optimized size (15 mm)

while a larger diameter (16 mm) was identified for patient B. This can be attributed

to the existence of multiple local minima. A larger branch size can slightly reduce

power loss, however, this was shown in a recent multi-scale modeling study to have

negligible effect on cardiac work load.54 The trunk length had a variable influence

on HFD with no particular trend for the values considered. The anastomosis

locations for the Y-graft play a more important role than the trunk length in

regulating HFD. However, we believe that the branch size and trunk length may

have a pronounced impact on wall shear stress (WSS) levels in the graft. Future

work should consider thrombotic risk in the surgical design, and thus a further

study with WSS and residence time constraints is warranted.

5.3.3 Technical considerations for Y-graft implantation

Space constraints are the main concern for surgical implantation of the Y-

graft. Since 20 mm grafts with an offset are routinely implanted in our institution,

our surgeons maintain there should be enough space for a Y-graft with 12 or 14

mm branches. The space available, however, can vary dramatically from patient

to patient, making some patients better candidates than others. The Y-graft may

also not be applicable to patients with abnormal anatomy and/or a history of

clotting disorder. Special care should be taken for these patients to identify an

optimal graft prior to surgery. Technical feasibility must be confirmed by a clinical

study.

Page 155: Surgical design for the Fontan procedure using computational fluid

131

5.4 Limitations

A HFD of 50/50 is assumed to be optimal in this study, and all Y-graft de-

signs are evaluated according to this criterion. Although the exact value required

for the prevention of PAVMs is still unknown, the goal of distributing hepatic flow

evenly is supported by clinical practice. As a step towards a better understand-

ing of the relationship between Fontan geometry and HFD, we chose 50/50 as a

reasonable target value which could be adjusted in future studies.

In this study, a single objective optimization with a simple bound con-

straint was performed. However it is clear that surgical design is multi-factorial

and physiological states can change over time. In addition, discrepancies from

surgical implementation may result in a large deviation in post-operative perfor-

mance. Multiple objectives, constraints and robustness should be addressed in

future studies.135

For the idealized model, the Y-graft is restricted to lie in a plane. Additional

variables for future study could include out-of-plane anastomosis angles and trunk

size to increase design flexibility though as mentioned previously, space within the

chest cavity is limited. The geometry of the SVC-PA junction should be exam-

ined in future studies to determine if the Glenn procedure should be performed

differently in patients with a planned Y-graft procedure.

Our assumption that the PA is enlarged to accommodate a large graft

anastomosis is based on common surgical practice. The work of Dobrin et al.145

Page 156: Surgical design for the Fontan procedure using computational fluid

132

showed that the maximum size of the anastomosis is determined by the Gore-tex

graft because the arteries are much more compliant than the Gore-tex material.

Although we do not model the process of vessel enlargement in this study, models

are constructed under the guidance of a surgeon to replicate realistic anastomoses.

Future work will focus on incorporating finite element modeling into the model

construction such that the process of anastomosing the graft in patient-specific

models can be implemented automatically.

The HFD is quantified numerically. Although the flow solver that provides

velocity fields for particle tracking has been validated against experiments and

theoretical solutions, predictions of patients’ HFD still need to be validated in vivo

and in vitro. Patients’ lung perfusion data will be used to compare with numerical

simulations in our future work, and comparisons to in vitro models should also be

made.

In this study, the HFD was divided into two parts (IVC-RPA and IVC-

LPA). However, we found that little hepatic flow was delivered to the RUL in

patient B though the overall hepatic flow split is close to 50/50. It is still unknown

whether an overall even hepatic flow split with a localized lack of hepatic flow in

a lobe can cause PAVMs. Future work would look at the incidence of PAVMs in

the RUL and consider controlling the HFD for each lobe.

A rigid wall assumption is used in this study. Our recent work has demon-

strated that HFD is insensitive to the use of rigid vs. deformable walls, which

increased confidence in our HFD predictions.57 However future predictions of

Page 157: Surgical design for the Fontan procedure using computational fluid

133

thrombotic risk linked to wall shear stress will likely require fluid structure in-

teraction, as there are large differences in WSS between rigid and flexible wall

simulations.

For the patient specific cases, flow measurements were taken from preop-

erative PCMRI data. Although we accounted for changes in IVC inflow patterns

after the Fontan procedure by scaling the IVC waveform, a multi-scale closed loop

model54 should be incorporated in future studies.

5.5 Conclusion

In this study, we coupled Lagrangian particle tracking to an optimization

framework to investigate the effect of geometry on HFD in a Y-graft Fontan con-

figuration. Y-graft models with unequal-sized branches were compared to models

with equal-sized branches. Optimized models with equal-sized branches are able

to distribute hepatic flow equally well as unequal-sized branches with lower en-

ergy loss under highly uneven pulmonary flow split conditions. In addition, the

theoretical optima are achieved using equal-sized branches over a large range of

pulmonary flow splits. Thus we do not recommend unequal-sized branches for fu-

ture Y-graft designs. A flared SVC anastomosis impacts optimal geometry of the

Y-graft by resulting in a more distal anastomosis for the branch on the flared side

and a more proximal one on the other side compared to the non-flared case. In

idealized models, the Y-graft design is more robust to changes in the IVC/SVC

Page 158: Surgical design for the Fontan procedure using computational fluid

134

flow ratio than the offset design. However, a patient-specific test did not support

this finding.

Two underperforming Y-grafts have been successfully optimized for patient-

specific cases by using semi-idealized Glenn models that incorporated key patient-

specific attributes, namely a patient-specific SVC and a curved PA path. Com-

pared to the original designs, these optimized Y-grafts for patients A and B im-

proved HFD by 79% and 94%, respectively. The strategy of using semi-idealized

models for optimization avoids a costly trial-and-error design process, requiring

laborious manual model revisions. We also found that ignoring the effect the right

upper lobe when it is adjacent to the SVC may result in failure to improve HFD in

patient-specific models. This study emphasizes that an optimization plan should

be tailored for each patient within the context of the overall framework we have

presented.

5.6 Acknowledgments

This work was supported by the American Heart Association, a Burroughs

Wellcome Fund Career Award at the Scientific Interface, a Leducq Foundation

Network of Excellence grant and an INRIA associated team project grant. The

authors wish to thank Charles Audet, Sebastien Le Digabel and Mohan V. Reddy

for sharing their expertise on optimization and pediatric cardiac surgery , and

Frandics Chan for image-data acquisition and expertise.

Page 159: Surgical design for the Fontan procedure using computational fluid

135

Chapter 5, in full, is a reprint of the material as it appears in Yang, W.,

Feinstein, J. A., Shadden, S. C., Vignon-Clementel, I. E. and Marsden, A. L.

Optimization of a Y-graft Design for Improved Hepatic Flow Distribution in the

Fontan Circulation. J. Biomech. Engrg., accepted.

Page 160: Surgical design for the Fontan procedure using computational fluid

Chapter 6

Simulations and validation for the

first cohort of Y-graft Fontan

patients

Simulations on idealized and patient specific models have shown that overall

the Y-graft design improves energy loss, SVC pressure and hepatic flow distribu-

tion (HFD) though no one-size-fits-all Y-graft design exists.2, 3, 136 Since Optiflo

introduces extra synthetic materials and technical difficulties, the dual bifurcated

design has not been studied further. Based on the previous simulation results, the

Y-graft design has been translated into clinical use in two institutions.146 In a pi-

lot study at Lucile Packard Children’s hospital at Stanford University, six patients

underwent a Y-graft EC Fontan surgery between June 2010 and March 2011. The

technical success demonstrated the feasibility of the Y-graft, which was a major

136

Page 161: Surgical design for the Fontan procedure using computational fluid

137

concern due to limited space in the chest cavity.

Previous hemodynamic evaluations of the Y-graft were based on virtual

surgeries.2,3 Although CFD for blood flow modeling has been validated against

theoretical solutions, in-vitro and in-vivo experiments,94, 117, 147 there is a lack of

in-vivo validation for Fontan patients in the literature. The study in this chapter

has two main goals: 1) to evaluate post-operative hemodynamic performance in

the first cohort of Y-graft Fontan patients, and 2) validate simulation predictions

of hepatic flow distribution against in-vivo clinical data. Based on the importance

of HFD, simulation-derived HFD and patients’ lung perfusion data were chosen to

validate the credibility of flow simulations for surgical design.

6.1 Methods

6.1.1 Surgical technique and clinical data acquisition

The Y-graft implanted in six patients (YF1 to YF6) was custom made for

each patient by the surgeon with an 18 mm trunk and 12 mm branches. Patients

received standard post-operative care, with no change in international normal-

ized ratio (INR) target. Although the six surgeries were technically successfully,

thrombus was found in the left branch in patient YF5 3 months after the Fontan

procedure.

Catheterization was routinely performed immediately prior to the Fontan

procedure. PAVMs were evident in the RPA of patient YF5. Early diffuse PAVMs

Page 162: Surgical design for the Fontan procedure using computational fluid

138

were found in the right upper lobe of patient YF4.

Early (< 1 month) and six-month postoperative magnetic resonance imag-

ing (MRI) scans were performed on patients YF1, YF2 and YF3. Pulmonary and

vena cava flow was measured by phase contrast MRI (PC-MRI). For patients YF4

and YF6, only CT images were collected. To quantify the HFD in vivo, a lung

perfusion scan was performed on these three patients in the early post-operative

stage. Authors who performed flow simulations were blinded to lung perfusion

data until all corresponding simulation results were reported. Patient YF5 under-

went pre-Fontan and 3 month post-operative MRI scans. MRI and lung perfusion

scans were performed with an institutional review board approved protocol.

6.1.2 Model construction

Following our previous work, patient specific models were constructed us-

ing a custom version of the open source Simvascular software package.2 From the

acquired image data, centerline paths were created in the Y-graft, SVC and PAs,

and segmentations of the vessel lumen were created in all vessels. Finally, a 3D

solid model was created by lofting all segmentations. Since patient YF5 developed

thrombus in the left branch of the Y-graft, an additional virtual unblocked Y-graft

was constructed for comparison. Patient specific models and MRI/CT images

are shown in Figure 6.1. To examine the influence of graft size and anastomosis

location on wall shear stress (WSS) and HFD, two modified Y-graft designs con-

structed for patient YF5. First, the right and left branch sizes were changed to 14

Page 163: Surgical design for the Fontan procedure using computational fluid

139

YF1

YF4

YF3YF2

YF5 YF6

Figure 6.1: Post-operative MRI/CT images and models. Since patients YF5 de-veloped thrombus in the left branch, an unblocked Y-graft was reconstructed forstudy.

mm and 10 mm, respectively (referred to as R14-L10). Then, model R12-L12 was

created to have a distal anastomosis and a proximal anastomosis for the right and

left branches respectively without changing the branch size.

Page 164: Surgical design for the Fontan procedure using computational fluid

140

6.1.3 Flow simulation and boundary conditions

To simulate blood flow, a 3D finite element Navier-Stokes solver was em-

ployed with a Newtonian approximation for the viscosity and a rigid wall assump-

tion.94 Pulsatile flow boundary conditions were applied to the IVC and SVC inlets.

To account for the respiratory effects in IVC flow, MRI-derived pulsatile inflow

data were superimposed with a respiratory model following our previous work.4

RCR boundary conditions were tuned to achieve the target Fontan pressure and

pulmonary flow split.3,5 Post-operative catheterization derived Fontan pressure of

11 mmHg was used as the target for patient 5, and 12 mmHg was assumed for

all other patients since no catheterization data was available. For patients YF4

and YF6, three pulmonary flow splits, 35/65, 55/45 and 75/25, were applied since

these patients were imaged with CT. The IVC and SVC inflow conditions for these

patients were taken from patient YF5 and scaled according to BSA. In the fol-

lowing paragraphs, we use the format, RPA/LPA, to present pulmonary flow split

and HFD. Table 2 lists all patients’ flow data and pulmonary flow splits used in

simulations.

To investigate the potential factors that led to the formation of thrombus in

patient YF5, three scenarios were studied. We first simulated the flow environment

at the time of the 3 month post-operative MRI, after formation of the thrombus.

A Y-graft model with a blocked left branch was constructed directly from the

post-operative MRI data and boundary conditions from the PCMRI acquired dur-

Page 165: Surgical design for the Fontan procedure using computational fluid

141

Table 6.1: Patients’ flow conditions used in simulations. The vena cava flow andpulmonary flow split were measured by PC-MRI except for patients YF4 and YF6,who had CT imaging. We use the format, RPA/LPA, to present pulmonary flowsplit. For patients YF1, YF2 and YF3, “early” and “6 month” denote measure-ments taken in the early (< 1 month) and 6 month post-operative stages, respec-tively. For patient YF5, pre-operative and 3 month post-operative measurementswere performed.

Data acquisition Patient IVC ml/s SVC ml/s Pulmonary flow split

Measured

YF1early: 6 early: 14 early: 81/19

6 month: 10.7 6 month: 17.8 6 month: 70/30

YF2early: 11.8 early: 15 early: 60/406 month: 8.2 6 month: 16.8 6 month: 66/34

YF3early: 9.3 early: 9.5 early: 59/41

6 month: 21.5 6 month: 15.3 6 month: 53/47

YF5pre-Fontan: 11.5 pre-Fontan: 14 pre-Fontan: 81/193 month: 11.7 3 month: 12.2 3 month: 54/46

ScaledYF4 11.4 11.9 35/65, 55/45, 75/25YF6 11.2 11.7 35/65, 55/45, 75/25

ing this scan were used. Second, a complete Y-graft was virtually reconstructed

by removing the branch blockage, and the same post-operative boundary condi-

tions were used. Third, we simulated the presumed flow conditions in the early

post-operative state using an unblocked Y-graft model with the pre-operative pul-

monary flow split, assuming that pulmonary remodeling progresses gradually and

the pulmonary flow split does not change significantly within a short period (less

than one month) after Fontan completion.

Lagrangian particle tracking was used to quantify the HFD.2, 139 Approxi-

mately 10,000 particles were released at the IVC inlet every 1/100 cycle for a cycle.

HFD was derived by computing the particle flux for particles traveling to the RPA

and LPA, respectively.3 In addition, previous work has shown that simulation

Page 166: Surgical design for the Fontan procedure using computational fluid

142

predictions of HFD are insensitive to the use of rigid vs. deformable walls.58

6.2 Results

Results are divided into three parts. We first present validation and lon-

gitudinal results for patients YF1, YF2 and YF3. Second, HFD under a range

of pulmonary flow splits for patients YF4 and YF6 is shown. Third, the issue of

thrombosis in patient YF5 is explored in detail to identify possible causes. For

patient YF5 three different scenarios were simulated and compared in order to

identify possible factors that triggered thrombus formation.

6.2.1 Simulation vs. lung perfusion

Figure 6.2a shows a comparison between simulation- and lung perfusion-

derived HFD. In the early post-operative stage, simulations agree within 10% of

with in-vivo measurements for all three patients. In patients YF1 and YF3, the

SVC flow blocked hepatic flow in the right and left branch respectively due to

a more proximal anastomosis. In patient YF3, a distal anastomosis for the left

branch which bypassed the stenosis in the LPA and a medial anastomosis for the

right branch resulted in over 60% of hepatic flow streaming to the LPA despite the

fact that the RPA received 60% of the total systemic venous flow.

Page 167: Surgical design for the Fontan procedure using computational fluid

143

YF3YF2YF1

a)

b)

0%10%20%30%40%50%60%70%80%90%

100%

Simula�on Lung perfusion

IVC-LPAIVC-RPA

PFS: 81/19

0%10%20%30%40%50%60%70%80%90%

100%

0 mon. post-op 6 mon. post op

IVC-LPAIVC-RPA

PFS: 81/19 PFS: 70/30

0%10%20%30%40%50%60%70%80%90%

100%

Simula�on Lung perfusion

IVC-LPAIVC-RPA

PFS: 60/40

0%10%20%30%40%50%60%70%80%90%

100%

Simula�on Lung perfusion

IVC-LPAIVC-RPA

PFS: 59/41

0%10%20%30%40%50%60%70%80%90%

100%

0 mon. post-op 6 mon. post-op

IVC-LPAIVC-RPA

PFS: 59/41 PFS: 53/47

0%10%20%30%40%50%60%70%80%90%

100%

0 mon. post-op 6 mon. post-op

IVC-LPAIVC-RPA

PFS: 60/40 PFS: 66/34

Figure 6.2: a) Comparison between early post-operative simulation-derived HFDand lung perfusion data for patients YF1, YF2 and YF3. HFD in the early post-operative stage was quantified by simulation and lung perfusion. b) Changes inHFD from the early to six-month post-operative stages derived from simulations.

6.2.2 Longitudinal HFD

HFD was predicted by particle tracking at the < one month and six month

time points, and pulmonary flow split was measured by MRI (Figure 6.2b). Lon-

gitudinally, HFD in patients YF1 and YF3 became more even as pulmonary flow

split approached 50/50 and IVC flow increased. In contrast, an opposite trend was

found in patient 2, but the changes in pulmonary flow split and HFD were minor.

The average change in HFD for these three patients is an 11% improvement and

in pulmonary flow split is 7.7%.

Page 168: Surgical design for the Fontan procedure using computational fluid

144

6.2.3 HFD estimation without in vivo flow conditions

Since PCMRI flow rates were not available for patients YF4 and YF6, HFD

was quantified under three assumed pulmonary flow splits of 35/65, 55/45 and

75/25. In patient YF4, this resulted in percentages of hepatic flow streaming to

the RPA of 55%, 69% and 88%, respectively. In patient YF6, results are 61%, 85%

and 95%, respectively. RPA-skewed HFD was attributed to the blockage effect of

SVC flow on the proximal anastomosis for the left branch.

6.2.4 Thrombus investigation

Because patient YF5 had an occurrence of thrombosis in the left branch,

this patient is examined in more detail for possible causes. The patient’s HFD

under different flow conditions is summarized in Table 6.2. In the early post-

operative period, hepatic flow in patient YF5 was highly skewed to the RPA due

to low resistance. Patient YF5’s HFD was still skewed due to the thrombus in the

left branch even though the patient’s pulmonary flow split was found to be 54/46

at the three month post-operative time point. If the left branch remained patent,

the HFD would be improved achieving 63/37 at the three month post-operative

time point.

To examine whether patient YF5 is different from other Y-graft patients

in WSS, Table 6.2 and Figure 6.3 show the time-averaged WSS magnitude for

patients YF1, YF2, YF3 and YF5. For patient YF5, there is a persistent region

Page 169: Surgical design for the Fontan procedure using computational fluid

145

Table 6.2: Mean WSS magnitude for Y-graft branches and HFD. Compared toother patients, patient YF5 had low WSS in the left branch in the early postoperative stage but the WSS in the left branch increased in the 3 month post-operative stage in which the pulmonary flow split changed from 81/19 to 54/46.The mean WSS for patient YF3 is low due to a lower cardiac output. The modifiedY-grafts for patient YF5 increased mean WSS in the left branch in the early post-operative stage compared to the original Y-graft. All Y-graft designs for patientYF5 skewed hepatic flow to the RPA with PAVMs in the early post-operative stagebut the HFD was improved in the 3 month post-operative stage.

Type Patient (flow cond.) Left branch Right branch HFD(dynes/cm2) (dynes/cm2)

Initial

YF1 (early) 5.0 10.8 63/37YF2 (early) 8.4 11.3 83/17YF3 (early) 4.2 5.4 35/65YF5 (early) 4.6 25 88/12

YF5 (3 mon.)9.1 22 63/37 (unclotted)N.A. 19.7 82/18 (clotted)

Modified

YF5 R14-L10 (early) 9.4 19.5 95/5YF5 R14-L10 (3 mon.) 12.7 13.8 72/28YF5 R12-L12 (early) 11.7 13.2 99/1YF5 R12-L12 (3 mon.) 9.9 8.1 85/15

Page 170: Surgical design for the Fontan procedure using computational fluid

146

of low WSS for the left branch for the early post-operative case when we have

a measured pulmonary flow split of 81/19. Overall, mean WSS is higher in the

right branch than the left branch. The WSS in the left branch increased in our

simulation of the later time point when pulmonary flow split is changed to 54/46.

To quantify the low WSS area which we hypothesize is related to thrombus

formation, we computed the branch surface area in which the WSS value is lower

than a threshold value. While it is known that low WSS is conducive to thrombosis

formation, the exact threshold value remains unknown, and is certain to vary

among patients.148 Figure 6.4 shows the percentage area of low WSS with different

threshold values for each branch in patients YF1, YF2, YF3 and YF5. Compared

to the early post-operative results for patients YF1, YF2 and YF3, YF5’s left

branch had a larger portion of low WSS for all threshold values used, though the

mean WSS magnitude for the left branch in patient YF3 was also low due to a

lower cardiac output. However, the percentage area of low WSS was significantly

reduced in the three month post-operative stage. For the right branch, patient

YF5 had the smallest low WSS area in both post-operative stages at all threshold

levels.

Velocity fields confirmed that an area of flow stagnation is evident in the

left branch of patient YF5 under a pulmonary flow split of 81/19. In models YF4

and YF6 with HFD similar to YF5, less flow stagnation region is observed due to

the impingement of the SVC jet.

Table 6.2 shows that the mean WSS increased with decreasing graft size,

Page 171: Surgical design for the Fontan procedure using computational fluid

147

YF3YF2YF1

YF5 YF5 R14-L10 YF5 R12-L12

Figure 6.3: Time-averaged WSS magnitude for patients YF1, YF2 YF3 and YF5in the early post-operative stage. YF5 R14-L10 and YF5 R12-L12 are two modifiedY-graft designs for patient YF5. In the baseline model for patient YF5, a distalanastomosis for the left branch and a highly skewed pulmonary flow split resultedin larger low WSS area in the left branch. In model R12-L12, the WSS in theleft branch was enhanced due to a proximal anastomosis that allowed SVC flow towash the left branch.

00.10.20.30.40.50.60.70.80.9

1

0.5 1 2 4

YF1

YF2

YF3

YF5 early post-opYF5 3 mon. post-op

cri�

cal a

rea

le� branch

τ0

0.10.20.30.40.50.60.70.80.9

1

0.5 1 2 4

YF1

YF2

YF3

YF5 early post-op.YF5 3 mon. post-op

τ

cri�

cal a

rea

right branch

Figure 6.4: Percentage of low WSS region for two branches in patients YF1, YF2,YF3 and YF5. For each threshold value τ , the low WSS area relative to eachbranch surface was computed.

Page 172: Surgical design for the Fontan procedure using computational fluid

148

00.10.20.30.40.50.60.70.80.9

1

0.5 1 2 4

R14-L10 le� branchR12-L12 le� branch

cri�

cal a

rea

τ

pulmonary flow split: 81/19

00.10.20.30.40.50.60.70.80.9

1

0.5 1 2 4

R14-L10 le� branchR12-L12 le� branch

cri�

cal a

rea

τ

3 mon post-op.pulmonary flow split: 54/46

early post-op.

Figure 6.5: Percentage of low WSS region for patient YF5’s modified Y-grafts. Inmodel R12-L12, the low WSS area in the left branch can be effectively minimizedby using a proximal anastomosis in which the SVC jet impinged the wall and theimpact of the SVC jet on the WSS was reduced with increasing LPA flow. Com-pared to the baseline model (Figure 6.4), model R14-L10 has a similar percentageof low WSS area for the left branch in the early post-operative stage for thresholdvalues below 2 dynes/cm2.

and that the proximal anastomosis resulted in less low WSS area in the left branch

due to the effect of the SVC jet. In Figure 6.5, model R12-L12 effectively minimized

the low WSS area in the left branch by using a proximal anastomosis. Compared to

the baseline model (see Figure 6.4), model R14-L10 achieved a similar percentage

of low WSS area for the left branch in the early post-operative stage for threshold

values below 2 dynes/cm2.

6.3 Discussion

In this study, we evaluated the hemodynamic performance for the first co-

hort of Y-graft Fontan patients. For the first time, we obtained MRI measurements

at two time points, which provided a direct comparison between the early and six-

month post-operative stages for Y-graft patients. Although only short-term data

Page 173: Surgical design for the Fontan procedure using computational fluid

149

are available, measurements at two post-operative time points show that the ra-

tios of caval flow and pulmonary flow split are not constant in time. Within six

months after Fontan completion, a uniform increase of 33% was observed in the

SVC flow. The IVC flow increased significantly by 105% in two patients, YF1

and YF3, but decreased by 31% in patient YF2. Although the systemic venous

flow increased in all patients, the IVC flow was still less than the SVC flow in all

patients except YF3 at the six-month post-operative time point. Similarly, Fogel

et al.33 and Houlind et al.34 previously reported that the SVC contributed more

flow to the venous return in lateral tunnel Fontan patients. This indicates that

Y-graft patients likely follow the same trend in systemic venous flow distribution

as other Fontan patients. At least four out of six patients in this study had a

RPA-predominant split, consistent with the measurements of Houlind et al.34

In validating simulations against the lung perfusion data, good agreement

was obtained in the three patients studied. The largest difference (10%) between

the simulation-derived HFD and lung perfusion in patient YF2 is still accept-

able, considering potential sources of uncertainty in both simulation and clinical

measurement. Although patient specific data were incorporated into previous sim-

ulation studies, simulation-derived results had not been directly validated against

clinical data in a blinded fashion in prior studies. The significance of this study

is a validation that flow simulations correctly quantify patients’ HFD, which is

an important parameter for Fontan surgical design. This provides a basis for fu-

ture predictive studies, albeit more patients should be used to achieve statistical

Page 174: Surgical design for the Fontan procedure using computational fluid

150

significance.

HFD is largely influenced by the anastomosis locations and amount of SVC

flow. In the first cohort of Y-graft patients, the Y-graft was anastomosed to the

PAs proximally relative to the SVC. For an uneven pulmonary flow split, a proximal

anastomosis on the PA side with lower resistance is beneficial for reducing skewed

hepatic flow. In patient YF1, the hepatic flow was relatively even despite a severely

skewed pulmonary flow split, since the SVC jet blocked a part of hepatic flow

streaming to the RPA. However, the same effect caused a highly uneven HFD in

patient YF2, who had a mildly uneven pulmonary flow split. These in-vivo results

were consistent with our previous virtual studies.3 Limited space in the chest cavity

is a major concern for implanting a Y-graft. Although the Y-graft in this study was

anastomosed proximally, it does not imply that the branch cannot be anastomosed

more distally. With more surgical experience, more distal anastomoses could be

made in some patients to improve HFD. For example, anastomosing the left branch

more distally could potentially avoid the SVC flow impingement in the left branch

in patients YF2, YF4 and YF6.

Unknown flow boundary conditions for patients YF4 and YF6 made simula-

tion results uncertain for these patients. Since there was good agreement between

simulation and lung perfusion for patients with MRI scans, the boundary condi-

tions are the major source of uncertainty for patients YF4 and YF6. Under a

range of pulmonary flow splits, patients YF4 and YF6 showed a consistent trend

that the hepatic flow streaming to the RPA increased with increasing RPA flow.

Page 175: Surgical design for the Fontan procedure using computational fluid

151

Since patient YF4 had PAVMs in the right upper lobe, we infer that HFD might

have changed from severely RPA-skewed to moderate RPA-skewed with PAVM

regression because PAVMs usually result in lower pulmonary resistance.13, 23

Since caval flow ratios and pulmonary resistances change with age, HFD is

unlikely to remain constant over time. We found that HFD values moved changed

towards 50/50 over time in patients who had increased IVC flow and a less skewed

pulmonary flow split. This agrees with our prior exercise simulation findings in,3

such that increasing IVC flow could mitigate the adverse effects of high SVC flow

and uneven pulmonary resistances on HFD.

The pre and post operative measurements in patient YF5 provide some in-

teresting insight into pulmonary remodeling in patients with PAVMs. Patient YF5

had a pre-operative pulmonary flow split of 81/19. It is well known that the lung

with PAVMs typically has a low pulmonary resistance due to a precapillary con-

nection between the systemic and pulmonary venous return.13, 23 At the 3 month

post-operative examination, the pulmonary flow distribution was nearly even and

PAVMs in the RPA had regressed. Based on the simulation results for patient

YF5, we can infer that an uneven pulmonary flow split for patients with PAVMs

prior to the Fontan procedure would likely change dramatically post surgery, and

that streaming hepatic flow to the malformed lung reversed the PAVMs in this

case, resulting in a less skewed pulmonary flow split.

For patients without PAVMs, the question of how pulmonary flow split

changes over time remains unclear and long term follow up is needed. Our data

Page 176: Surgical design for the Fontan procedure using computational fluid

152

suggest that the pulmonary flow split in patients without PAVMs is relatively

stable within six months. Recently, Yin et al.149 reported a five year follow-up

study on the pulmonary perfusion of 43 Fontan patients . Their results showed

that there are small differences (3% and 2%, respectively) in the pulmonary and

IVC flow perfusion between the early post-operative and follow-up groups, though

pulmonary vascular resistance and relative perfusion for upper and lower lobe

lung showed statistically significant differences after 5 years.149 However perfusion

data prior to the Fontan surgery were not available in their study. If there were

significant changes in pulmonary flow split, they likely occurred during the early

post-operative stage because a small difference was found between the early and

5 year post-operative time points. If these trends were universal, HFD would

remain relatively stable with possible larger changes in the relative distribution for

the upper and lower lobes. Relative hepatic and pulmonary flow distribution in

each lung should be examined in future clinical studies.

The risk of thrombosis is another important issue which should be consid-

ered in addition to HFD. In patient YF5, thrombus was found in the left branch,

even though PAVMs in the right lung had regressed. Although the choice of crit-

ical WSS value for thrombosis is still unclear, calculations with different critical

values consistently showed that a low WSS region was evident in the left branch

for patient YF5 in the early post-operative stage. These results indicate that low

WSS and flow stasis in the left branch increased thrombotic risk and consequently

led to complete occlusion of the left branch. Moreover, simulations with an un-

Page 177: Surgical design for the Fontan procedure using computational fluid

153

blocked Y-graft and a pulmonary flow split of 54/46 suggested that the thrombus

likely formed before the PAVM regression, which caused the pulmonary flow split

to drop from 81/19 to 54/46. Had PAVM regression occurred earlier, the increased

WSS level in the left branch may have prevented thrombus formation. Without

further clinical evidence and data to support this timing, these findings, however,

remain speculative.

Although the presumed early development of thrombus in the left branch

in patient YF5 may be driven by multiple factors, our findings suggest that low

WSS and flow stasis are likely to be important causes. Thus, for patient YF5, the

immediate surgical design goal should have been to distribute the hepatic flow to

the lung with PAVMs without introducing a large low WSS area, while at the same

time keeping the long-term goal to achieve even HFD after PAVMs are regressed.

Compared to the original Y-graft for patient YF5, an 18-12-12 Y-graft

with a proximal anastomose for the left branch (YF5 R12-L12) minimized the

low WSS area in the simulated early post-operative stage, though HFD was still

skewed at the early time point. However, HFD is later improved with increasing

IVC flow.3 The unequal sized Y-graft (YF5 R14-L10) did not show substantial

improvement in minimizing the lowWSS area. Previous optimal designs considered

a single objective solely. The formation of thrombus in patient YF5 showed that

a single objective design without considering the dynamic process of pulmonary

remodeling and thrombotic risk is inadequate. Future studies should incorporate

multi-objective constrained optimization to balance the need for even HFD with

Page 178: Surgical design for the Fontan procedure using computational fluid

154

thrombotic risk. Our results suggest that the the first three months post Fontan

may be a critical time period to prevent thrombosis in the Y-graft for patients

with PAVMs, and that the risk likely drops after PAVM regression. As pointed

out in our previous studies, the superiority of the Y-graft cannot be guaranteed

without a customized optimal design for each patient. In addition to optimizing

the Y-graft geometry to minimize low WSS area, more aggressive anticoagulation

therapy may be needed for patients during the initial high risk period. MRI and

CFD could be used in these cases to monitor the WSS level in Y-graft Fontan

patients, providing extra information for their treatment.

Recently, Haggerty et al.150 reported post-operative simulation results for

five Y-graft Fontan patients at Children’s Healthcare of Atlanta showing that com-

pared to virtual T-junction and offset designs, the Y-graft design resulted in im-

proved HFD but little difference in the connection resistance. These findings are

likely due to the use of a smaller branch size (9mm and 10mm) used for the Y-graft

and more idealized graft geometry for the virtual tube-shaped models. Instead of

custom Y-grafts, commercial bifurcated aorto-illiac grafts were used in the Y-graft

Fontan surgery in their study.150 Although use of an “off the shelf” graft is consis-

tent and well controlled in terms of fabrication, it is less flexible than the custom

Y-graft in the choice of geometry such as trunk-branch size and bifurcation angle.

Since Fontan patients show a variety of anatomic patterns, a custom graft may

be more suitable for customized Fontan design. In addition, our findings indicate

that anastomosis location is likely equally or more important than the graft branch

Page 179: Surgical design for the Fontan procedure using computational fluid

155

diameter in determining WSS and HFD values.

6.4 Limitations

The follow up time in this study was limited and the number of study

patients was relatively small. It is still too early to answer questions such as how

pulmonary and hepatic flow distributions change over a long period and whether

the Y-graft results in better mid-term or long term outcomes than traditional

designs. Longitudinal and serial data collection including the pre-Fontan stage

should be performed on multiple patients in future studies.

Owing to a lack of PC-MRI measurements in the early post-operative stage

for patients YF4, YF5 and YF6, flow boundary conditions were approximated.

Since patient YF5 underwent pre-operative and three month post-operative MRI

examinations, results for the early post-operative stage were less uncertain com-

pared to patients YF4 and YF6 who had CT images only.

The prediction of post-operative pulmonary flow split from pre-operative

data should be addressed systematically by multi-scale modeling in future studies.

Rigid wall and Newtonian flow assumptions were employed in this study. Long et

al.58 showed that differences in HFD and energy efficiency due to wall compliance

are small but that WSS was over-predicted by up to 17% in rest conditions by rigid

wall simulations.57 Future accurate predictions of thrombotic risk may therefore

necessitate use of fluid structure interaction, however the relative values used in

Page 180: Surgical design for the Fontan procedure using computational fluid

156

this study may be adequate for risk assessment.

6.5 Conclusions

By comparing in-vivo lung perfusion data, the accuracy of simulation-

derived HFD was validated. Although the technical success of the Y-graft Fontan

surgery demonstrated the feasibility of the Y-graft concept, simulations for the

first six patients show that a proximal anastomosis of the left branch resulted in

uneven HFD due to the SVC blockage effect in some patients. Overall, hepatic

flow was skewed to the RPA in the early post-operative stage and HFD was im-

proved as the IVC flow increased and pulmonary flow split became less unequal.

It was confirmed that the overall pulmonary flow split changes over time. A pa-

tient with unilateral PAVMs showed a significant change in pulmonary flow split

after Fontan completion due to the regression of PAVMs while variations in other

patients were less pronounced. Simulations can provide insight in patients with

adverse events. Although sufficient hepatic flow was channeled to the malformed

right lung to resolve PAVMs, thrombus likely developed in one branch with stag-

nant flow. Compared to other patients, YF5’s left branch had a larger region of

low WSS in the early post-operative period. Thus a plausible explanation is that

the thrombus in the left branch developed due to flow stasis soon after Fontan

completion, and that PAVM regression occurred after this. Therefore, particular

attention should be paid to WSS values and flow stasis in the early post-operative

Page 181: Surgical design for the Fontan procedure using computational fluid

157

period in order to reduce the thrombotic risk in the Y-graft. This should be ex-

amined during the pre-operative surgical planning phase, particularly for patients

with existing PAVMs.

6.6 Acknowledgments

This study was supported by the American Heart Association, a Burroughs

Wellcome Fund Carreer Award at the Scientific Interface, a Leducq Foudantion

Network of Excellent Grant, a NSF CAREER Award and a UCSD Kaplan Fel-

lowship. We thank Shawn Shadden, Irene Vignon-Clementel, Mahdi Esmaily

Moghadam, and John Lamberti for their expertise in numerical simulations and

pediatric cardiac surgery, as well as Christina Ngo for model construction. We also

wish to acknowledge the use of Simvascular (simtk.org, www.osmsc.com).

Chapter 6, in full, is a reprint of the material as it appears in Yang, W.,

Chan, F. P., Feinstein, Reddy, V. M., Marsden, A. L., and Feinstein, J. A. Flow

Simulations and Validation for the First Cohort of Y-graft Fontan Patients., in

preparation.

Page 182: Surgical design for the Fontan procedure using computational fluid

Chapter 7

Conclusions and future work

7.1 Conclusions

In this dissertation, we focused on hemodynamics of a novel Y-graft de-

signed for the Fontan procedure using numerical simulations. This work extended

the preliminary results obtained by Marsden and colleagues.2, 4, 76 We applied a

derivative-free optimization framework to evaluate and improve upon designs for

the Fontan surgery. Simulations in idealized Y-graft models showed that a 24mm

branch size (maximum size) with a large bifurcation angle achieved the highest en-

ergy efficiency at rest and the graft size decreased with increasing IVC flow under

exercise conditions. Optimal designs for the rest condition were significantly influ-

enced by the wall shear stress (WSS) constraint showing a trade-off relationship

between the energy efficiency and low WSS area. However this impact was less

significant under exercise conditions because a smaller graft size resulted in less

158

Page 183: Surgical design for the Fontan procedure using computational fluid

159

energy loss. Although this is the first study that applied formal optimal design

to the Fontan procedure, the use of idealized models resulted in low energy loss

values that are not realistic in the in-vivo Fontan circulation. Later surgical prac-

tice suggested that a Y-graft with 24 mm branches may not be feasible in 2-4 year

old Fontan patients. In addition, a recent multiscale modeling study by Baretta et

al.54 showed that differences in energy loss with different Fontan geometries had

negligible effects on the ventricular pressure-volume loop. Thus, the hypothesis

that lower hydrodynamic energy loss improves clinical outcomes is not supported

by recent studies, although further investigations are required to solidify these

findings. Based on our current results, the importance of energy efficiency in the

Fontan geometry is secondary to other factors.

Studies in multiple patient specific models showed that the hemodynamic

performance of the Y-graft design was patient specific. We found that the pul-

monary resistance, anastomosis location, flow condition and stenosis treatment

play important roles in local hemodynamic performance. Instead of energy loss,

more attention was paid to the pulmonary flow distribution (HFD). The SVC flow

was found to be a major reason causing skewed HFD because the downward SVC

jet blocked the IVC flow, resulting in unilateral streaming with as much as 97%.

Compared to the traditional T-junction and offset designs, the Y-graft design dis-

tributed hepatic flow to two lungs more evenly with moderate improvement in

energy loss and SVC pressure. However, the cases of underperforming Y-graft de-

signs demonstrated the importance and needs for customizing surgical designs for

Page 184: Surgical design for the Fontan procedure using computational fluid

160

individual patients.

The findings in the patient specific study motivated us to investigate the

necessity of the use of unequal branches for patients with highly uneven pulmonary

flow splits. We demonstrated that optimized Y-graft with equal sized branches can

achieve the same target HFD with less energy loss for a large range of pulmonary

flow splits compared to the design with unequal sized branches. Anastomosis

locations for the Y-graft are more important in determining HFD than the graft

size. Although the SVC jet blockage effect was attributed to skewed HFD, it

can be utilized to optimize HFD by anastomosing the branch more proximally

(towards the SVC) for the PA with a smaller resistance value. We effectively

improved previously underperforming Y-grafts in two patients by optimizing semi-

idealized models. Comparisons between the semi-idealized and patient specific

models showed that the semi-idealized model can approximate the patient specific

model for HFD prediction with treatments for the PA and right upper lobe adjacent

to the SVC.

Based on these simulation results, the Y-graft design has been translated

into clinical use. The technical success demonstrated the feasibility of the Y-graft

design. Simulation derived HFD showed excellent agreement with in vivo lung per-

fusion data for three patients in the immediate post-operative stage. However non-

optimized Y-grafts did not achieve a 50/50 split for hepatic flow, channeling more

hepatic flow to the right lung in most patients. The six month post-operative simu-

lations and MRI measurements showed that increased IVC flow rate and more even

Page 185: Surgical design for the Fontan procedure using computational fluid

161

pulmonary flow split improved HFD, compared to the immediate post-operative

data. Investigations on thrombus formation demonstrated that skewing hepatic

flow to the malformed lung might facilitate the regression of PAVMs but resulted

in larger areas of low WSS and increased flow stagnation, which might be a pos-

sible factor causing thrombus. This case suggested that future surgical design

should carefully balance the need of hepatic flow for the lung with PAVMs and

the risk of thrombosis due to low WSS, together with considerations of pulmonary

remodeling.

7.2 Future work

7.2.1 Pre-operative prediction and assessment

Post-operative validation with lung perfusion data demonstrated the capa-

bility of numerical simulations to correctly calculate physiologically relevant pa-

rameters with MRI-derived boundary conditions. Current computational tools are

able to evaluate individual Fontan patient’s hemodynamic performance with rea-

sonable accuracy. However clinicians also need a tool capable of predicting post-

operative performance based on pre-operative data in order to choose the best

surgical plan. Therefore our future studies should focus on predictive simulation

capabilities for the Fontan procedure.

Although similar issues have been involved in the comparison study using

multiple patient specific models in Chapter 4, the assumption that patients’ pul-

Page 186: Surgical design for the Fontan procedure using computational fluid

162

monary flow split and cardiac output remain unchanged limits the scope of the

study within the immediate post-operative period. Our post-operative follow up

showed that there are nonnegligible changes in the pulmonary flow split and caval

flow within six month after Fontan completion. Therefore future predictive simula-

tions have to account for these changes in the boundary conditions. The multiscale

modeling methods54 that couple closed-loop lumped parameter networks to a 3D

flow solver offer a promising means to systematically model the effects of physio-

logic changes due to the surgery.

We still know little about Fontan patients’ hemodynamic evolution. The

study on the first cohort of Y-Fontan patients has not yet provided enough inputs

to predict the outcomes of the Y-graft design due to a small sample size and a

short follow up period. Therefore, launching a long term follow up study with

a sufficient sample size is crucial to determine the actual outcomes and provide

first-hand data for predictive modeling.

7.2.2 Patient specific optimal design

As we showed in this work, the variability of the Fontan geometry lead to

dramatically different flow fields, and idealized models are unable to fully char-

acterize individual Fontan patient’s hemodynamics. Therefore, patient specific

designs are necessary. A few attempts have been made in patient specific design

optimization.151,152 The major limitations included the use of a trial-and-error ap-

proach and optimized Y-graft derived from the semi-idealized models. Although

Page 187: Surgical design for the Fontan procedure using computational fluid

163

some treatments were applied to the semi-idealized model in order to approximate

the corresponding patient specific model, discrepancies were inevitable and manual

interventions were needed. In addition, the assumption about the size of the end-

to-side anastomosis is empirical and qualitative. In future studies, a physics-based

model manipulation method for the end-to-side anastomosis will be developed such

that one can manipulate the Y-graft design in patient specific models without la-

borious manual interventions. One possible way is to model the surface of the

PA and graft as an elastic structure. The geometry of the graft could then be

determined by applying forces to deform the graft and the end-to-side anastomosis

could be modeled by pressurizing the elastic structure.

The semi-idealized model can be used as a low fidelity model which is less

expensive to evaluate compared to the patient specific model. Recent optimization

studies took advantages of low fidelity models to enhance the efficiency of expensive

simulation based optimization.144

7.2.3 Validation against 4D MRI

Although simulation-derived HFD has been validated against to lung per-

fusion data, the in-vivo velocity field and WSS have not been compared. With

the advent of new imaging technology, 4D MRI allows one to measure in-vivo

time-dependent 3D flow. Although current spatial resolution limits our ability to

quantify WSS and highly dynamic flow structures,39 4D MRI provides a valuable

tool to comprehensively compare with CFD results and we can expect to be able to

Page 188: Surgical design for the Fontan procedure using computational fluid

164

perform direct validations for WSS calculations and turbulence modeling in blood

flow in the near future.

Page 189: Surgical design for the Fontan procedure using computational fluid

Bibliography

[1] D. D. Soerensen, K. Pekkan, D. de Zelicourt, S. Sharma, K. Kanter, M. Fogel,and A. Yoganathan. Introduction of a new optimized total cavopulmonaryconnection. Ann. Thorac. Surg., 83(6):2182–90, 2007.

[2] A. L. Marsden, A. J. Bernstein, V. M. Reddy, S. Shadden, R. L. Spilker,F. P. Chan, C. A. Taylor, and J. A. Feinstein. Evaluation of a novel Y-shaped extracardiac Fontan baffle using computational fluid dynamics. J.Thocac. Cardio. Surg., 137(2):394–403, 2009.

[3] W. Yang, I. E. Vignon-Clementel, G. Troianowski, V. M. Reddy, J. A. Fein-stein, and A. L. Marsden. Hepatic blood flow distribution and performancein traditional and Y-graft Fontan geometries: A case series computationalfluid dynamics study. J. Thorac. Cardiovasc. Surg., 143:1086–1097, 2012.

[4] A. L. Marsden, I. E. Vignon-Clementel, F. Chan, J. A. Feinstein, and C. A.Taylor. Effects of exercise and respiration on hemodynamic efficiency inCFD simulations of the total cavopulmonary connection. Ann. Biomed. Eng.,35(2):250–263, 2007.

[5] G. Troianowski, C. A. Taylor, J. A. Feinstein, and I. Vignon-Clementel.Three-dimensional simulations in Glenn patients: clinically based boundaryconditions, hemodynamic results and sensitivity to input data. J. Biomech.Eng., 133(11), 2011. DOI:10.1115/1.4005377.

[6] Roger et al. Heart disease and stroke statistics - 2011 update: A report fromthe american heart association, 2011.

[7] M. A. Canfield, M. A. Honein, N. Yuskiv, J. Xing, C. T. Mai, J. S. Collins,O. Devine, J. Petrini, T. A. Ramadhani, C. A. Hobbs, and R. S. Kirby.National estimates and race/ethnic-specific variation of selected birth defectsin the united states, 1999-2001. Birth Defects Res A Clin Mol Teratol.,76(11):747–756, 2006.

[8] F. Fontan and E. Baudet. Surgical repair of tricuspid atresia. Thorax, 26:240–248, 1971.

[9] M. R. de Leval. The Fontan circulation: a challenge to William Harvey?Nat. Clin. Pract. Cardiovasc. Med., 2:202–208, 2005.

165

Page 190: Surgical design for the Fontan procedure using computational fluid

166

[10] B. S. Marino. Outcomes after the Fontan procedure. Curr. Opin. Pediatr.,14:620–626, 2002.

[11] A. M. Rudolph. Congenital Diseases of the Heart: Clinical-PhysiologicalConsiderations. Wiley-Blackwell, Hoboken, 2009.

[12] B. W. Migliavacca, V. Zak, L. A. Sleeper, S. M. Paridon, S. D. Colan,T. Geva, L. Mahony, J. S. Li, R. E. Breitbart, R Margossian, R. V. Williams,W. M. Gersony, and A. M. Atz. Laboratory measures of exercise capacity andventricular characteristics and function are weakly associated with functionalhealth status after fontan procedure. Circulation, 121:34–42, 2010.

[13] D. A. de Zelicourt, A. L. Marsden, M. A. Fogel, and A. P. Yoganathan.Imaging and patient-specific simulations for the fontan surgery: Currentmethodologies and clinical applications. Progress in Pediatric Cardiology,30:31–44, 2010.

[14] S.M. Paridon, P. D. Mitchell, S.D. Colan, R.V. Williams, A. Blaufox, J. S. Li,R. Margossian, S. Mital, J. Russell, and J. Rhodes. A cross-sectional studyof exercise performance during the first 2 decades of life after the fontanoperation. J Am Coll Cardiol, 52:99–107, 2008.

[15] M.H. Gewillig, U.R. Lundstrom, C. Bull, R.K. Wyse, and J.E Deanfield. Ex-ercise responses in patients with congenital heart disease after fontan repair:patterns and determinants of performance. J Am Coll Cardiol, 15:1424–1432,1990.

[16] A. Giardini, A. Hager, C. P. Napoleone, and F. M. Picchio. Natural historyof exercise capacity after the Fontan operation: a longitudinal study. Ann.Thorac. Surg., 85(3):818–821, 2008.

[17] A. Giardini, C. P. Napoleone, S. Specchia, A. Donti, R. Formigari, G. Op-pido, G. Gargiuloand, and F. M. Picchio. Conversion of atriopulmonaryfontan to extracardiac total cavopulmonary connection improves cardiopul-monary function. Int J Cardiol, 113(3):341–344, 2006.

[18] A. C. Guyton, B Abernathy, J. B. Langston, B. N. Kaufmann, and H. M.Fairchild. Relative importance of venous and arterial resistances in control-ling venous return and cardiac output. Am. J. Physiol., 196(5):1008–1014,1959.

[19] K. S. Sundareswaran, K. Pekkan, L. P. Dasi, K. Whitehead, S. Sharma,K. Kanter, M. A. Fogel, and A. P. Yoganathan. The total cavopulmonaryconnection resistance: a significant impact on single ventricle hemodynamicsat rest and exercise. Am J Physiol Heart Circ Physiol, 295:H2427–H2435,2008.

[20] P. D. Coon, J. Rychik, R. T. Novello, P. S. Ro, J. W. Gaynor, and T. L.Spray. Thrombus formation after the Fontan operation. Ann. Thorac. Surg.,71(6):1990–4, 2001.

Page 191: Surgical design for the Fontan procedure using computational fluid

167

[21] L. K. Shirai, D. N. Rosenthal, B. A. Reitz, R. C. Robbins, and A. M. Du-bin. Arrhythmias and thromboembolic complications after the extracardiacfontan operation. J. Thorac. Cardio. Surg., 115(3):499–505, 1998.

[22] V. Alexi-Meskishvili, S. Ovroutski, P. Ewert, I. Dahnert, F. Berger, P. E.Lange, and R. Hetzer. Optimal conduit size for extracardiac Fontan opera-tion. Eur. J. of Cardio-Thorac. Surg., 18:690–695, 2000.

[23] B. W. Duncan and S. Desai. Pulmonary arteriovenous malformations aftercavopulmonary anastomosis. Ann. Thorac. Surg., 76:1759–1766, 2003.

[24] J. R. Grossage and G. Kanj. Pulmonary arteriovenous malformations. AmJ Respir Crit Care Med, 158:643–661, 1998.

[25] H. Uemura, T. Yagihara, R. Hattori, Y. Kawahira, S. Tsukano, andK. Watanabe. Redirection of hepatic venous drainage after total cavopul-monary shunt in left isomerism. Ann. Thorac. Surg., 68:1731–1735, 1999.

[26] N. A. Pike, L. A. Vricella, J. A. Feinstein, M. D. Black, and B. A. Reitz.Regression of severe pulmonary arteriovenous malformations after Fontanrevision and hepatic factor rerouting. Ann. Thorac. Surg., 78:697–9, 2004.

[27] D. B. McElhinney, G. R. Marx, A. C. Marshall, J. E. Mayer, and P. J. delNido. Cavopulmonary pathway modification in patients with heterotaxy andnewly diagnosed or persistent pulmonary arteriovenous malformations aftera modified Fontan operation. J. Thocac. Cardio. Surg., 141(6):1362–1370,2011.

[28] Y. Imoto, A. Sese, and K. Joh. Redirection of the hepatic venous flow for thetreatment of pulmonary arteriovenous malformations after Fontan operation.Pediatr. Cardiol, 27:490–492, 2006.

[29] M. S. Pearce, J. A. Salotti, M. P. Little, K. McHugh, C. Lee, K. P. Kim, N. L.Howe, G. M. Ronckers, P. Rajaraman, A. W. Craft, L. Parker, and A. B.de Gonzalez. Radiation exposure from ct scans in childhood and subsequentrisk of leukaemia and brain tumours: a retrospective cohort study. TheLancet, 2012. doi:10.1016/S0140-6736(12)60815-0.

[30] M. M. Samyn. A review of the complementary information available withcardiac magnetic resonance imaging and multi-slice computed tomography(ct) during the study of congenital heart disease. Int. J. Cardiovas. Imag.,20:569–578, 2004.

[31] M. A. Seliem, J. Murphy, J. Vetter, S. Heyman, and W. Norwood. Lung per-fusion patterns after bidirectional cavopulmonary anastomosis (hemi-Fontanprocedure). Pediatr. Cardiol, 18:191–196, 1997.

[32] M. J. Shah, J. Rychik, M. A. Fogel, J. O. Murphy, and M. L. Jacobs. Abnor-mal distribution of pulmonary blood flow after the Glenn shunt or Fontanoperation: risk of development of arteriovenous fistulae. Circulation, 72:471–479, 1985.

Page 192: Surgical design for the Fontan procedure using computational fluid

168

[33] M. A. Fogel, P. M. Weinberg, J. Rychik, A. Hubbard, M. Jacobs, T. L.Spray, and J. Haselgrove. Caval contribution to flow in the branch pulmonaryarteries of Fontan patients with a novel application of magnetic resonancepresaturation pulse. Circulation, 99:1215–1221, 1999.

[34] K. Houlind, E. Stenbog, K. Sorensen, K. Emmertsen, O. K. Hansen, L. Ry-bro, and V. E. Hjortdal. Pulmonary and caval flow dynamics after totalcavopulmonary connection. Heart, 81:67–72, 1999.

[35] A. Hager, S. Fratz, M. Schwaiger, R. Lange, J. Hess, and H. Stern. Pul-monary blood flow patterns in patients with fontan circulation. Ann. Thorac.Surg., 85:186–191, 2008.

[36] V. E. Hjortdal, K. Emmertsen, E. Stenbog, T. Frund, M. Rahbek Schmidt,O. Kromann, K. Sorensen, and E. M. Pedersen. Effects of exercise andrespiration on blood flow in total cavopulmonary connection: A real-timemagnetic resonance flow study. Circulation, 108(10):1227–1231, 2003.

[37] E. Be’eri, S. E. Maier, M. J. Landzberg, T. Chung, and T. Geva. In vivo eval-uation of fontan pathway flow dynamics by multidimensional phase-velocitymagnetic resonance imaging. Circulation, 98(25):2873–2882, 1998.

[38] K. S. Sundareswaran, C. M. Haggerty, D. de Zelicourt, L. P. Dasi, K. Pekkan,D. H. Frakes, A. J. Powell, K. Kanter, M. A. Fogel, and A. P. Yoganathan.Visualization of flow structures in fontan patients using 3-dimensionalphase contrast magnetic resonance imaging. J. Thorac. Cardiovasc. Surg.,143:1108–1116, 2012.

[39] M. Markl, J. Geiger, P. J. Kilner, D. Foll, B. Stiller, F. Beyersdorf, R. Arnold,and A. Frydrychowicz. Time-resolved three-dimensional magnetic resonancevelocity mapping of cardiovascular flow paths in volunteers and patients withfontan circulation. Eur. J. Cardio. Thorac. Surg., 39:206–212, 2011.

[40] C. M. Gibson, L. Diaz, K. Kandarpa, F. M. Sacks, R. C. Pasternak, T. San-dor, C. Feldman, and P. H. Stone. Relation of vessel wall shear stressto atherosclerosis progression in human coronary arteries. Arterioscler.Thromb., 13:310–315, 1993.

[41] T. A. Tasciyan, R. Banerjee, Y. I. Cho, and R. Kim. Two dimensional pul-satile hemodynamic analysis in the magnetic resonance angiography interpre-tation of a stenosed carotid arterial bifurcation. Med. Phys., 20:1059–1070,1993.

[42] C. A. Taylor and C. A. Figueroa. Patient-specific modeling of cardiovascularmechanics. Annu. Rev. Biomed. Eng., 11:109–134, 2009.

[43] G. Dubini, M. R. de Leval, R. Pietrabissa, F. M. Montevecchi, andR. Fumero. A numerical fluid mechanical study of repaired congenital heartdefects: Application to the total cavopulmonary connection. J. Biomech.,29(1):111–121, 1996.

Page 193: Surgical design for the Fontan procedure using computational fluid

169

[44] M. R. de Leval, G. Dubini, F. Migliavacca, H. Jalali, G. camporini, A. Red-ington, and R. Pietrabissa. Use of computational fluid dynamics in the designof surgical procedures: application to the study of competitive flows in cavo-pulmonary connections. J. Thorac. Cardiovasc. Surg., 111(3):502–13, 1996.

[45] F. Migliavacca, G. Dubini, E. L. Bove, and M. R. de Leval. Computa-tional fluid dynamics simulations in realistic 3-D geometries of the totalcavopulmonary anastomosis: the influence of the inferior caval anastomosis.J. Biomech. Eng., 125:805–813, 2003.

[46] K. K. Whitehead, K. Pekkan, H. D. Kitahima, S. M. Paridon, A. P. Yo-ganathan, and M. A. Fogel. Nonlinear power loss during exercise in single-ventricle patients after the Fontan: insights from computational fluid dy-namics. Circulation, 116:I–165 – I–171, 2007.

[47] K. Ryu, T. M. Healy, A. E. Ensley, S. Sharma, C. Lucas, and A. P. Yo-ganathan. Importance of accurate geometry in the study of the total cavopul-monary connection: Computational simulations and in vitro experiments.Ann. Biomed. Eng., 29:844–853, 2001.

[48] L. P. Dasi, R. Krishnankuttyrema, H. Katajima, K. Pekkan, K. Sun-dareswaran, M. A. Fogel, S. Sharma, K. Whitehead, K. Kanter, and A. P.Yoganathan. Fontan hemodynamics: importance of pulmonary artery diam-eter. J. Thorac. Cardiovasc. Surg., 137(3):560–564, 2009.

[49] C. G. De Groff and R. Shandas. Designing the optimal total cavopulmonaryconnection: pulsatile versus steady flow experiments. Med. Sci. Monit.,8(3):MT41–MT45, 2002.

[50] I. E. Vignon-Clementel, A. L. Marsden, and J. A. Feinstein. A primer on com-putational simulation in congenital heart disease for the clinician. Progressin Pediatric Cardiology, 30(1-2):3–13, 2010.

[51] A. L. Marsden, V. Mohan Reddy, S. C. Shadden, F. P. Chan, C. A. Taylor,and J. A. Feinstein. A new multi-parameter approach to computationalsimulation for Fontan assessment and redesign. Congenital Heart disease,5(2):104–117, 2010.

[52] E. L. Bove, M. R. de Leval, F. Migliavacca, G. Guadagni, and G. Dubini.Computational fluid dynamics in the evaluation of hemodynamic perfor-mance of cavopulmonary connections after the Norwood procedure for hy-poplastic left heart syndrome. J. Thorac. Cardiovasc. Surg., 126:1040–1047,2003.

[53] L. P. Dasi, K. Whitehead, K. Pekkan, D. de Zelicourt, H. Katajima, K. Sun-dareswaran, K. Kanter, M. A. Fogel, and A. P. Yoganathan. Pulmonaryhepatic flow distribution in total cavopulmonary connections: Extracardiacversus intracardiac. J. Thorac. Cardiovasc. Surg., 141:207–214, 2011.

Page 194: Surgical design for the Fontan procedure using computational fluid

170

[54] A. Baretta, C. Corsini, W. Yang, I. E. Vignon-Clementel, A. L. Marsden,T. Y. Hsia, G. Dubini, F. Migliavacca, and G. Pennati. Virtual surgeriesin patients with congenital heart disease: a multiscale modelling test case.Phil. Trans. R. Soc., 369(1954):4316–4330, 2011.

[55] W. Orlando, J. Hertzberg, R. Shandas, and C. DeGroff. Reverse flow incompliant vessels and its implications for the fontan procedure: numericalstudies. Biomed. Sci. Instrum., 38:321–326, 2002.

[56] W. Orlando, J. Hertzberg, R. Shandas, and C. DeGroff. Efficiency differencesin computational simulations of the total cavo-pulmonary circulation withand without compliant vessel walls. Biomed. Sci. Instrum., 81(3):220, 2006.

[57] Y. Bazilevs, M.-C. Hsu, D.J. Besnon, S. Sankaran, and A. L. Marsden. Com-putational fluid-structure interaction: Methods and application to a totalcavopulmonary connection. Comput Mech, 45(1):77–89, 2008.

[58] C. C. Long, M-C. Hsu, Y. Bazilevs, J. A. Feinstein, and A. L. Marsden.Fluidcstructure interaction simulations of the fontan procedure using variablewall properties. Int. J. Numer. Meth. Biomed. Engng., 28:513–527, 2012.

[59] M. R. de Leval, P. Kilner, M. Gewillig, and C. Bull. Total cavopulmonaryconnection: a logical alternative to atriopulmonary connection for complexFontan operations. experimental studies and early clinical experience. J.Thorac. Cardiovasc. Surg., 96:682–695, 1988.

[60] H. T. Low, Y. T. Chew, and C. N. Lee. Flow studies on atriopulmonaryand cavopulmonary connections of the fontan operations for congenital heartdefects. J. Biomech. Eng., 15:303–307, 1993.

[61] S. Sharma, S. Goudy, P. Walker, S. Panchal, A. Ensley, K. Kanter, V. Tam,D. Fyfe, and A. Yoganathan. In vitro flow experiments for determination ofoptimal geometry of total cavopulmonary connection for surgical repair ofchildren with functional single ventricle. J. Am. Coll. Cardiol, 27(5):1264–1269, 1996.

[62] A. E. Ensley, P. Lynch, G. P. Chatzimavroudis, C. Lucas, S. Sharma, andA. P. Yoganathan. Toward designing the optimal total cavopulmonary con-nection: an in vitro study. Ann. Thorac. Surg., 68:1384–1390, 2000.

[63] C.G. DeGroff, J.D. Carlton, C.E. Weinberg, M.C. Ellison, R. Shandas, andL. Valdes-Cruz. Effect of vessel size on the flow efficiency of the total cavopul-monary connection: In vitro studies. Pediatr. Cardiol., 23(2):171–177, 2002.

[64] A. C. Lardo, S. A. Webber, I. Friehs, P. J. del Nido, and E. G. Cape. Fluiddynamic comparison of intra-atrial and extracardiac total cavopulmonaryconnections. J. Thorac. Cardiovasc. Surg., 117(4):697–704, 1999.

[65] D. A. de Zelicourt, K. Pekkan, L. Wills, K. Kanter, J. Forbess, S. Sharma,M. Fogel, and A. P. Yoganathan. In vitro flow analysis of a patient-specific

Page 195: Surgical design for the Fontan procedure using computational fluid

171

intraatrial total cavopulmonary connection. Ann. Thorac. Surg., 79:2094–2102, 2005.

[66] Y. Khunatorn, R. Shandas, C. G. DeGroff, and S. Mahalingam. Comparisonof in vitro velocity measurements in a scaled total cavopulmonary connectionwith computational predictions. Ann Biomed Eng, 31:810–822, 2003.

[67] K. Pekkan, D. de Zelicourt, L. Ge, F. Sotiropoulo, D. Frakes, M. A. Fogel,and A. Yoganathan. Physics-driven cfd modeling of complex anatomicalcardiovascular flows-a tcpc case study. Ann Biomed Eng, 33(3):284–300,2005.

[68] P. G. Walker, T. T. Howea, R. L. Daviesa, J. F., and K. G. Wattersonb.Distribution of hepatic venous blood in the total cavopulmonary connection:an in vitro study. Eur. J. Cardio. Thorac. Surg., 17(6):658–665, 2000.

[69] P. G. Walker, G. F. Oweis, and K. G. Watterson. Distribution of hepaticvenous blood in the total cavo pulmonary connection: An in vitro study intothe effects of connection geometry. J Biomech Eng., 123(6):558–564, 2001.

[70] R. S. Figliola, A. Giardini, T. Conover, T. A. Camp, G. Biglino, J. Chiulli,and T-Y. Hsia. In vitro simulation and validation of the circulation withcongenital heart defects. Progress in Pediatric Cardiology, 30:71–80, 2010.

[71] M. P. Bendsoe and O. Sigmund. Topology Optimization: Theory, Methods,and Applications. Springer, Berlin, 2003.

[72] M. A. Abramson. Mixed variable optimization of a load-bearing thermalinsulation system using a filter pattern search algorithm. Optimization andEngineering, 5(2):157–177, 2004.

[73] G. Papageorgakis and D. Assanis. Optimizing gaseous fuel-air mixing indirect injection engines using an rng based k-ε model. SAE Technical Paper980315, 1998.

[74] B. Mohammadi and O. Pironneau. Applied Shape Optimization for Fluids.Oxford University Press, Oxford, 2001.

[75] B. Mohammadi and O. Pironneau. Shape optimization in fluidmechanics.Annu. Rev. Fluid Mech., 36:255–279, 2004.

[76] A. L. Marsden, J. A. Feinstein, and C. A. Taylor. A computational frame-work for derivative-free optimization of cardiovascular geometries. Comput.Methods Appl. Mech. Engrg., 197(21-24):1890–1905, 2008.

[77] A. Jameson. Aerodynamic design via control theory. J. Sci. Comput., 3:233–260, 1988.

[78] A. L. Marsden, M. Wang, J. E. Dennis, Jr., and P. Moin. Optimal aeroacous-tic shape design using the surrogate management framework. Optimizationand Engineering, 5(2):235–262, 2004. Special Issue: Surrogate Optimization.

Page 196: Surgical design for the Fontan procedure using computational fluid

172

[79] A. L. Marsden, M. Wang, J. E. Dennis, Jr., and P. Moin. Suppressionof airfoil vortex-shedding noise via derivative-free optimization. Physics ofFluids, 16(10):L83–L86, 2004.

[80] A. L. Marsden, M. Wang, J. E. Dennis, Jr., and P. Moin. Trailing-edge noisereduction using derivative-free optimization and large-eddy simulation. J.Fluid Mech., 572:13–36, 2007.

[81] A. N. Brooks and Thomas J. R. Hughes. Streamline upwind/petrov-galerkinformulations for convection dominated flows with particular emphasis on theincompressible navier-stokes equations. Comput. Meth. Appl. Mech. Engrg.,32(1-3):199–259, 1981.

[82] T. J. R. Hughes. The Finite element method: linear static and dynamic finiteelement analysis. Dover, Mineola, 2000.

[83] O. C. Zienkiewicz and R. L. Taylor. The Finite Element Method: FluidDynamics. Butterworth-Heinemann, Philedelphia, 2005.

[84] J. Donea and A. Huerta. Finite Element Methods for flow problems. Wiley,2003.

[85] I. E. Vignon-Clementel, C. A. Figueroa, K. E. Jansen, and C. A. Taylor.Outflow boundary conditions for three-dimensional finite element modelingof blood flow and pressure in arteries. Comput. Meth. Appl. Mech. Engrg.,195(29-32):3776–3796, 2006.

[86] C. Audet and J. E. Dennis, Jr. Mesh adaptive direct search algorithms forconstrained optimization. SIAM Journal on Optimization, 17(1):2–11, 2006.

[87] G. Pelosi. The finite-element method, part i: R. l. courant [historical corner].Antennas and Propagation Magazine, IEEE, 49(2):180–182, 2007.

[88] E. Stein. Olgierd c. zienkiewicz, a pioneer in the development of the finiteelement method in engineering science. Steel Construction, 2(4):264–272,2009.

[89] G. Strang and G. Fix. An Analysis of The Finite Element Method. PrenticeHall, 1973.

[90] P. Lax. Feng kang. SIAM News, 26(11), 1993.

[91] K. Feng. Finite element method and natural boundary reduction. pages1439–1453, Warsaw, Poland, 1983. Proceedings of International Congress ofMathematicians.

[92] D-H Yu. Natural Boundary Integral Method and Its Applications. KluwerAcad Publ, Dordrecht, Netherlands, 2002.

[93] D. Givoli. Natural boundary integral method and its applications. Appl.Mech. Rev., 56(5):B65, 2003.

Page 197: Surgical design for the Fontan procedure using computational fluid

173

[94] C. A. Taylor, T. J. R. Hughes, and C. K. Zarins. Finite element modeling ofblood flow in arteries. Comput. Method. Appl. Mech. Engrg., 158(1-2):155–196, 1998.

[95] K. E. Jansen, C. H. Whiting, and G. M. Hulbert. A generalized-α method forintegrating the filtered navierstokes equations with a stabilized finite elementmethod. Comput. Meth. Appl. Mech. Engrg., 190(3-4):305–319, 2000.

[96] I. E. Vignon-Clementel, C. A. Figueroa, K. E. Jansen, and C. A. Taylor.Outflow boundary conditions for 3D simulations of non-periodic blood flowand pressure fields in deformable arteries. Comput. Meth. Biomech. Biomed.Engrg, 13(5):625–640, 2010.

[97] D. Givoli and J. B. Keller. A finite element method for large domains.Comput. Methods Appl. Mech. Engrg., 76(1):41–66, 1989.

[98] A. J. Booker, J. E. Dennis, Jr., P. D. Frank, D. B. Serafini, V. Torczon, andM. W. Trosset. A rigorous framework for optimization of expensive functionsby surrogates. Structural Optimization, 17(1):1–13, 1999.

[99] M. D. McKay, W. J. Conover, and R. J. Beckman. A comparison of threemethods for selecting values of input variables in the analysis of output froma computer code. Technometrics, 21:239–245, 1979.

[100] V. Torczon. On the convergence of pattern search algorithms. SIAM J.Optimization, 7:1–25, 1997.

[101] D. B. Serafini. A Framework for Managing Models in Nonlinear Optimiza-tion of Computationally Expensive Functions. PhD thesis, Rice University,Houston, TX, 1998.

[102] C. Audet and J. E. Dennis, Jr. Analysis of generalized pattern searches.SIAM Journal on Optimization, 13(3):889–903, 2003.

[103] C. Audet and J. E. Dennis, Jr. A pattern search filter method for non-linear programming without derivatives. SIAM Journal on Optimization,14(4):980–1010, 2004.

[104] D. G. Krige. A statistical approach to some mine valuations and allied prob-lems at the Witwatersrand. PhD thesis, University of Witwatersrand, Wit-watersrand, South Africa, 1951.

[105] G. Matheron. Principles of geostatistics. Economic Geology, 58:1246–1266,1963.

[106] G. Matheron. The intrinsic random functions and their applications. Adv.Appl. Prob., 5:439–468, 1973.

[107] J. Sacks, W. J. Welch, T. J Mitchell, and H. P. Wynn. Design and analysisof computer experiments. Statistical Science, 4(4):409–423, 1989.

[108] T. W. Simpson, J. J. Korte, T. M. Mauery, and F. Mistree. Comparison of re-sponse surface and kriging models for multidisciplinary design optimization.AIAA Paper 98-4755, 1998.

Page 198: Surgical design for the Fontan procedure using computational fluid

174

[109] D. R. Jones, M. Schonlau, and W. J. Welch. Efficient global optimization ofexpensive black box functions. Journal of Global Optimization, 13(4):445–492, 1998.

[110] W. Van Beers and J. P. C. Kleijnen. Kriging for interpolation in randomsimulation. Journal of the Operational Research Society, 54:255–262, 2003.

[111] J. P. C. Kleijnen. Kriging metamodeling in simulation: A review. EuropeanJournal of Operational Research, 192:707–716, 2009.

[112] S. N. Lophaven, H. B. Nielsen, and J. Søndergaard. DACE: A MATLABKriging toolbox version 2.0. Technical Report IMM-TR-2002-12, TechnicalUniversity of Denmark, Copenhagen, 2002.

[113] N. Wilson, K. Wang, R. Dutton, and C. A. Taylor. A software frameworkfor creating patient specific geometric models from medical imaging data forsimulation based medical planning of vascular surgery. Lect. Notes Comput.Sc., 2208:449–456, 2001.

[114] Jeanette P. Schmidt, Scott L. Delp, Michael A. Sherman, Charles A. Taylor,Vijay S. Pande, and Russ B. Altman. The simbios national center: Systemsbiology in motion. Proceedings of the IEEE,special issue on ComputationalSystem Biology., 96(8):1266 – 1280, 2008.

[115] C. H. Whiting and K. E. Jansen. A stabilized finite element method for theincompressible Navier-Stokes equations using a hierarchical basis. Interna-tional Journal for Numerical Methods in Fluids, 35(1):93–116, 2001.

[116] C. A. Figueroa, I. E. Vignon-Clementel, K. E. Jansen, T. J.R. Hughes, andC. A. Taylor. A coupled momentum method for modeling blood flow in three-dimensional deformable arteries. Computer Methods in Applied Mechanicsand Engineering, 195(41-43):5685–5706, 2006.

[117] J. P. Ku, M. T. Draney, F. R. Arko, W. A. Lee, F. Chan, N. J. Pelc, C. K.Zarins, and C. A. Taylor. In vivo validation of numerical predictions of bloodflow in arterial bypass grafts. Ann. Biomed. Eng., 30(6):743–752, 2002.

[118] J. P. Ku, C. J. Elkins, and C. A. Taylor. Comparison of CFD and MRI flowand velocities in an in vitro large artery bypass graft model. Ann. Biomed.Eng., 33(3):257 – 269, 2005.

[119] T. J. Kulik, J. L. Bass, B. P. Fuhrman, J. H. Moller, and J. E. Lock. Exerciseinduced pulmonary vasoconstriction. Brit. Heart J., 50:59–64, 1983.

[120] M. S. Shephard and M. K. Georges. Automatic three-dimensional meshgeneration by the finite octree technique. Int. J Numer. Meth. Eng., 32:709–749, 1991.

[121] M. A. Abramson, C. Audet, J. E. Dennis, Jr., , and S. Le Digabel. Or-thomads: A deterministic mads instance with orthogonal directions. SIAMJ. Optim., 20(2):948–966, 2009.

Page 199: Surgical design for the Fontan procedure using computational fluid

175

[122] R. Fletcher and S. Leyffer. Nonlinear programming without a penalty func-tion. Mathematical Programming, 91:239–269, 2002.

[123] T. Hsia, F. Migliavacca, S. Pittaccio, A. Radaelli, G. Dubini, G. Pennati,and M. de Leval. Computational fluid dynamic study of flow optimizationin realistic models of the total cavopulmonary connections. J. Surg. Res.,116(2):305–313, 2004.

[124] D. N. Rosenthal, A. H. Friedman, C. S. Kleinman, G. S. Kopf, L. E. Rosen-feld, and W. E. Hellenbrand. Thromboembolic complications after Fontanoperations. Circulation, 92(Suppl II):287–293, 1995.

[125] U. K. Chowdhury, B. Airan, S. S. Kothari, S. Talwar, A. Saxena, R. Singh,G. K. Subramaniam, K. K. Pradeep, C. D. Patel, and V. Venugopal. Specificissues after extracardiac Fontan operation: ventricular function, growth po-tential, arrhythmia, and thromboembolism. Ann. Thorac. Surg., 80(2):665–72, 2005.

[126] E. F. Grabowski. Thrombolysis, flow and vessel wall interactions. J VascInterv Radiol., 6:25s–29s, 1995.

[127] J. Loscalzo and A. I. Schafer. Thrombosis and hemorrhage. LippincottWilliams and Wilkins, Philedelphia, 2003.

[128] J. A. Lopez and J. Chen. Pathophysiology of venous thrombosis. ThrombosisResearch, 123(Suppl. 4):S30–S34, 2009.

[129] S. Wessler. Thrombosis in the presence of vascular stasis. American Journalof Medicine, 33:648–666, 1962.

[130] H. A. Walker and M. A. Gatzoulis. Prophylactic anticoagulation followingthe Fontan operation. Heart, 91(7):854–856, 2005.

[131] Y Ikeda, M Handa, K Kawano, T Kamata, M Murata, Y Araki, H Anbo,Y Kawai, K Watanabe, and I Itagaki. The role of von willebrand factor andfibrinogen in platelet aggregation under varying shear stress. J. Clin. Invest.,87(4):1234–1240, 1991.

[132] F. J. H. Gijsen, F. N. van de Vosse, and J. D. Janssen. The influence of thenon-newtonian properties of blood on the flow in large arteries: steady flowin a carotid bifurcation model. J. Biomechanics, 32:601–608, 1999.

[133] F.J.H. Gijsen, E. Allanic, F.N. van de Vosse, and J.D. Janssen. The influ-ence of the non-newtonian properties of blood on the fow in large arteries:unsteady flow in a 903 curved tube. J. Biomechanics, 32:705–713, 1999.

[134] S. Sankaran. Stochastic optimization using a sparse grid collocation scheme.Probab. Eng. Mech., 24(3):382–396, 2009.

[135] S. Sankaran C. Audet and A. L. Marsden. A method for stochastic con-strained optimization using derivative-free surrogate pattern search and col-location. J. Comput. Phys., 229(12):4664–4682, 2010.

Page 200: Surgical design for the Fontan procedure using computational fluid

176

[136] W. Yang, J. A. Feinstein, and A. L. Marsden. Constrained optimizationof an idealized Y-shaped baffle for the Fontan surgery at rest and exercise.Comput. Meth. Appl. Mech. Engrg., 199(33-36):2135–2149, 2010.

[137] T. Hsia, S. Khambadkone, A. N. Redington, F. Migliavacca, J. E. Deanfield,and M. R. de Leval. Effects of respiration and gravity on infradiaphragmaticvenous flow in normal and Fontan patients. Circulation, 102 suppl III:III–148–III–153, 2000.

[138] R. L. Spilker, J. A. Feinstein, D. W. Parker, V. M. Reddy, and C. A.Taylor. Morphometry-based impedance boundary conditions for patient-specific modeling of blood flow in pulmonary arteries. Annals Biomed. Eng.,35(4):546–549, 2007.

[139] S. C. Shadden and C. A. Taylor. Characterization of coherent structures inthe cardiovascular system. Annals of Biomedical Engineering, 36(7):1152–1162, 2008.

[140] M. A. Salim, T. G. DiSessa, K. L. Arheart, and B. S. Alpert. Contributionof superior vena caval flow to total cardiac output in children: a Dopplerechocardiographic study. Circulation, 92:1860–1865, 1995.

[141] C. P. Cheng, R. J. Herfkins, A. L. Lightner, C. A. Taylor, and J. A. Feinstein.Blood flow conditions in the proximal pulmonary arteries and vena cavae:healthy children during upright cycling exercise. Am J Physiol Heart CircPhysiol, 287(2):H921–926, 2004.

[142] S Sharma, A. E. Ensley, K. Hopkins, G. P. Chatzimavrodis, T. M. Healy,V. K. H. Tam, K. R. Kanter, and A. P. Yoganathan. In vivo flow dynam-ics of the total cavopulmonary connection from three-dimensional multislicemagnetic resonance imaging. Ann. Thorac. Surg., 71(3):889–898, 2001.

[143] J. Muller, O. Sahni, X. Li, K. E. Jansen, M. S. Shephard, and C. A. Taylor.Anisotropic adaptive finite element method for modeling blood flow. ComputMethods Biomech Biomed Eng., 8(5):295–305, 2005.

[144] A. I. J. Forrester, A. Sobester, and A. J. Keane. Multi-fidelity optimizationvia surrogate modelling. Proc. R. Soc. A, 463:3251–3269, 2007.

[145] P. B. Dobrin, P. Mirande, S. Kang, Q. S. Dong, and R. Mrkvicka. Me-chanics of end-to-end artery-to-PTFE graft anastomoses. Ann. Vasc. Surg.,12(4):317–23, 1998.

[146] K. R. Kanter, C. M. Haggerty, M. Restrepo, D. de Zelicourt, J. Rossignac,W. J. Parks, and A. P. Yoganathan. Preliminary clinical experience witha bifurcated y-graft fontan procedure. The Western Thoracic Surgical As-sociation 37th Annual Meeting , June 22-25, 2011; Colorado Springs, CO.,2011.

[147] E. O. Kung, A. S. Les, F. Medina, R. B. Wicker, McConnel M. V., andC. A. Taylor. In vitro validation of finite-element model of aaa hemodynam-

Page 201: Surgical design for the Fontan procedure using computational fluid

177

ics incorporating realistic outlet boundary conditions. J. Biomech. Eng.,133(4):041103, 2011.

[148] F. Loth, P. F. Fischer, and H. S. Bassiouny. Blood flow in end-to-side anas-tomoses. Annu. Rev. Fluid Mech., 40:367–393, 2008.

[149] Z. Yin, H. Wang, Z. Wang, H. Zhu, R. Zhang, M. Hou, and M. Fang. Ra-dionuclide and angiographic assessment of pulmonary perfusion after fontanprocedure: Comparative interim outcomes. Ann. Thorac. Surg., 93:620–625,2012.

[150] C. M. Haggerty, K. R. Kanter, M. Restrepo, D. A. de Zelicourt, W. J. Parks,J. Rossignac, M. A. Fogel, and A. P. Yoganathan. Simulating hemodynamicsof the fontan y-graft based on patient-specific in vivo connections. J. Thorac.Cardiovasc. Surg., 2012. 10.1016/j.jtcvs.2012.03.076.

[151] W. Yang, J. A. Feinstein, S. C. Shadden, I. E. Vignon-Clementel, and A. L.Marsden. Optimization of a Y-graft design for improved hepatic flow distri-bution in the Fontan circulation. J. Biomech. Engrg., 2012. in press.

[152] S. Sankaran, M. Esmaily Moghadam, A.M. Kahn, J. Guccione, E. Tseng,and A.L. Marsden. Patient-specific multiscale modeling of blood flow forcoronary artery bypass graft surgery. Ann Biomed Eng, 40(1):2228–2242,2012.