You are on page 1of 7

ARTICLE IN PRESS

Journal of Biomechanics 40 (2007) 2716–2722


www.elsevier.com/locate/jbiomech
www.JBiomech.com

Numerical simulation of saccular aneurysm hemodynamics:


Influence of morphology on rupture risk
Brent Utter, Jenn Stroud Rossmann
Mechanical Engineering Department, Lafayette College, Easton, PA 18042, USA
Accepted 11 January 2007

Abstract

The governing equations for pulsatile fluid flow were solved in their finite volume formulation in order to simulate blood flow in a
variety of three-dimensional aneurysm geometries. The influence of geometric factors on flow patterns and fluid mechanical forces was
studied with the goal of identifying the risk of aneurysm rupture. Aneurysm morphology was characterized by quantitative shape indices
reflecting the three dimensionality of the vasculature derived from clinical studies. Recirculation zones and secondary flows were
observed in aneurysms and arteries. Regions of extreme and alternating shear stress were observed and identified as sites for potential
aneurysm rupture. The ellipticity of an aneurysm was observed to be strongly correlated with wall shear stress at the aneurysm fundus,
while its non-sphericity, volume, and degree of undulation were more weakly correlated.
r 2007 Elsevier Ltd. All rights reserved.

Keywords: Numerical simulation; Aneurysm; Intracranial; Morphology; Shape index; Rupture

1. Introduction stood to have the greatest influence on aneurysm growth


and rupture (Liepsch, 2001; Hoi et al., 2004). Experiments
Aneurysms pose a grave health risk. If left untreated, such as those of Gobin et al. (1994) and observations
aneurysms may continue to expand until rupture, causing including those of Strother (1995) have demonstrated the
hemorrhage, complications of local organ function, and existence of a recirculation zone inside a saccular aneur-
possibly death. Aneurysm rupture has high rates of mortality ysm. This region of slow flow is believed to promote the
and morbidity (e.g. Kassell et al., 1990). Of particular interest formation of thrombi or blood clots (Hademenos and
due to their location are intracranial aneurysms (ICAs), Massoud, 1998), as recirculating flow also does in
commonly found at bifurcations in the circle of Willis. Since otherwise healthy arteries. Foutrakis et al. (1999) con-
ICAs are asymptomatic prior to rupture (Vega et al., 2002), cluded that in the absence of disease, the formation and
intact ICAs are often discovered by neurosurgeons who are growth of saccular aneurysms are due to mechanical forces.
not looking for them. Once an ICA is found, the choice of The search for reliable rupture predictors that can be
treatment is complicated by the limited diagnostic criteria to obtained from available in vivo information has focused
predict rupture risk and by the risks of surgical intervention primarily on aneurysm size, aspect ratio, and irregularity.
itself. A detailed understanding of the local hemodynamics, Hademenos et al. (1994) improved the traditional Laplace’s
of the effects of vascular wall remodeling on the flow law method of determining an aneurysm’s ‘‘critical radius’’
patterns, and of long-term adaptation to surgical interven- by accounting for the wall’s viscoelasticity, and determined
tion, is vital to useful clinical treatment of aneurysms. a radius of 4.8 mm above which rupture was likely.
While wall structure and behavior are relevant to any However, because many large ICAs remain intact, while
investigation of aneurysmal vessels, fluid flow is under- some small ICAs rupture (e.g. Beck et al., 2003), size is not
a reliable predictor of rupture risk.
Corresponding author. Tel.: +1 610 330 5406; fax: +1 610 330 5059. An aneurysm’s aspect ratio (AR) has also been suggested
E-mail address: rossmanj@lafayette.edu (J.S. Rossmann). as a rupture predictor (e.g. Parlea et al., 1999; Ujiie et al.,

0021-9290/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jbiomech.2007.01.011
ARTICLE IN PRESS
B. Utter, J.S. Rossmann / Journal of Biomechanics 40 (2007) 2716–2722 2717

1999; Nader-Sepahi et al., 2004). AR is the ratio of the


aneurysm’s height to its neck diameter. Nader-Sepahi et al.
(2004) measured the aspect ratio of 75 ruptured and 107
intact aneurysms and found the average AR values of each
group to be 2.70 and 1.80, respectively. However, the AR
of a physiological aneurysm is highly dependent on the
viewing angle or projection plane imaged; for multilobular
and other complex aneurysm geometries, a definitive value
of AR is difficult or impossible to obtain.
Rohde et al. (2005) used Fourier analysis to quantify the
irregularity of aneurysms. Although this method can only
be applied to a two-dimensional projection, the results of
this study suggested that irregular aneurysms were more
likely to rupture.
Ma et al. (2004) developed shape indices based on three-
dimensional morphologies, and calculated these indices for
clinically observed aneurysms. One year later, the authors
evaluated their indices by revisiting the same patient base
(Raghavan et al., 2005). Although the population consisted
of only 27 patients, the results suggested that a more Fig. 1. Base geometry for all simulations.
irregular geometry had a higher chance of rupture.
Fourteen individual aneurysm geometries are examined in this study.
Significantly, the authors found that ‘‘abnormal shape
Of these, 13 are derived from a 7 mm sphere. One 10 mm spherical
characteristics’’ were possible in aneurysms of all sizes. aneurysm is studied; this diameter just exceeds the ‘‘critical’’ value
Their results also indicated that the indices for non- predicted by Hademenos et al. (1994) and would be most likely to receive
sphericity and ellipticity were among the most statistically surgical intervention due to its size. The remaining aneurysms are not
significant indices. smooth spheres, but contain undulations, lobed and elliptical features
modeled on clinical images (e.g. Parlea et al., 1999; Shojima et al., 2004).
Shojima et al. (2004) obtained the geometries of 20
The degree of geometric irregularity can then be quantified using the shape
saccular ICAs using CT angiography, and calculated the indices of Ma et al. (2004), and the blood flow simulated.
local wall shear stress (WSS) acting at the fundi to be
significantly lower than the WSS acting on the surrounding
2.2. Shape indices
vasculature. The authors surmised that low values of WSS
on the aneurysmal wall promote the growth and eventual The indices used to evaluate aneurysms in this study depend on the
rupture, via the degeneration of endothelial cells, of volume and surface area of the aneurysm and of its convex hull (Ma et al.,
saccular ICAs. Also, in a result corroborating the conclu- 2004). The convex hull of an object is the smallest volume that completely
sions of Nader-Sepahi et al. (2004), Shojima et al. (2004) envelops the object and is also completely convex. The volume and surface
found an inverse relationship between an aneurysm’s aspect area of the convex hull for each aneurysm studied are computed using the
Qhull1 algorithm.
ratio and the shear stress acting on the aneurysmal wall. The non-sphericity index (NSI) is defined as
The goal of the current study is to correlate aneurysm
geometry, characterized by the shape indices of Ma et al. V 2=3
NSI ¼ 1  ð18pÞ1=3 ,
(2004), with the intra-aneurysmal fluid dynamics obtained S
using computational fluid dynamics (CFD). Such a where V and S are the volume and surface area of the aneurysm,
respectively. NSI increases as the aneurysm’s geometry deviates from the
correlation could inform the diagnosis and treatment of
shape of a sphere.
an intact aneurysm by elucidating the likelihood of The ellipticity index (EI) is defined as
rupture.
2=3
V CH
EI ¼ 1  ð18pÞ1=3 ,
S CH
2. Methods where VCH and SCH are the volume and surface area of the aneurysm’s
convex hull. EI is closely related to the aspect ratio, because the EI will be
2.1. Geometry of models large when the height of the aneurysm is large compared to the aneurysm’s
width. EI differs from AR in that EI will also be large if the aneurysm’s
This study is focused on aneurysms forming at arterial bifurcations in body is wide compared to its height.
the circle of Willis, and in particular, at the basilar artery bifurcation. The undulation index is defined as
Ingebrigtsen et al. (2004) measured the geometric features of 107  
V
intracranial artery bifurcations. For the basilar artery, they reported an UI ¼ 1  .
V CH
average parent diameter of 3.2 mm, one daughter vessel with diameter
2.0 mm and a 631 branching angle, and a second daughter vessel with
1
diameter 1.8 mm and 651 branch angle. These values are used to create the Qhull was developed by The National Science and Technology
surrounding vasculature of the three-dimensional aneurysms in the current Research Center for Computation and Visualization of Geometric
study, shown in Fig. 1. Structures (The Geometry Center), University of Minnesota.
ARTICLE IN PRESS
2718 B. Utter, J.S. Rossmann / Journal of Biomechanics 40 (2007) 2716–2722

UI characterizes the ‘‘lumpiness’’ of an aneurysm, and increases as the A representative volumetric flowrate in the basilar artery was obtained
aneurysm wall becomes more undulatory. using phase-contrast cine magnetic resonance by Marks et al. (1992). The
resulting velocity waveform used in the current study is shown in Fig. 2.
2.3. Computational method Four cycles are simulated to ensure periodic flow conditions, and the
results of the fourth are reported. Flow-through conditions are applied at
the outlets of the daughter vessels. At the walls, which are assumed to be
For the simulations carried out in this study, it is assumed that blood is
rigid, no-slip and no-penetration conditions are imposed. At the basilar
an incompressible, Newtonian fluid and that the flow is laminar and
artery inlet, the velocity profile is assumed to be uniform with the pulsatile
isothermal. The assumption of Newtonian behavior is based on the
waveform as shown in Fig. 2.
findings of Perktold et al. (1989) and others who found minimal changes in
The effect of the inlet velocity profile on the downstream flow was
arterial flow patterns when non-Newtonian effects were included. For such
examined, and the impact of the uniform flow assumption was found to be
a fluid, the continuity and Navier–Stokes equations are as follows:
minimal. As shown in Fig. 3, the velocity profile becomes nearly parabolic
r  u ¼ 0, upstream of the bifurcation and aneurysm region whether the inlet flow is
  uniform or parabolic, as expressed by the Womersley solution (Zamir,
qu
r þ u r u ¼  rp þ mr2 u , 2000). This is physiologically realistic: although the basilar artery receives
qt
blood from two vertebral arteries, it is sufficiently long to allow adequate
where u is the velocity, r the fluid density, p the pressure, and m the fluid mixing of the streams to achieve a nearly parabolic, pulsatile velocity
viscosity. Solution of these equations in their finite volume form is profile (e.g. Valencia et al., 2006).
accomplished through a commercial software package, CFD–ACE+ As the model was developed, grid refinement studies were performed to
(Singhal, 1998). The discretization is temporally implicit and spatially ensure an optimal balance between resolution and computational expense.
second order. Velocity–pressure coupling is achieved using the SIMPLEC These studies demonstrated that unstructured grids consisting of
algorithm initially suggested by Chorin (1967). 150,000–250,000 cells, depending on the morphology, adequately captured
the details of the flow without requiring unreasonable processing time.

3. Results

3.1. Flow patterns and forces

Fourteen different aneurysm geometries were created


and analyzed in the current study. Most were derived from
a 7 mm sphere. The calculated systolic WSS distribution,
pressure distribution and streamlines for the 7 mm sphe-
rical aneurysm are shown in Fig. 4.
The WSS and pressure distribution images shown in
Fig. 4 are rotated 401 about the vertical axis to show the
shoulders of the aneurysm more clearly. The peak values
calculated at the shoulders are a WSS of 14 N/m2 (140 dyn/
cm2) and a pressure of 515 N/m2 (3.86 mm Hg); these may
be compared with values in the surrounding vasculature
that vary from 2 to 5 dyn/cm2 and 0.5–2.9 mmHg. The
Fig. 2. Mean velocity of blood in the basilar artery (after Marks et al., WSS throughout most of the aneurysm is less than 1 N/m2
1992). (10 dyn/cm2) at systole, with the lowest values observed at

Fig. 3. Effect of inlet velocity profile on flow in healthy bifurcation: left, inlet profile is uniform; right, parabolic.
ARTICLE IN PRESS
B. Utter, J.S. Rossmann / Journal of Biomechanics 40 (2007) 2716–2722 2719

Fig. 4. WSS distribution, pressure distribution, and streamlines of intra-aneurysmal flow at systole for 7 mm spherical model. Streamlines are colored by
their distance from the basilar artery midplane.

Fig. 5. Sampling of aneurysm shapes studied: (a) high aspect ratio, (b) distorted 7 mm sphere, and (c) small irregular geometries.

the fundus. These results are in agreement with the findings aneurysmal flow is not decelerated to the same extent as
of other studies of aneurysmal flow, including those of it is in larger models. Despite the increased average flow
Valencia et al. (2006), who also observed low WSS at the and small size of this aneurysm, the average systolic WSS
aneurysm fundus. for this model differs from that in the 7 mm spherical
The instantaneous streamlines in Fig. 4 are colored by model by only 1 %.
their distance from the midplane. The streamlines, which Table 1 shows the spatially averaged systolic wall shear
begin in the basilar artery’s plane of symmetry, extend out stress and pressure, as well as the time-averaged WSS near
of plane by as much as 75% of the aneurysm radius. They the fundus, for several of the geometries examined.
demonstrate the three dimensionality of intra-aneurysmal
flow; recirculation zones and out-of-plane motion are 3.2. Shape indices
present that would not be captured by a two-dimensional
model. Over the 14 aneurysm models studied, the aneurysm
Similar regions of elevated WSS and pressure exist in volume varied from 43 to 517 mm3, the EI from 0.209 to
each model. The high aspect ratio model shown in Fig. 5(a) 0.273, UI from 0.002 to 0.1, and NSI from 0.223 to 0.339.
differs from the spherical models in that a higher average For analysis, all index values are normalized by the largest
pressure acts on the high aspect ratio aneurysm both at value in the set.
systole (7.3% higher than the 7 mm model) and diastole Fig. 6 shows that in general, as the EI increases, the
(6.14% higher). The high aspect ratio model has the highest average WSS acting on the aneurysmal wall decreases.
values of average pressure on the aneurysmal wall at both In both the systolic and diastolic series, the lowest EI
systole (545 N/m2, or 4.09 mmHg) and diastole (242 N/m2, represents the 10 mm geometry, whose volume, more
or 1.82 mmHg). than four times greater than the average volume of the
The aneurysm in Fig. 5(c) has the smallest volume of any thirteen other geometries, is its dominant morphological
aneurysm in this study. Due to its size, the intra- characteristic.
ARTICLE IN PRESS
2720 B. Utter, J.S. Rossmann / Journal of Biomechanics 40 (2007) 2716–2722

Table 1
Spatially averaged aneurysm WSS and pressure at systole, and temporally
averaged WSS at fundus, for several aneurysm models

Aneurysm Systolic Over cycle

Averaged WSS Averaged P Fundus WSS


(N/m2), dyn/cm2 (N/m2), mm Hg (N/m2), dyn/cm2

7 mm sphere 0.7199, 7.2 508, 3.81 0.0523, 0.52


Distorted 1 0.7688, 7.7 540, 4.05 0.0456, 0.46
Distorted 2 0.6722, 6.7 511, 3.83 0.0456, 0.46
Distorted 3 0.7231, 7.2 509, 3.82 0.0345, 0.35
Distorted 4 0.5928, 5.9 510, 3.83 0.0055, 0.05
10 mm sphere 0.2667, 2.7 507, 3.80 0.0172, 0.17
High AR 0.7725, 7.7 545, 4.09 0.0138, 0.14
Small 0.7702, 7.7 524, 3.93 0.0320, 0.32
irregular
Large 0.5022, 5.0 503, 3.77 0.0013, 0.01
irregular

Fig. 7. WSS near fundus versus time over one cardiac cycle. Systole
occurs 0.2 s into cycle (3.2 s in figure); diastolic or minimum inlet flow
conditions are at 0.75 s (here, 3.75 s).

Fig. 6. Average WSS of aneurysm at systole and diastole versus


normalized EI. Fig. 8. Variation in, and average, WSS at fundus over cycle versus EI.

The WSS at the fundus of each aneurysm model over normalized quantity is referred to as the ‘‘normalized
one cardiac cycle is shown in Fig. 7. The cyclic nature of change in WSS.’’ The WSS at the fundus, averaged over
the WSS is observable, with stress following the inlet the cycle, is also shown.
velocity waveform (systolic 0.2 s into cycle, diastolic at The correlation of normalized change in WSS with EI
0.75 s) with varying closeness. The low-flow conditions in (shown in Fig. 8) is most strongly linear with an r value of
the aneurysm appear to delay both systole and diastole in 0.924, while the linear correlations of Figs. 9 (NSI) and 10
the aneurysm dome. (UI) have less compelling r values of 0.704 and 0.474,
The WSS at the fundus is significantly lower than the respectively. The relative significance of these relationships
spatially averaged WSS throughout the cycle. As Fig. 7 is also demonstrated by p values of 2.29  106 (EI), 0.0049
shows, the fundus WSS values calculated in the larger (NSI), and 0.087 (UI), suggesting a strong relationship of
(10 mm diameter), more irregular (independent of aneurysm normalized change in WSS with EI, a significant but less
volume), and higher AR aneurysms are consistently lower strong one with NSI, and a weak one with UI.
than those in the 7 mm sphere and its undulatory variants.
To illustrate the magnitude of the alternation in WSS at 4. Discussion
the fundus of each aneurysm, the maximum ‘‘peak-to-
peak’’ change in WSS at the fundus is normalized by the Of the shape indices investigated, the EI has the strongest
average fundus WSS over the cycle. In Figs. 8–10, this correlation with wall shear stress in the aneurysm. The EI
ARTICLE IN PRESS
B. Utter, J.S. Rossmann / Journal of Biomechanics 40 (2007) 2716–2722 2721

compromised supply of nitric oxide usually released in


response to high WSS, result in the aggregation of red
blood cells, and in the buildup and adhesion of platelets
and leukocytes on the intimal surface. Subsequent effects
can include intimal damage, small thrombus formation,
and infiltration of white blood cells and fibrin inside the
aneurysm wall (Ujiie et al, 1999). The biological response
triggered by this low WSS may ultimately cause the
endothelial cells to degenerate via apoptosis. Pentimalli
et al. (2004) found higher levels of apoptosis in ruptured
than intact aneurysms, and also found a strong link
between apoptosis and aneurysm irregularity.
Oscillatory WSS at the neck’s distal end has previously
been suggested to promote aneurysm growth by weakening
the intima (Gonzalez et al., 1992). Fig. 7 illustrates the
cyclic nature of WSS acting on the fundus of the aneurysm.
Fig. 9. Variation in, and average, WSS at fundus over cycle versus NSI.
Figs. 9 and 10 show that aneurysms with increasing levels
of irregularity have increasing normalized differences of
WSS at their fundi. This unstable near-wall environment
may signal an increased risk of rupture.
Tension in the wall of the aneurysm may also lead to
rupture, especially if the wall is thinning via cell degenera-
tion. It is imperative for future work to consider the
interaction between the fluid flow and the arterial wall, as
the effects of vessel remodeling and the material properties
of the arterial wall (e.g. Toth et al., 1998) surely play an
active role in aneurysm growth and rupture. It is
impossible to say anything conclusive about mechanisms
for rupture, including the effects of cyclic loading or
fatigue, without considering the interactions of fluid and
solid mechanics. Humphrey and Canham (2000) provide
an elegant review urging elasticians to consider this
problem.
The current work indicates that aneurysm ellipticity (EI)
and, to a lesser extent, non-sphericity (NSI) are correlated
Fig. 10. Variation in, and average, WSS at fundus over cycle versus UI. with fluid mechanical forces and rupture risk. No definitive
correlation was observed between WSS and aneurysm
irregularity as quantified by the undulation index. How-
considers three dimensionality and aneurysm orientations ever, the full range of values of the shape indices has not
not accounted for by the aneurysm’s aspect ratio. As EI yet been explored. The current study is limited by the small
increases, the wall shear stress at the fundus decreases, population of aneurysms, and also by the physiological
suggesting an increasing risk of rupture with increasing accuracy of the aneurysm models, which are based on
aneurysm ellipticity. geometries obtained in clinical studies (e.g. Parlea et al.,
Fig. 8 shows that as EI increases, the peak-to-peak 1999; Shojima et al., 2004). Future studies will simulate
change in WSS at the fundus increases, and the time- flow through actual vasculature geometries obtained from
averaged wall shear stress at the fundus decreases. The CT angiography and MRI; the availability of clinical data
normalized change in WSS at the fundus is an important about the progression of these aneurysms will usefully
parameter, because significant oscillation of WSS at the supplement CFD analyses. Further study will also include
fundus may trigger a biological response from the consideration of the fluid particle residence times within the
endothelial cells lining the wall of the aneurysm. aneurysm, especially near the fundus, and the implications
The seemingly counterintuitive observation that aneur- for risk of rupture.
ysms commonly rupture at their fundi, in a response to low
WSS, can be understood when biological and mechanical
factors are considered together. Acknowledgments
Shojima et al. (2004) suggested that low WSS acting at
the fundus of saccular aneurysms was a direct cause of This work has been supported by the Lafayette
aneurysm rupture. These low-flow conditions, and the College EXCEL Scholars program and by NSF Grant
ARTICLE IN PRESS
2722 B. Utter, J.S. Rossmann / Journal of Biomechanics 40 (2007) 2716–2722

DBI-0442269, which provided the Linux cluster used for que—evaluation of normal subjects and patients with arteriovenous-
computations. malformations. Radiology 182 (2), 467–476.
Nader-Sepahi, A., Casimiro, M., Sen, J., Kirchen, N., 2004. Is aspect ratio
a reliable predictor of intracranial aneurysm rupture? Neurosurgery 54
References (6), 1343–1347.
Parlea, L., Fahrig, R., Holdsworth, D.W., Lownie, S.P., 1999. An analysis
Beck, J., Rohde, S., el Beltagy, M., Zimmerman, M., Berkefeld, J., Seifert, of the geometry of saccular intracranial aneurysms. American Journal
V., Raabe, A., 2003. Difference in configuration of ruptured and of Neuroradiology 20, 1079–1089.
unruptured intracranial aneurysms determined by biplanar digital Pentimalli, L., Modesti, A., Vignati, A., Marchese, E., Albanese, A., Di
subtraction angiography. Acta Neurochir 145, 861–865. Rocco, F., Coletti, A., Di Nardo, P., 2004. Role of apoptosis in
Chorin, A., 1967. A numerical method for solving incompressible viscous intracranial aneurysm rupture. Journal of Neurosurgery 101 (6),
flow problems. Journal of Computational Physics 2, 12–26. 1018–1025.
Foutrakis, G., Yonas, H., Sclabassi, R., 1999. Saccular aneurysm Perktold, K., Peter, R., Resch, M., 1989. Pulsatile non-Newtonian blood
formation in curved and bifurcating arteries. American Journal of flow simulation through a bifurcation with an aneurysm. Biorheology
Neuroradiology 20 (7), 1309–1317. 26 (6), 1011–1030.
Gobin, Y., Counord, J., Flaud, P., Duffaux, J., 1994. In vitro study of Raghavan, M., Ma, B., Harbaugh, R., 2005. Quantified aneurysm shape
hemodynamics in a giant saccular aneurysm model: influence of flow and rupture risk. Journal of Neurosurgery 102 (2), 355–362.
dynamics in the parent vessel and effects of coil embolization. Rohde, S., Lahmann, K., Beck, J., Nafe, R., Yan, B., Raabe, A.,
Neuroradiology 36, 530–536. Berkefeld, J., 2005. Fourier analysis of intracranial aneurysms:
Gonzalez, C., Cho, Y., Ortega, H., Moret, J., 1992. Intracranial towards an objective and quantitative evaluation of the shape of
aneurysms: flow analysis of their origin and progression. American aneurysms. Neuroradiology 47 (2), 121–126.
Journal of Neuroradiology 13 (1), 181–188. Shojima, M., Oshima, M., Takagi, K., Torii, R., 2004. Magnitude and role
Hademenos, G., Massoud, T., 1998. The Physics of Cerebrovascular of wall shear stress on cerebral aneurysm—computational fluid
Diseases. Springer, New York. dynamic study of 20 middle cerebral artery aneurysms. Stroke 35
Hademenos, G., Massoud, T., Valentino, D., Dickwiler, G., Vinuela, F., (11), 2500–2505.
1994. A nonlinear mathematical model for the development and Singhal, A.K., 1998. Key elements of verification and Validation of CFD
rupture of intracranial saccular aneurysms. Neurological Research 16 Software. In: Proceedings of the 29th AIAA Fluid Dynamics
(5), 376–384. Conference, AIAA 98-2639.
Hoi, Y., Meng, H., Woodward, S., Guterman, ., Bendok, B., Hanel, R., Strother, C.M., 1995. In vitro study of hemodynamics in a giant saccular
Hopkins, L., 2004. Effects of arterial geometry on aneurysm growth: aneurysm model: Influence of flow dynamics in the parent vessel and
three-dimensional computational fluid dynamics study. Journal of effects of coil embolization (letter; comment). Neuroradiology 37,
Neurosurgery 101 (4), 676–681. 159–161.
Humphrey, J., Canham, P., 2000. Structure, mechanical properties, and Toth, M., Nadasy, G.L., Nyary, I., Kerenyi, T., Orosz, M., Molnarka, G.,
mechanics of intracranial saccular aneurysms. Journal of Elasticity 61, Monos, E., 1998. Sterically inhomogeneous viscoelastic behavior of
49–81. human saccular cerebral aneurysms. Journal of Vascular Research 35,
Ingebrigtsen, T., Morgan, M., Faulder, K., Ingebrigsten, L., Sparr, T., 345–355.
Schirmer, H., 2004. Bifurcation geometry and the presence of cerebral Ujiie, H., Tachibana, H., Hiramatsu, O., Hazel, A., Matsumoto, T.,
artery aneurysms. Journal of Neurosurgery 101 (1), 108–113. Ogasawara, Y., Nakajima, H., Hori, T., Takakura, K., Kajiya, F., 1999.
Kassell, N.F., Torner, J.C., Haley, E.C., et al., 1990. the international Effects of size and shape (aspect ratio) on the hemodynamics of saccular
cooperative study on the timing of aneurysm surgery. 1. Overall aneurysms: a possible index for surgical treatment of intracranial
management results. Journal of Neurosurgery 73 (1), 18–36. aneurysms. Neurosurgery 45 (1), 119–129 (discussion 129–30).
Liepsch, D., 2001. An introduction to biofluid mechanics—basic models Valencia, A.A., Guzman, A.M., Finol, E.A., Amon, C.H., 2006. Blood
and applications. Journal of Biomechanics 35, 415–435. flow dynamics in saccular aneurysm models of the basilar artery.
Ma, B., Harbaugh, R., Raghavan, M., 2004. Three-dimensional geome- Journal of Biomechanical Engineering (128), 516–526.
trical characterization of cerebral aneurysms. Annals of Biomedical Vega, C., Kwoon, J.V., Lavine, S.D., 2002. Intracranial aneurysms:
Engineering 32 (2), 264–273. current evidence and clinical practice. American Family Physician 66
Marks, M., Pelc, N., Ross, M., Elzmann, D., 1992. Determination of (4), 601–608.
cerebral blood-flow with a phase-contrast cine MR imaging techni- Zamir, M., 2000. The Physics of Pulsatile Flow. Springer, New York.

You might also like