Next Article in Journal
Data-Driven Insights into Labor Progression with Gaussian Processes
Previous Article in Journal
U-Net Convolutional Neural Network for Real-Time Prediction of the Number of Cultured Corneal Endothelial Cells for Cellular Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multi-Dimensional Modeling of Cerebral Hemodynamics: A Systematic Review

by
Jana Korte
1,2,*,
Ehlar Sophie Klopp
1,3 and
Philipp Berg
1,3
1
Research Campus STIMULATE, University of Magdeburg, 39106 Magdeburg, Germany
2
Department of Fluid Dynamics and Technical Flows, University of Magdeburg, 39106 Magdeburg, Germany
3
Department of Medical Engineering, University of Magdeburg, 39106 Magdeburg, Germany
*
Author to whom correspondence should be addressed.
Bioengineering 2024, 11(1), 72; https://doi.org/10.3390/bioengineering11010072
Submission received: 11 December 2023 / Revised: 23 December 2023 / Accepted: 6 January 2024 / Published: 11 January 2024
(This article belongs to the Section Regenerative Engineering)

Abstract

:
The Circle of Willis (CoW) describes the arterial system in the human brain enabling the neurovascular blood supply. Neurovascular diseases like intracranial aneurysms (IAs) can occur within the CoW and carry the risk of rupture, which can lead to subarachnoid hemorrhage. The assessment of hemodynamic information in these pathologies is crucial for their understanding regarding detection, diagnosis and treatment. Multi-dimensional in silico approaches exist to evaluate these hemodynamics based on patient-specific input data. The approaches comprise low-scale (zero-dimensional, one-dimensional) and high-scale (three-dimensional) models as well as multi-scale coupled models. The input data can be derived from medical imaging, numerical models, literature-based assumptions or from measurements within healthy subjects. Thus, the most realistic description of neurovascular hemodynamics is still controversial. Within this systematic review, first, the models of the three scales (0D, 1D, 3D) and second, the multi-scale models, which are coupled versions of the three scales, were discussed. Current best practices in describing neurovascular hemodynamics most realistically and their clinical applicablility were elucidated. The performance of 3D simulation entails high computational expenses, which could be reduced by analyzing solely the region of interest in detail. Medical imaging to establish patient-specific boundary conditions is usually rare, and thus, lower dimensional models provide a realistic mimicking of the surrounding hemodynamics. Multi-scale coupling, however, is computationally expensive as well, especially when taking all dimensions into account. In conclusion, the 0D–1D–3D multi-scale approach provides the most realistic outcome; nevertheless, it is least applicable. A 1D–3D multi-scale model can be considered regarding a beneficial trade-off between realistic results and applicable performance.

1. Introduction

The circle of Willis (CoW) is the main neurovascular artery system supplying the brain with oxygen and nutrients. The CoW has three main supplying arteries connecting the anterior and posterior blood circulation: the left and right anterior internal carotid arteries and the posterior basilar artery [1]. These ascend from the aortic arch and combine from the common carotid arteries’ anterior and the left and right posterior vertebral arteries. The anatomy of the aorta and the CoW is subject-specific with differences in bifurcations and artery connections. This should be considered when examining the patient-specific blood flow in the CoW, which is crucial for the study of neurovascular diseases. These diseases comprise, among others, stenoses of the vessel and intracranial aneurysms (IAs). IAs are dilations of the neurovasculature, which carry the risk of rupture. IA rupture can lead to subarachnoid hemorrhage with fatal consequences like stroke or death of the patient [2]. Knowledge about the hemodynamics in IAs can provide information about the pathology’s growth and help understand the rupture risk [3,4]. Understanding the pathology is important for clinical applications like early-stage detection and therapy planning, which in the following can reduce fatal consequences. Therapies nowadays mostly comprise minimally invasive endovascular treatment like stenting or coiling [5]. The important hemodynamic information can be assessed non-invasively through medical imaging like magnetic resonance imaging (MRI) or echocardiography [6]. Though these techniques can give realistic in vivo information about the blood flow, high restriction regarding the spatial and temporal resolution exist and no wall-related parameters can be extracted [7]. The use of non-invasive methods, such as in silico numerical modeling, is required here, as the mentioned disadvantages can be overcome by the numerical evaluation of the realistic flow field on the basis of patient-specific data. Different modeling approaches exist to investigate the blood flow through the CoW and the brain-supplying arteries [8,9]. These have a dimensional (D) range from 0D, solving ordinary differential equations (ODEs), to 1D, solving partial differential equations (PDEs), to 3D numerical simulations, solving PDEs in full volumetric space. The numerical analysis of neurovascular pathologies was studied already in 1988 with 0D computer simulations based on animal trials [10] and in 1991 with 1D computer simulations [11]. In recent studies, these low-scale models are often used to supply boundary conditions for 3D approaches. Lower-dimensional simulations provide hemodynamic information on lower cost and real-time periods and thus can serve as boundary conditions for higher-dimensional simulations [12].
Three-dimensional simulations provide the most highly resolved insight into hemodynamics but nevertheless are expensive and time-consuming [13,14]. Therefore, mostly only a small part of the vessel containing the pathology is analyzed [15]. Moreover, since patient-specific real-time flow data are difficult to capture and are therefore rarely available, the surrounding information, which is used as boundary conditions, is often based on assumptions or taken from the literature [16]. In addition to receiving the boundary conditions from low-scale models, coupled multi-scale approaches exist, directly exchanging flow and pressure information [17,18]. Multi-scale coupling carries the advantage of incorporating and exchanging information over the dimensions and models, leading to a more realistic simulation of the patient-specific flow [19]. Challenges regarding the multi-scale analysis of hemodynamics arise considering the individual structure of the CoW anatomy, the complex morphology of the pathologies and the patient-specific as well as posture- and motion-specific pulsatile flow. Moreover, modeling challenges regarding the interaction between different dimensions are a crucial factor in the development of the most realistic blood flow reproduction of neurovascular pathologies. Subsequently, regardless of the wide-ranging development, validation and application of numerical low-, high- and multi-scale models within the past years, up to now no decision on the most realistic and applicable model could be derived. Clinical applicability relies on this decision, which needs further investigation to provide significant results. This paper aims to review the existing approaches and numerical models, focusing on multi-scale modeling of the CoW and the supplying arteries, namely the inflow into the brain circulation. Recent research will be discussed and described accordingly. First, the existing models, comprising 0D, 1D and 3D models, will be discussed. Second, the review of the multi-scale coupling of these models, looking at the 0D–1D, 0D–3D, 1D–3D and 0D–1D–3D coupling, will be performed. The most important findings according to a realistic reproduction of the blood flow and the applicability of the respective approaches will be presented. In the following, the crucial weaknesses of the models regarding these factors will be discussed and the required future steps will be extracted.

2. Methods

The literature search was carried out using the keywords “circle of willis hemodynamic blood flow simulation” in the databases PubMed and ScienceDirect. Exclusion of articles was structured based on Atkinson et al. and Moher et al. [20,21].
As described in Figure 1, articles were taken into account within the database search and further extended by a backward (screening of reference list) and forward search (screening of citations) as well as a direct contact search (communication to individual researchers) [21]. After duplicates were removed, this search returned 564 results in total for a first screening. This included a filtering based on title and abstract to determine whether articles could contribute to this work. As such, articles in which the neurovascular system was not mentioned or solely an experimental approach was performed were excluded. In the next step, the full texts of the remaining articles were read. If it became clear while reading that the articles did not contribute to the method development of low-, high- or multi-scale neurovascular hemodynamic analysis, these were also excluded. In total, 115 articles were analyzed in this review, comprising different model dimensions and different multi-scale models. The studies were firstly categorized by the developed methods based on their model dimensions in three categories: 0D, 1D and 3D models. Secondly, the multi-scale models were categorized in four categories: 0D–1D, 0D–3D, 1D–3D and 0D–1D–3D models. Within the former, the models and their applications are presented and discussed. Concerning the latter, a review was conducted on the methods upon which the coupling was based. The review finishes with an evaluation of the models’ benefits and their relations and an outlook on important future work in this field. A summary of chosen articles analyzed within this study is presented in Table 1 to give an overview of representative studies for each of the dimensions (0D, 1D, 3D, multi-scale).

3. Mathematical Models

Different spatial scales exist to describe the hemodynamics in the complex cerebrovascular structure numerically. These numerical models entail specific advantages and disadvantages regarding computing hours, provided information and proximity to reality depending on the scale. In the following section, the numerical models based on three spatial scales (0D, 1D, 3D) are introduced and the application for the neurovasculature is reviewed.

3.1. Zero-Dimensional Models

Zero-dimensional models are based on the similarity of the uniform distribution of pressure, volume, and flow in the human arteries compared to electrical current in circuits. Here, the vascular system is simplified into compartments describing the numerical connection using Windkessel (WK) equations [1,136]. This makes them valuable tools for studying pressure and flow distribution under specific physiological conditions, providing insights into the overall cardiovascular dynamics. The models follow Kirchhoff’s laws, yielding differential–algebraic equations and allowing efficient simulation of the circulatory system, accounting for the heart, venous system and regulatory dynamics with minimal computational resources [129,136]. Blood flow is described through principles like mass conservation and Poiseuille’s law for steady flow. These principles find an analog in electrical circuits with elements like resistance (R), inductance (L), capacitance (C), voltage (U) and current (I) [136], as presented in Figure 2.
The equations used within the electrical circuit are set as follows:
Ohm slaw : I = Δ U R P o i s e u i l l e s l a w : Q = Δ P R
C d P i n = ( Q i n Q o u t ) d t
L d P o u t = ( R · Q o u t ) d t
Values of R, L, and C are determined based on vessel geometry and mechanics. The underlying formulas are presented in the following:
R = 8 π μ l 3 V 2
L = 9 ρ l 4 π r 2
C = 3 π l r 3 2 E h
with ρ = fluid density, l = vessel length, r = vessel radius, E = elasticity module, h = vessel thickness [33,50,137]. WK models were developed to represent these principles effectively. Initially, the two-element WK model emerged, offering a reasonable portrayal of afterload on the heart and aortic pressure during diastole, although with less accuracy during systole. To address this, a three-element WK model was introduced, and further enhancements led to the four-element WK model [46]. For a more comprehensive understanding of microcirculation dynamics, researchers have explored five-element and six-element WK models. These models describe individual compartments and can be combined to create multi-compartment models, offering flexibility to tailor the models for specific research purposes [136].
  • Validation
One of the first WK computer simulations was implemented by Burkhoff et al. with the model parameters based on an in vivo animal study in 1988 [10]. Ten years later, Burattini et al. used an in vivo study as well to estimate a viscoelastic WK model. Using viscoelasticity optimizes the WK model compared to an elastic one [26]. First validation studies in regard to change in position (sitting to standing) were performed in 2004 by Olufsen et al. [38]. The study showed an increase in hemodynamic pressure and velocity when a position change is performed. This investigation is crucial for finding the true model parameters, since body movement and position change are neglected in conversational studies. Regarding the true WK model parameters, Pope et al. carried out a sensitivity analysis [39]. Here, five parameters were found to be crucial for reliability. Furthermore, the impact of aging was estimated and an increase in cerebral systemic resistance as well as peak systolic ventricular pressure was found. A reduction with aging was found only for arterial compliance. One of the first model reductions was performed by Vande Vosse et al., showing the advantage of lower computing costs but surely showing the disadvantage of accuracy [138].
  • Application
Zero-dimensional models in the field of cerebrovascular research were variously applied to either further develop the models or analyze different pathologies [33]. Abdi et al. investigated the impact of a stenotic vessel in the WK model [22]. A RCL network was used starting from the carotic communicating artery and results were validated with the literature. Zhang et al. focused on the variations in the anatomy of the circle of Willis based on different positions of the subject (supination to standing) [50]. Within this model, they further investigated the impact of a stenosis and validated their results with Doppler ultrasound measurements. Sedraie et al. utilized the model to analyze a stenosis and validation in this case was performed with the literature as well [42]. In contrast, Sethaput et al. investigated the impact of different arterial formations of the CoW on the blood flow with regard to occurring brain injuries [139]. Later on, McConnell et al. created a steady-state model of the neurovascular network and carried out simulations investigating different variations in occlusion and stenosis of the CoW [34]. They found that autoregulation is a crucial factor in collateral flow through the CoW. To optimize the model parameters, Lal et al. captured in vivo data and performed further development of the models [32]. Furthermore, Chnafa et al. validated the model with three-dimensional computational fluid dynamics data and experimental in vitro data [27]. The limitation in this study was the assumption of no energy loss over the bifurcations, which is then named as the crucial discrepancy to reality.

3.2. One-Dimensional Models

Using the cylindrical shape of blood vessels, 1D models are created to represent hemodynamic changes along arteries [1,129,140]. These models assume laminar blood flow with an impermeable, sliding wall. Some consider blood seepage into the wall [141,142], but radial and tangential velocities are usually neglected [143,144]. The 1D equations for mass and momentum conservation can be derived from control volumes or by integrating the incompressible Navier–Stokes equations across a cross-section [143]. The equations are presented in the following, with A: cross-sectional area, Q: flow, P: pressure, t: time, z: vessel length, ρ : density, K R : blood resistance [1,114].
A t + Q z = 0
Q t + z ( Q 2 A ) + A ρ P z + K R Q A = 0
To solve the system, an equation describing the material properties of the wall (parameters A and P) is required [1,114,132], with h: artery wall thickness, index 0: reference, E: elasticity module, σ : Poisson number.
P P 0 = E h 0 r 0 ( 1 σ 2 ) ( A A 0 1 )
Regarding the impact of the geometry, namely bifurcations within the vessels, flow conditions at these locations are described with the mass conservation and the overall pressure continuity. The equations are shown in the following, with indices 1 to 3 representing the proximal (1) and distal (2,3) parent arteries and U representing the mean axial velocity [23,85,145,146]:
A 1 U 1 = A 2 U 2 + A 3 U 3
P 1 + 1 2 ρ U 1 2 = P 2 + 1 2 ρ U 2 2 + P 3 + 1 2 ρ U 3 2
One-dimensional models are commonly used to simulate blood flow changes, averaged blood pressure and velocity. The arterial network is segmented and connected by nodes, with each segment modeled as a compliant tube described by the axial coordinate [143]. These models are valuable for analyzing wave propagation in the vascular system and understanding pressure dynamics in a large portion of the arterial network with reasonable computational effort [23,54,58,129,145,147]. Depending on the purpose, 1D models of the CoW vary in geometric and structural complexity.
The 1D models were created by different research groups over the recent years using several solvers and input data. Regarding the latter, some studies took in vivo data as an input and modeled the geometry of the intracranial arteries [29,64,125]. Further studies considered the flow from the left ventricle to the brain-supplying arteries as input [11,59,85,114,148].
  • Validation
Olufsen et al. examined the flow in large arteries with a 1D model and compared the blood flow and pressure results with MRI measurements, which showed good agreement [62]. Huang et al. developed a 1D model, which considered the heart, lungs and neurovascular arteries. Here, the results were validated with in vivo flow data taken from the literature [61]. After normalizing the captured in vivo data, due to low resolution and variation in subjects, these could be aligned with the simulated data. Park et al. as well confirmed their results from the 1D model with in vivo blood flow data from six volunteers [63]. In vivo validation was further performed by Zhang et al. using SPECT data and by Reymond et al. using Doppler measurements [65,114]. Alastruey et al. further performed validation against in vitro data through imitating the human arteries with tubes, imitating blood with a glycerol–water mixture and measuring pressure and flow parameters [23]. Ryu et al. compared their results to the literature [66], whereas Moorhead et al. compared them to higher scale models (3D) [147].
  • Application
One-dimensional models were applied to differing variations of the CoW [23,56]. Blanco et al. analyzed a broad network with a representation of 2000 vessels and compared it to a reduced 86-vessel network [149]. Within their study, the variation in complexity showed an impact on the hemodynamics when pathologies are present. One-dimensional models were applied to pathologies like stenosis [61,150] and were widely used to examine the systemic behavior of the cerebral hemodynamics search, thus assessing the collateral capacity of the CoW. Regarding changes in the pressure and flow waveforms, Yu et al. investigated the caused changes due to aging and smoking. Alastruey et al. focused on the changes due to different pulse waves. The evaluation of pressure dynamics can be performed following these aspects: [23,54,82,119,129,147].
Moreover, a part of the application is the post-surgery analysis of the hemodynamics. One-dimensional modeling can be applied to analyze the impact of implanting a prothesis or after carotid intestinectomy [23,37]. In recent studies, also the wall modeling was considered in 1D models. Park et al. performed a precise vascular wall modeling for quantifying the interaction between the blood flow and the wall biomechanics. They compared the linear elastic model with the viscoelastic Kelvin–Voigt model and verified the data with in vivo measurements. Here, the comparison showed that when using the linear elastic model only a 4% difference was found, being even lower when the viscoelastic model was used. Regarding autoregulation, Moorhead et al. applied transient resistance at the outlets to take vasodilation and vasoconstriction into account [147]. This was also approached by Ryu et al. and Alastruey et al., who applied a resistance at the outlets, which is independent from the flow conditions in their 1D model of the aorta to the cerebrovascular vessels [30,66].
Most 1D models include simplifications and neglect wall shear stress occurrence after Womersley and use only a reduced artery system [13,35,53,151]. Herein, Formaggia et al. implemented a full body circulation without the cerebral vessels and neglected the elasticity [85]. In contrast, Reymond et al. created a full-body circuit model using a viscoelastic wall modeling of the cerebral vessels and the left ventricle [65]. Within this model, formulas for resistance and convective acceleration were included and the results show an overestimation of the pressure when compared to in vivo data [1].
Despite these wall modeling approaches, 1D models carry a disadvantage regarding the analysis of pathologies. Comprising a vascular network with reduced information, the impact of morphology and cross-sections of specific pathological vessels are not taken into account [70,119,152]. To address this concern, 3D models are proposed, especially in terms of high-resolution analysis.

3.3. Three-Dimensional Models

Apart from reducing the vessel geometries to a 0D or 1D structure, the inclusion of the full vessel structure is implemented within the use of 3D models. These 3D models are firstly based on the Navier–Stokes equation comprising the mass and momentum conservation equations. Secondly, they are based on on a volumetric mesh, created within on a vessel geometry and the incompressible Navier–Stokes-Equations solved on each mesh point [153].
The mass conservation is a general continuity equation as shown in Equation (12).
ρ t + ( ρ u ) = 0
Due to incompressible flow conditions, the following formula is applied.
u = 0
The momentum equation is a continuum equation as the mass conservation and thus described as follows:
ρ ( d u d t + u · u ) = p + η 2 u + f
with ρ = blood density, u = velocity, p = pressure, η = blood viscosity, f = body force [1].
Several assumptions have to be implemented in the modeling of computational fluid dynamic (CFD), whereas the Navier–Stokes equations are solved in a 3D space over time [82]. CFD, applied to real 3D arterial structures from clinical imaging, enables precise assessment of flow patterns and shear stresses in complex geometries. This involves modeling continuous blood flows using the mentioned continuum equations, which are then discretized into matrix equations, solved and used to obtain hemodynamic solutions.
Together with these equations, the further assumptions to compute the local flow field and depending hemodynamic parameters are described in the following.
First, blood as a non-Newtonian fluid can be modeled with the Carreau model [80,119,154,155] or the Carreau–Yasuda model to mimic the non-Newtonian behavior [13]. Nevertheless, blood is mostly handled as Newtonian fluid with constant viscosity [15,16,156]. Berg et al., Fan et al. and Razavi et al. compared the impact of the choice of fluid modeling and found no significant difference between non-Newtonian and Newtonian blood composition in laminar flows [68,73,80].
Second, the vessel walls are elastic and in a few studies modeled as such [72,76]. Fluid structure interaction (FSI) models are used to represent blood flow interactions with vessel walls in three dimensions, considering the dynamic interaction between blood as a fluid and the vessel walls [82]. The simplification has led to overestimation of wall shear stress in some local regions [82]. Several research groups considered wall thickness and found out that WSS depends on the IA’s geometry [63,157]. Moreover, FSI was applied to pathologies like wall-near thrombus formation [18] or analysis of vessel flow after brain trauma [158]. Nevertheless, in most studies focusing on cerebral vessel flow, walls are assumed to be rigid [68,159].
Third, the complex choice of the boundary condition (BC) plays an important role in 3D hemodynamic modeling [160,161]. The most realistic inlet boundary condition will for sure be the in vivo curve measured with medical imaging, as applied in recent studies [68,75,76,83]. Nevertheless, due to the usual absence of in vivo data and restrictions in data capture regarding the spatial and temporal resolution, most studies take inlet flow from sample flow measurements in human subjects or assumptions based on the literature [72,162,163]. Regarding the outlet BC, the choice is as complex as for the inlet BC, since different approaches exist and the verification is rare due to medical image restrictions. In older studies, specific blood pressure or a zero pressure condition is used [80,81,164]. In recent research, novel splitting approaches are applied to better mimic the outflow conditions [16,27,162]. Moreover, 0D resistance models are often used at the vessel outlets [1,68,117,165], which will be further examined in Section 4.2. Urquiza et al. stated that the pressure BC has an impact on the overall pulse through the vessel [132]. Therefore, a precise pressure BC is crucial to avoid velocity discrepancies due to changes in pulse.
Three-dimensional models are therefore used to compute the highly resolved neurovascular hemodynamics in complex vessel structures [1]. In particular, the impact of the morphology and prescribed boundary conditions on these hemodynamics can be analyzed using 3D models [16,166]. The geometries can be allocated into two categories [1]: idealized models, which are created artificially based on the literature [75,80,81,117,165,167], and patient-specific models, which are based on in vivo data captured with medical imaging techniques [1,13,68,72,80,83,123,125,135,168].
  • Validation
Three-dimensional models were verified using the comparison to in vivo data or to in vitro data. The latter comprise phase-contrast magnetic resonance imaging (PC-MRI) or particle-image velocimetry (PIV). Roloff et al. performed the comparison of numerically assessed IA hemodynamics to PC-MRI and PIV measurements [169] and found good agreement between the modalities. Nevertheless, they found out that within MRI the velocity can be underestimated. Brindise et al. performed a multi-modality study as well, comparing the effect of the flow field on wall shear parameters [170]. Here, non-dimensional parameters seem to provide the most stabile results. Liang et al. compared their CFD results solely to PIV [123] and received good agreement. Korte and Gaidzik et al. performed a comparison of CFD to 4D flow MRI in phantom models based on mimicking the experiments with the simulations and received comparable flow fields as well [171].
  • Application
Three-dimensional simulations were applied to physiological and pathological phenomena of particular parts of the cerebral arteries. In particular, intracranial aneurysms and vessel stenoses were taken into account in several studies [68,124,167,172]. Also, the analysis of carotid bifurcation and carotid and other variations in stenosis and IA surgery is driven from 3D CFD data [77,117,173]. Other research groups have focused on the reversed Robin Hood syndrome [165], embolic particle pathlines [72] and the simulation of contrast agent and analysis of vessel wall parameters [69]. Despite the analysis of IA hemodynamics, also the treatment effect (TE) of IAs is analysed using 3D models, which provides the opportunity for TE prediction. This virtual analysis can especially be helpful for physicians. Namely, the virtual analysis of a bypass surgery [83] or cardiopulmonary bypass procedures [78] can be analyzed. Moreover, endovascular mechanical recanalization [125] and the application of flow diverter stents were studied in the past [15,174]. The effect of devices used for IA treatment, like coil and WEB, can be studied virtually and with pre- and post-treatment analysis. Furthermore, novel devices can be tested and the efficacy shown virtually [171,175].

4. Multi-Scale Models

Various modeling scales have different advantages to mathematically describe the hemodynamics in the body circulation. Modeling hemodynamic interactions with vessel walls is complex, especially when analyzing pathologies like IAs and plaques, which require detailed local flow data. Three-dimensional models are ideal for these cases but depend on accurate boundary conditions. While simulating the entire arterial system in 3D would be most accurate, it is computationally infeasible. Hence, an efficient approach is to use 3D models for specific regions of interest and lower-dimensional models for the rest of the network [117]. These multi-scale models combine different spatial scales and couple 3D simulations with 0D and 1D models to introduce more realistic constraints [135].

4.1. Coupling 0D–1D Models

Formaggia et al. introduced a 0D–1D coupling of the descending aorta (1D) and the cardiovascular system (0D) with 30 compartments [176]. The coupling was achieved both proximally and distally at specific interfaces, ensuring a smooth transition from 0D to 1D using an iterative transitioning between time steps.
A similar algorithm was used by Pontrelli et al. to investigate the pressure distribution on the arterial wall model [177]. Liang et al. developed a closed-loop model to investigate the influence of aortic and arterial stenosis in different locations [37,86]. The utilized 1D model was built from 55 arterial parameters taken from [35,87] and incorporated elastic properties. The interface conditions allowed them to derive unknown flow and pressure variables through extrapolated Riemannian invariants, approximated via numerical iteration due to the lack of analytical solutions at interfaces. They also applied this method to study left ventricular arterial coupling in the context of age-related hemodynamic changes. The model was further extended to include the cerebral circulation and investigate the influence of the CoW anatomy on hyperfusion after carotid artery surgery.
Zhang et al. adopted a more patient-specific approach, combining 0D modeling from single-photon emission computed tomography (SPECT) data and 1D modeling from MRI data to obtain individualized hemodynamic information in the CoW [114]. The peripheral resistance of each efferent CoW artery was adjusted to match the average flow over a cardiac cycle with the reference data, minimizing peripheral resistance errors. The coupling algorithm closely resembled that of Liang et al. [37,86,90]. Validation using PC-MRI data from efferent arteries confirmed the model’s ability to capture individual variations in CoW blood flow.
Devault et al. presented a 0D–1D coupled model using Doppler ultrasound data for validation of the intracranial velocity [84]. Notably, they introduced a Kalman filter to enhance the determination of the outflow boundary condition, excluding a subset of data used for validation. Alastruey et al. also solved the 0D–1D coupled model through the Riemann problem at the bifurcations [30]. The model demonstrated that introducing a resistance equivalent to the blood vessel’s characteristic impedance at the inlet of a terminal Windkessel model is crucial for preventing the creation of unrealistic wave reflections.
Mueller and Toro et al. focused on the representation of the cardiovascular system including the venous system in the head and neck area [178]. Large cerebral arteries and veins were modeled with a 1D model and the cardiovascular and pulmonary system were modeled with 0D models. The 1D model provides the pressure terms to the 0D model so that the latter can reproduce the flow. Arterial results were validated with data from the literature and venous results were validated with PC-MRI data.
Weber et al. implemented an open-source 1D–0D solver with a 0D modeling of the cardiovascular hemodynamics and a 1D modeling of the cerebrovascular system [110]. A viscoelastic modeling of the walls was applied and parameters were based on the literature.

4.2. Coupling 0D–3D Models

Sun et al. coupled a closed-loop 0D model (LPM) with a 3D CoW model extracted from CT angriographic data. They focused on the analysis of stenosed arteries in the CoW and the resulting effect of stenosis formation onto CoW hemodynamics.
Yu et al. presented 3D models of the vertebral arteries based on MRI with a 0D model of the CoW. The study showed the importance of unequally distributed vertebral arteries on the magnitude of the blood flow [148].
For Lee et al., the model coupling of 0D–3D was crucial to estimate an overview of the most important aspects of the coronary and neurovascular flow reserve [179]. In the study, the limitations of simulation approaches were discussed and suggestions presented to overcome these.
Morbiducci et al. conducted a study using a CFD model to analyze the carotid bifurcation and coupled it with a vascular bed modeled as a 0D model [120]. A comparison of fixed outflow rates to a 0D model was carried out. Their findings highlighted that employing more realistic boundary conditions leads to greater accuracy in understanding the role of local hemodynamics in vascular diseases. Regarding the BCs, a fixed velocity profile is set at the inlet and a RCR network at the outlets. The coupling between the two submodels required the flow and pressures to be transferred at the interface. To overcome the complexity of the transfer between the models, in recent studies, commercial solvers and implemented application programming interfaces (API) were employed. To be mentioned here is the ANSYS Fluent Code, which was used by several researchers [32,68,120,176]. Nevertheless, due to an increased likelihood of errors in the computation when using default codes with high complexity, most researchers have preferred to use a non-extensive LP model under simplified assumptions. Moreover, the previous method limits the multi-scale model of the cardiovascular system to a single 3D model, whereas coupling two or three detailed 3D simulations with extended LP modeling of the rest of the cardiovascular system offers a greater opportunity for broader applications [117]. Kashefi et al. introduced a CAD algorithm for efficiently coupling multiple 3D models with a 0D model, substantially reducing processing time. The model was applied to investigate the carotid bifurcation, focusing on identifying regions at risk for atherosclerotic lesions. In contrast to earlier studies, within this study the Matlab toolbox Simulink was used for analyzing the 0D model, while the Fluent software was employed solely for parallel 3D computations [117].

4.3. Coupling 1D–3D Models

Urquiza et al. presented a 1D–3D model of the entire arterial tree, including a 3D finite element model of the carotid bifurcation coupled with a 1D model for the remaining part of the arterial tree [132]. The results were consistent with reported flow patterns in the literature and the solely executed 1D model, indicating the model’s utility in understanding the formation and development of arterial diseases.
Ho et al. used 1D and 3D models constructed based on CT images to investigate hemodynamics in IAs [122]. The resulting wall shear stress stands in agreement with the previous literature data and was validated with ultrasound measurements. Although CFD results were within an acceptable range compared to the literature, diastolic velocity was overestimated. This was traced back to the rigid wall model and therefore assumed that the use of an elastic wall model is crucial for realistic investigation of carotid artery or aortic hemodynamics [122]. However, adding elastic wall properties requires high computational power and will not affect regional parameters crucially. Thus, this is neglected in many models, especially multi-scale models, due to already high computational costs [68,80,81,164,180]. Still, the arterial wall has a high effect on pressure wave propagation throughout the vascular network. This is why Passerini et al. set the focus of their study on the effect of the wall compliance at the transfer region from the 3D to 1D model [129] using a resistor–inductor–capacitor network at the inlet and outlet.
Liang et al. modeled an anterior communicating artery aneurysm with the 1D–3D approach [123]. Based on MRI, a patient-specific 1D model with elastic walls and a 3D model of the aneurysm area with rigid walls were created. Additionally, a 1D model based on population-wide data was developed and integrated with the aneurysm model. The comparison revealed differences in the flow field and patient-specific sensitivities of hemodynamic parameters. Especially in vortex structure, in wall shear stress magnitude and distribution, and in oscillatory shear index, variations were observed. In vitro data were captured for validating the underlying results.
Sutalo et al. coupled a 3D CoW model with a 1D model, representing the peripheral cerebral vasculature [131]. They found out that stenosed peripheral vessels of the posterior cerebral artery influence the flow in vertebral arteries and of the anterior and middle cerebral arteries influence the internal carotid arteries.

4.4. Coupling 0D–1D–3D Models

Formaggia et al. presented an early full 0D–1D–3D coupling [176] using a 3D approach for computing the high-resolved flow inside local vessels based on medical imaging and a 1D approach to calculate the peripheral vessels based on the literature. The flow in the capillaries was numerically assessed with a 0D model, with parameters based on the literature as well and the three scale models coupled to build one system. Nevertheless, in real-world investigations, this approach was applied only in a few studies [140].
Oshima et al. applied a full-scale coupled model to different CoW anatomies to investigate the hemodynamic changes due to a missing posterior communicating artery and the free-stream outflow conditions [135]. The latter was achieved by comparing the free outflow with a multi-scale 0D/1D model for peripheral vessels. Results were compared to previous work by Tanaka et al. and Gobin et al. [181,182]. The findings showed that applying multi-scale boundary condition brought the flow closer to those calculated by the literature.
Regarding the 0D–1D–3D modeling of cerebrovascular flow, Blanco et al. investigated the flow in patient-specific intracranial aneurysms [183] and the influence of the heart rate on the flow in the carotid arteries [17]. In each case, the 3D model described the local area (blood flow), the 1D model described the systemic and peripheral arteries (wave propagation) and the 0D model described the heart, venous and pulmonary circulation (microvascular system). They achieved creating a robust model with high scalability in components and a multi-time-stepping technique to reduce the computing time.
Liang et al. investigated the pre- and postoperative patient-specific hemodynamics using a 3D model coupled with a 0D–1D cardiovascular system model [134]. The 1D submodel contains the 87 largest arteries, including in particular the 40 largest cerebral arteries, to account for hemodynamic regulation within the CoW, and the 0D submodel contains the remaining cardiovascular part, while the 3D model locally represents the middle cerebral artery bifurcation. Here as well, walls were assumed to be rigid in terms of lower computational costs. Liang et al. highlighted the importance of multi-scale models in their work by comparing them with a simple three-dimensional model. Regarding the solving of the equations, direct coupling of a rigid 3D model with an elastic 1D model leads to discontinuities of the impedance at the connection interfaces between the models. This can lead to improper wave reflections and potential numerical instability. To address these issues, a 0D interface model was set between the 1D and 3D models, based on Passerini et al. [129].

5. Summary and Conclusions

In this systematic review, the recent research in multi-scale modeling of the neurovascular system to describe detailed and realistic hemodynamics was presented and discussed.
  • Multi-scale models as a realistic description of hemodynamic parameters within the circle of Willis
Coupling of differently scaled models could bring the advantage of taking the full body circulation into account when analyzing neurovascular hemodynamics. Since the performance of full 3D simulation of the entire vasculature carries a high computational effort, the coupled process shows the advantage of reducing this, especially if only small regions of interest need to be investigated in detail. Up to now, detailed neurovascular hemodynamics were mostly performed using 3D simulations with assumed boundary conditions. Performing 3D simulations with patient-specific boundary conditions measured with medical imaging like PC-MRI could give the most realistic outcome. However, drawbacks also exist within these measurements due to low resolution and noise as well as non-availability.
  • Advantages derived from multi-scale modeling
If medical image data are not available or not suitable to serve as boundary conditions for 3D modeling, the use of 1D modeling as boundary condition might be beneficial. With numerical 1D modeling in contrast to medical imaging, no harm of the patient and no use of hardware resources is necessary. Moreover, using patient-specific input data for these 1D models might make them similarly realistic to the measured data [123]. In the following, if the model can be adjusted specifically to each patient, a coupled 1D–3D simulation could give a proficient representation of detailed hemodynamics in specific vessel structures. Additionally, different flow variations, pathologies and vessel formations can be analyzed comparatively using the numerical analysis without interacting with the patient [160]. This also provides the advantage of testing treatment options with regard to prediction of treatment outcomes.
  • Limitations of the current status regarding multi-scale modeling
In order to advance the progress of applying the numerical methods to the clinic, currently existing limitations in multi-scale numerical modeling of neurovascular hemodynamics should be stressed out. First to mention are the high computational time and costs, which aggravate the clinical applicability crucially due to a fast-paced clinical routine. Although the flow results are accurate and can give information about hemodynamic behavior, diagnosis prediction and treatment outcome, the processing is still time-consuming and profound statistical data is missing. Second, modeling with a high degree of detail can result in discontinuities of the coupling system which can result in non-fitting flow and pressure curves [134]. Third, a lack of validation still exists regarding multi-scale modeling. This is due to a difficulty in validation itself, which is restricted by low-resolution medical images in in vivo data and by a distance from reality in in vitro data.
  • Remaining investigation needed for future improvement
The presented mathematical and multi-scale models feature a high numerical standard to generate realistic and patient-specific hemodynamic analysis. Nevertheless, some improvements are necessary to implement these models in the clinical routine. Herein, in vivo validation is still needed to proove the applicability of the derived results, and statistical analysis should be performed, to receive significant results. Moreover, a reduction in modeling effort should be addressed to be more clinically applicable. Especially in terms of fully coupled codes, this effort could already be reduced through standardizing the coupling process. Prescribed modeling within coupling libraries can offer useful possibilities regarding a broader application [184]. These could be further addressed and implemented in future studies.

Author Contributions

Conceptualization, J.K. and P.B.; methodology, J.K. and E.S.K.; investigation, J.K. and E.S.K.; writing—original draft preparation, J.K.; writing—review and editing, J.K., E.S.K. and P.B.; visualization, J.K.; supervision, P.B.; project administration, J.K.; funding acquisition, P.B. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge financial support for this publication from the German Research Foundation (SPP2311, project number: 465189657) and the German Federal Ministry of Education and Research within the Research Campus STIMULATE (grant no. 13GW0473A).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ACross-Sectional Area
CCapacitance
CFDComputational Fluid Dynamics
CoWCircle of Willis
CTComputed Tomography
DDimensional
FSIFluid Structure Interaction
ICurrent
IAIntracranial Aneurysm
K R Blood Resistance
LInductance
LPMLumped Parameter Model
MRIMagnetic Resonance Imaging
ODEOrdinary Differential Equation
PPressure
PCPhase Contrast
PDEPartial Differential Equation
PIVParticle-Image Velocimetry
QFlowrate
RResistance
SPECTSingle-Photon Emission Computed Tomography
tTime
UVoltage
WKWindkessel
zVessel Length

References

  1. Liu, H.; Wang, D.; Leng, X.; Zheng, D.; Chen, F.; Wong, L.; Shi, L.; Leung, T. State-of-the-art computational models of circle of willis with physiological applications: A Review. IEEE Access 2020, 8, 156261–156273. [Google Scholar] [CrossRef]
  2. Linn, F.; Rinkel, G.; Algra, A.; Van Gijn, J. Incidence of subarachnoid hemorrhage: Role of region, year, and rate of computed tomography: A meta-analysis. Stroke 1996, 27, 625–629. [Google Scholar] [CrossRef]
  3. Cebral, J.R.; Mut, F.; Weir, J.; Putman, C.M. Association of hemodynamic characteristics and cerebral aneurysm rupture. AJNR Am. J. Neuroradiol. 2011, 32, 264–270. [Google Scholar] [CrossRef]
  4. Meng, H.; Tutino, V.; Xiang, J.; Siddiqui, A. High WSS or how WSS? Complex interactions of hemodynamics with intracranial aneurysm initiation, growth, and rupture: Toward a unifying hypothesis. AJNR Am. J. Neuroradiol. 2013, 35, 1254–1262. [Google Scholar] [CrossRef]
  5. Voß, S.; Beuing, O.; Janiga, G.; Berg, P. Stent-induced vessel deformation after intracranial aneurysm treatment—A hemodynamic pilot study. Comput. Biol. Med. 2019, 111, 103338. [Google Scholar] [CrossRef] [PubMed]
  6. Markl, M.; Chan, F.P.; Alley, M.T.; Wedding, K.L.; Draney, M.T.; Elkins, C.J.; Parker, D.W.; Wicker, R.; Taylor, C.A.; Herfkens, R.J.; et al. Time-resolved three-dimensional phase-contrast MRI. J. Magn. Reson. Imaging 2003, 17, 499–506. [Google Scholar] [CrossRef] [PubMed]
  7. Pravdivtseva, M.; Gaidzik, F.; Berg, P.; Ulloa, P.; Larsen, N.; Jansen, O.; Hövener, J.B.; Salehi Ravesh, M. Influence of spatial resolution and compressed sense acceleration factor on flow quantification with 4d flow mri at 3 tesla. Tomography 2022, 8, 457–478. [Google Scholar] [CrossRef]
  8. Shi, Y.; Lawford, P.; Hose, R. Review of zero-d and 1-D models of blood flow in the cardiovascular system. Biomed. Eng. Online 2011, 10, 1–38. [Google Scholar] [CrossRef]
  9. Berg, P.; Saalfeld, S.; Voß, S.; Beuing, O.; Janiga, G. A review on the reliability of hemodynamic modeling in intracranial aneurysms: Why computational fluid dynamics alone cannot solve the equation. Neurosurg. Focus 2019, 47, E15. [Google Scholar] [CrossRef]
  10. Burkhoff, D.; Alexander, J.; Schipke, J. Assessment of windkessel as a model of aortic input impedance. Am. J. Physiol.-Heart Circ. Physiol. 1988, 255, H742–H753. [Google Scholar] [CrossRef]
  11. Fitchett, D. LV-arterial coupling: Interactive model to predict effect of wave reflections on LV energetics. Am. J. Physiol. 1991, 261, H1026–H1033. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, M.; Peng, F.; Tong, X.; Feng, X.; Li, Y.; Chen, H.; Niu, H.; Zhang, B.; Song, G.; Li, Y.; et al. Associations between haemodynamics and wall enhancement of intracranial aneurysm. Stroke Vasc. Neurol. 2021, 6, 467–475. [Google Scholar] [CrossRef] [PubMed]
  13. Moore, S.M.; Moorhead, K.T.; Chase, J.G.; David, T.; Fink, J. One-dimensional and three-dimensional models of cerebrovascular flow. J. Biomech. Eng. 2005, 127, 440–449. [Google Scholar] [CrossRef]
  14. Antiga, L.; Piccinelli, M.; Botti, L.; Ene-Iordache, B.; Remuzzi, A.; Steinman, D.A. An image-based modeling framework for patient-specific computational hemodynamics. Med. Biol. Eng. Comput. 2008, 46, 1097–1112. [Google Scholar] [CrossRef]
  15. Voß, S.; Beuing, O.; Janiga, G.; Berg, P. Multiple Aneurysms AnaTomy CHallenge 2018 (MATCH)-Phase Ib: Effect of morphology on hemodynamics. PLoS ONE 2019, 14, e0216813. [Google Scholar] [CrossRef] [PubMed]
  16. Korte, J.; Voß, S.; Janiga, G.; Beuing, O.; Behme, D.; Saalfeld, S.; Berg, P. Is accurate lumen segmentation more important than outlet boundary condition in image-based blood flow simulations for intracranial aneurysms? Cardiovasc. Eng. Technol. 2023, 14, 617–630. [Google Scholar] [CrossRef]
  17. Blanco, P.J.; Urquiza, S.A.; Feijóo, R.A. Assessing the influence of heart rate in local hemodynamics through coupled 3D-1D-0D models. Int. J. Numer. Methods Biomed. Eng. 2010, 26, 890–903. [Google Scholar] [CrossRef]
  18. Grinberg, L.; Fedosov, D.A.; Karniadakis, G.E. Parallel multiscale simulations of a brain aneurysm. J. Comput. Phys. 2013, 244, 131–147. [Google Scholar] [CrossRef]
  19. Kim, H.J.; Vignon-Clementel, I.E.; Figueroa, C.A.; LaDisa, J.F.; Jansen, K.E.; Feinstein, J.A.; Taylor, C.A. On coupling a lumped parameter heart model and a three-dimensional finite element aorta model. Ann. Biomed. Eng. 2009, 37, 2153–2169. [Google Scholar] [CrossRef]
  20. Moher, D.; Liberati, A.; Tetzlaff, J.; Altman, D.G.; PRISMA Group. Preferred reporting items for systematic reviews and meta-analyses: The PRISMA statement. BMJ 2009, 339, b2535. [Google Scholar] [CrossRef]
  21. Atkinson, K.M.; Koenka, A.C.; Sanchez, C.E.; Moshontz, H.; Cooper, H. Reporting standards for literature searches and report inclusion criteria: Making research syntheses more transparent and easy to replicate. Res. Synth. Methods 2014, 6, 87–95. [Google Scholar] [CrossRef] [PubMed]
  22. Abdi, K.A. Modeling the circle of willis using electrical analogy method under both normal and pathological circumstances. J. Biomed. Phys. Eng. 2013, 3, 45. [Google Scholar]
  23. Alastruey, J.; Parker, K.; Peiro, J.; Byrd, S.; Sherwin, S. Modeling the circle of willis to assess the effects of anatomical variations and occlusions on cerebral flows. J. Biomech. 2007, 40, 1794–1805. [Google Scholar] [CrossRef] [PubMed]
  24. Olufsen, M.; Nadim, A. On deriving lumped models for blood flow and pressure in the systemic arteries. Math. Biosci. Eng. MBE 2004, 1, 61–80. [Google Scholar] [CrossRef] [PubMed]
  25. Rideout, V.C. Mathematical and Computer Modeling of Physiological Systems; Prentice Hall: Upper Saddle River, NJ, USA, 1991. [Google Scholar]
  26. Burattini, R.; Natalucci, S. Complex and frequency-dependent compliance of viscoelastic windkessel resolves contradictions in elastic windkessels. Med. Eng. Phys. 1998, 20, 502–514. [Google Scholar] [CrossRef] [PubMed]
  27. Chnafa, C.; Brina, O.; Pereira, V.; Steinman, D. Better than nothing: A rational approach for minimizing the impact of outflow strategy on cerebrovascular simulations. Am. J. Neuroradiol. 2017, 39, 337–343. [Google Scholar] [CrossRef]
  28. Manini, S.; Antiga, L.; Botti, L.; Remuzzi, A. pyNS: An open-source framework for 0d haemodynamic modelling. Ann. Biomed. Eng. 2014, 43, 1461–1473. [Google Scholar] [CrossRef]
  29. Connolly, M.; He, X.; Gonzalez, N.; Vespa, P.; DiStefano, J.; Hu, X. Reproduction of consistent pulse-waveform changes using a computational model of the cerebral circulatory system. Med. Eng. Phys. 2014, 36, 354–363. [Google Scholar] [CrossRef]
  30. Alastruey, J.; Moore, S.M.; Parker, K.H.; David, T.; Peiró, J.; Sherwin, S.J. Reduced modelling of blood flow in the cerebral circulation: Coupling 1-D, 0-D and cerebral auto-regulation models. Int. J. Numer. Methods Fluids 2008, 56, 1061–1067. [Google Scholar] [CrossRef]
  31. Ursino, M.; Lodi, C. A simple mathematical model of the interaction between intracranial pressure and cerebral hemodynamics. J. Appl. Physiol. 1997, 82, 1256–1269. [Google Scholar] [CrossRef]
  32. Lal, R.; Nicoud, F.; Bars, E.L.; Deverdun, J.; Molino, F.; Costalat, V.; Mohammadi, B.; Lal, R.; Nicoud, F.; Bars, E.L.; et al. Non invasive blood flow features estimation in cerebral arteries from uncertain medical data. Biomed. Eng. 2017, 45, 2574–2591. [Google Scholar] [CrossRef] [PubMed]
  33. Li, X.; Bai, J.; Cui, S.; Wang, S. Simulation study of the cardiovascular functional status in hypertensive situation. Comput. Biol. Med. 2002, 32, 345–362. [Google Scholar] [CrossRef] [PubMed]
  34. McConnell, F.K.; Payne, S. The dual role of cerebral autoregulation and collateral flow in the circle of willis after major vessel occlusion. IEEE Trans. Biomed. Eng. 2017, 64, 1793–1802. [Google Scholar] [CrossRef]
  35. Stergiopulos, N.; Young, D.; Rogge, T. Computer simulation of arterial flow with applications to arterial and aortic stenoses. J. Biomech. 1993, 25, 1477–1488. [Google Scholar] [CrossRef] [PubMed]
  36. Fahrig, R.; Nikolov, H.; Fox, A.; Holdsworth, D. A three-dimensional cerebrovascular flow phantom. Med. Phys. 1999, 26, 1589–1599. [Google Scholar] [CrossRef] [PubMed]
  37. Liang, F.; Fukasaku, K.; Liu, H.; Takagi, S. A computational model study of the influence of the anatomy of the circle of willis on cerebral hyperperfusion following carotid artery surgery. Biomed. Eng. Online 2011, 10, 84. [Google Scholar] [CrossRef] [PubMed]
  38. Olufsen, M.; Tran, H.; Ottesen, J. Modeling cerebral blood flow control during posture change from sitting to standing. Cardiovasc. Eng. 2004, 4, 47–58. [Google Scholar] [CrossRef]
  39. Pope, S.R.; Ellwein, L.M.; Zapata, C.L.; Novak, V.; Kelley, C.T.; Olufsen, M.S. Estimation and identification of parameters in a lumped cerebrovascular model. Math. Biosci. Eng. 2009, 6, 93–115. [Google Scholar] [CrossRef]
  40. Ottesen, J.; Olufsen, M.; Larsen, J. Applied Mathematical Models in Human Physiology; Society for Industrial and Applied Mathematics: Philadelphia, PA, USA, 2004. [Google Scholar]
  41. Heldt, T. Computational Models of Cardiovascular Response to Orthostatic Stress. Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, USA, 2005. [Google Scholar]
  42. Sadraie, S.H.; Abdi, M.; Navidbakhsh, M.; Hassani, K.; Kaka, G.R. Modeling the heart beat, circle of willis and related cerebral stenosis using an equivalent electronic circuit. Biomed. Eng. Appl. Basis Commun. 2014, 26, 1450052. [Google Scholar] [CrossRef]
  43. MacIntosh, B.; Sideso, E.; Donahue, M.; Chappell, M.; Günther, M.; Handa, A.; Kennedy, J.; Jezzard, P. Intracranial hemodynamics is altered by carotid artery disease and after endarterectomy a dynamic magnetic resonance angiography study. Stroke J. Cereb. Circ. 2011, 42, 979–984. [Google Scholar] [CrossRef]
  44. O’Rourke, M. Accurate measurement of arterial pressure. J. Hum. Hypertens. 2003, 17, 445–447. [Google Scholar] [CrossRef] [PubMed]
  45. Hoksbergen, A.W.J.; Fülesdi, B.; Legemate, D.A.; Csiba, L. Collateral configuration of the circle of willis. Stroke 2000, 13, 1346–1351. [Google Scholar] [CrossRef] [PubMed]
  46. Westerhof, N.; Bosman, F.; Vries, C.J.D.; Noordergraaf, A. Analog studies of the human systemic arterial tree. J. Biomech. 1969, 2, 121–143. [Google Scholar] [CrossRef] [PubMed]
  47. Bergel, D. The Visco-Elastic Properties of the Arterial Wall. Ph.D. Thesis, Queen Mary University of London, London, UK, 1960. [Google Scholar]
  48. Patel, D.J.; Freitas, F.M.D.; Greenfield, J.C.; Fry, D.L. Relationship of radius to pressure along the aorta in living dogs. J. Appl. Physiol. 1963, 18, 1111–1117. [Google Scholar] [CrossRef]
  49. Noordergraaf, A.; Verdouw, P.D.; Boom, H.B. The use of an analog computer in a circulation model. Prog. Cardiovasc. Dis. 1963, 5, 419–439. [Google Scholar] [CrossRef]
  50. Zhang, C.; Li, S.; Pu, F.; Fan, Y.; Li, D. The effect of anatomic variations of circle of willis on cerebral blood distribution during posture change from supination to standing: A model study. Bio-Med. Mater. Eng. 2014, 24, 2371–2380. [Google Scholar] [CrossRef]
  51. Zhang, C.; Wang, L.; Li, X.; Li, S.; Pu, F.; Fan, Y.; Li, D. Modeling the circle of Willis to assess the effect of anatomical variations on the development of unilateral internal carotid artery stenosis. Bio-Med. Mater. Eng. 2014, 24, 491–499. [Google Scholar] [CrossRef]
  52. Alastruey, J. Numerical Modelling of Pulse Wave Propagation in the Cardiovascular System: Development, Validation and Clinical Applications. Ph.D. Thesis, Departments of Aeronautics and Bioengineering, Imperial College London, London, UK, 2006. [Google Scholar]
  53. Bessems, D.; Rutten, M.; van de Vosse, F. A wave propagation model of blood flow in large vessels using an approximate velocity profile function. J. Fluid Mech. 2007, 580, 145–168. [Google Scholar] [CrossRef]
  54. Cassot, F.; Zagzoule, M.; Marc-Vergnes, J.P. Hemodynamic role of the circle of Willis in stenoses of internal carotid arteries. An analytical solution of a linear model. J. Biomech. 2000, 33, 395–405. [Google Scholar] [CrossRef]
  55. Hillen, B.; Hoogstratent, H.W.; Post, L. A mathematical model of the flow in the circle of willis. J. Biomech. 1986, 19, 187–194. [Google Scholar] [CrossRef]
  56. Ferrandez, A.; David, T.; Brown, M.D. Numerical models of auto-regulation and blood flow in the cerebral circulation. Comput. Methods Biomech. Biomed. Eng. 2002, 5, 7–19. [Google Scholar] [CrossRef] [PubMed]
  57. Ferrández, A.; David, T.; Bamford, J.; Scott, J.; Guthrie, A. Computational models of blood flow in the circle of willis. Comput. Methods Biomech. Biomed. Eng. 2001, 4, 1–26. [Google Scholar] [CrossRef] [PubMed]
  58. Formaggia, L.; Lamponi, D.; Quarteroni, A. One-dimensional models for blood flow in arteries. J. Eng. Math. 2003, 47, 251–276. [Google Scholar] [CrossRef]
  59. Huang, P.G.; Muller, L.O. Simulation of one-dimensional blood flow in networks of human vessels using a novel TVD scheme. Int. J. Numer. Methods Biomed. Eng. 2015, 31, e02701. [Google Scholar] [CrossRef]
  60. Mueller, L.; Toro, E. Well-balanced high-order solver for blood flow in networks of vessels with variable properties. Int. J. Numer. Methods Biomed. Eng. 2013, 29, 1388–1411. [Google Scholar] [CrossRef]
  61. Huang, G.P.; Yu, H.; Yang, Z.; Schwieterman, R.; Ludwig, B. 1D simulation of blood flow characteristics in the circle of Willis using THINkS. Comput. Methods Biomech. Biomed. Eng. 2018, 21, 389–397. [Google Scholar] [CrossRef] [PubMed]
  62. Olufsen, M.S.; Peskin, C.S.; Kim, W.Y.; Pedersen, E.M.; Nadim, A.; Larsen, J. Numerical simulation and experimental validation of blood flow in arteries with structured-tree outflow conditions. Ann. Biomed. Eng. 2000, 28, 1281–1299. [Google Scholar] [CrossRef] [PubMed]
  63. Park, C.S.; Alaraj, A.; Du, X.; Charbel, F.T.; Linninger, A.A. An efficient full space-time discretization method for subject-specific hemodynamic simulations of cerebral arterial blood flow with distensible wall mechanics. J. Biomech. 2019, 87, 37–47. [Google Scholar] [CrossRef]
  64. Perdikaris, P.; Grinberg, L.; Karniadakis, G.E. An effective fractal-tree closure model for simulating blood flow in large arterial networks. Ann. Biomed. Eng. 2015, 43, 1432–1442. [Google Scholar] [CrossRef]
  65. Reymond, P.; Merenda, F.; Perren, F.; Rü, D.; Stergiopulos, N. Validation of a one-dimensional model of the systemic arterial tree. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, 208–222. [Google Scholar] [CrossRef]
  66. Ryu, J.; Hu, X.; Shadden, S.C. A coupled lumped-parameter and distributed network model for cerebral pulse-wave hemodynamics. J. Biomech. Eng. 2015, 137, 101009. [Google Scholar] [CrossRef]
  67. Moore, S.; David, T.; Chase, J.; Arnold, J.; Fink, J. 3D models of blood flow in the cerebral vasculature. J. Biomech. 2006, 39, 1454–1463. [Google Scholar] [CrossRef] [PubMed]
  68. Berg, P.; Stucht, D.; Janiga, G.; Beuing, O.; Speck, O.; Thévenin, D. Cerebral blood flow in a healthy circle of willis and two intracranial aneurysms: Computational fluid dynamics versus four-dimensional phase-contrast magnetic resonance imaging. J. Biomech. Eng. 2014, 136, 041003. [Google Scholar] [CrossRef] [PubMed]
  69. Cebral, J.R.; Castro, M.A.; Soto, O.; Löhner, R.; Alperin, N. Blood-flow models of the circle of Willis from magnetic resonance data. J. Eng. Math. 2003, 47, 369–386. [Google Scholar] [CrossRef]
  70. Chaichana, T.; Sun, Z.; Jewkes, J. Computation of hemodynamics in the left coronary artery with variable angulations. J. Biomech. 2011, 44, 1869–1878. [Google Scholar] [CrossRef] [PubMed]
  71. Vlachopoulos, C.; O’Rourke, M.; Nichols, W.W. McDonald’s Blood Flow in Arteries: Theoretical, Experimental and Clinical Principles; CRC Press: Boca Raton, FL, USA, 2011; Volume 6. [Google Scholar] [CrossRef]
  72. Fabbri, D.; Long, Q.; Das, S.; Pinelli, M. Computational modelling of emboli travel trajectories in cerebral arteries: Influence of microembolic particle size and density. Biomech. Model. Mechanobiol. 2014, 13, 289–302. [Google Scholar] [CrossRef] [PubMed]
  73. Fan, Y.; Jiang, W.; Zou, Y.; Li, J.; Chen, J.; Deng, X. Numerical simulation of pulsatile non-Newtonian flow in the carotid artery bifurcation. Acta Mech. Sin. Lixue Xuebao 2009, 25, 249–255. [Google Scholar] [CrossRef]
  74. Bharadvaj, B.K.; Mabon, R.F.; Giddens, D.P. Steady flow in a model of the human carotid bifurcation. Part I—Flow visualization. J. Biomech. 1982, 15, 349–362. [Google Scholar] [CrossRef]
  75. Ghaffari, M.; Tangen, K.; Alaraj, A.; Du, X.; Charbel, F.T.; Linninger, A.A. Large-scale subject-specific cerebral arterial tree modeling using automated parametric mesh generation for blood flow simulation. Comput. Biol. Med. 2017, 91, 353–365. [Google Scholar] [CrossRef]
  76. Grinberg, L.; Cheever, E.; Anor, T.; Madsen, J.R.; Karniadakis, G.E. Modeling blood flow circulation in intracranial arterial networks: A comparative 3D/1D simulation study. Ann. Biomed. Eng. 2011, 39, 297–309. [Google Scholar] [CrossRef]
  77. Mamatyukov, M.; Mikheev, I.; Parshin, D.; Khe, A.; Cherevko, A.; Orlov, K.; Chupakhin, A. Comprehensive research of human brain hemodynamics: Clinical monitoring and computer simulations. AIP Conf. Proc. 2018, 2027, 020009. [Google Scholar] [CrossRef]
  78. Piskin, S.; Ündar, A.; Pekkan, K. Computational modeling of neonatal cardiopulmonary bypass hemodynamics with full circle of willis anatomy. Artif. Organs 2015, 39, E164–E175. [Google Scholar] [CrossRef] [PubMed]
  79. Pekkan, K.; Dur, O.; Sundareswaran, K.; Kanter, K.; Fogel, M.; Yoganathan, A.; Undar, A. Neonatal aortic arch hemodynamics and perfusion during cardiopulmonary bypass. J. Biomech. Eng. 2009, 130, 061012. [Google Scholar] [CrossRef] [PubMed]
  80. Razavi, S.E.; Sahebjam, R. Numerical simulation of the blood flow behavior in the circle of Willis. BioImpacts 2014, 4, 89–94. [Google Scholar] [CrossRef]
  81. Ren, Y.; Chen, Q.; Li, Z.Y. A 3D numerical study of the collateral capacity of the Circle of Willis with anatomical variation in the posterior circulation. Biomed. Eng. Online 2015, 14, S11. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, Z.Q.; Xu, L.J.; Liu, R.; Liu, X.S.; Zhao, B.; Liang, F.Y. Importance of incorporating systemic cerebroarterial hemodynamics into computational modeling of blood flow in intracranial aneurysm. J. Hydrodyn. 2020, 32, 510–522. [Google Scholar] [CrossRef]
  83. Zhu, F.; Qian, Y.; Xu, B.; Gu, Y.; Karunanithi, K.; Zhu, W.; Chen, L.; Mao, Y.; Morgan, M.K. Quantitative assessment of changes in hemodynamics of the internal carotid artery after bypass surgery for moyamoya disease. J. Neurosurg. 2018, 129, 677–683. [Google Scholar] [CrossRef]
  84. Devault, K.; Gremaud, P.A.; Olufsen, M.S.; Novak, V.; Zhao, P.; Vernières, G. Blood flow in the circle of willis: Modeling and calibration. Multiscale Model. Simul. 2008, 7, 888–909. [Google Scholar] [CrossRef]
  85. Formaggia, L.; Lamponi, D.; Tuveri, M.; Veneziani, A. Numerical modeling of 1D arterial networks coupled with a lumped parameters description of the heart. Comput. Methods Biomech. Biomed. Eng. 2006, 9, 273–288. [Google Scholar] [CrossRef]
  86. Liang, F.; Takagi, S.; Himeno, R.; Liu, H. Multi-scale modeling of the human cardiovascular system with applications to aortic valvular and arterial stenoses. Med. Biol. Eng. Comput. 2009, 47, 743–755. [Google Scholar] [CrossRef]
  87. Olufsen, M. Modeling the Arterial System with Reference to an Anesthesia Simulator. Ph.D. Thesis, Rotskilde University, Roskilde, Denmark, 1998. [Google Scholar]
  88. Wang, J.; Parker, K. Wave propagation in a model of the arterial circulation. J. Biomech. 2004, 37, 457–470. [Google Scholar] [CrossRef] [PubMed]
  89. McEniery, C.M.; Yasmin, N.; Hall, I.R.; Qasem, A.; Wilkinson, I.B.; Cockcroft, J.R.; ACCT Investigators. Normal vascular aging: Differential effects on wave reflection and aortic pulse wave velocity. J. Am. Coll. Cardiol. 2005, 46, 1753–1760. [Google Scholar] [CrossRef] [PubMed]
  90. Liang, F.Y.; Takagi, S.; Himeno, R.; Liu, H. Biomechanical characterization of ventricular-arterial coupling during aging: A multi-scale model study. J. Biomech. 2009, 42, 692–704. [Google Scholar] [CrossRef] [PubMed]
  91. Sheng, C.; Sarwal, S.; Watts, K.; Marble, A. Computational simulation of blood flow in human systemic circulation incorporating an external force field. Med. Biol. Eng. Comput. 1995, 33, 8–17. [Google Scholar] [CrossRef] [PubMed]
  92. Davatgar, M. Numerical Simulation of Blood flow in the Systemic Vasculature Incorporatin Gravitational Force with Application to Cerebral Circulation. Ph.D. Thesis, Graduate School of Biomedical Engineering, University of New South Wales, Kensington, Australia, 2006. [Google Scholar]
  93. Werner, J.; Siskin, G.; Mandato, K.; Englander, M.; Herr, A. Review of venous anatomy for venographic interpretation in chronic cerebrospinal venous insufficiency. J. Vasc. Interv. Radiol. 2011, 22, 1681–1690. [Google Scholar] [CrossRef] [PubMed]
  94. Fortune, J.; Feustel, P. Effect of patient position on size and location of the subclavian vein for percutaneous puncture. Arch. Surg. 2003, 138, 996–1000. [Google Scholar] [CrossRef]
  95. Stringer, M.; Restieaux, M.; Fisher, A.; Crosado, B. The vertebral venous plexuses: The internal veins are muscular and external veins have valves. Clin. Anat. 2012, 25, 609–618. [Google Scholar] [CrossRef]
  96. Liu, W.; Wu, S.; Lin, X. Anatomical observation of the retromandibular vein by mandibular angle ostectomy. Chin. J. Tissue Eng. Res. 2010, 14, 9113–9116. [Google Scholar] [CrossRef]
  97. Tanoue, S.; Kiyosue, H.; Okahara, M.; Sagara, Y.; Hori, Y.; Kashiwagi, J.; Mori, H. Para-cavernous sinus venous structures: Anatomic variations and pathologic conditions evaluated on fat-suppressed 3D fast gradient-echo MR images. AJNR Am. J. Neuroradiol. 2006, 27, 1083–1089. [Google Scholar]
  98. Caruso, R.; Rosenbaum, A.; Chang, J.; Joy, S. Craniocervical junction venous anatomy on enhanced mr images: The suboccipital cavernous sinus. AJNR Am. J. Neuroradiol. 1999, 20, 1127–1131. [Google Scholar]
  99. Balak, N.; Ersoy, G.; Uslu, U.; Tanriover, N.; Tapul, L.; Cetin, G.; Isik, N.; Elmaci, I. Microsurgical and histomorphometric study of the occipital sinus: Quantitative measurements using a novel approach of stereology. Clin. Anat. 2010, 23, 386–393. [Google Scholar] [CrossRef]
  100. Louis, R.G.; Loukas, M.; Wartmann, C.T.; Tubbs, R.S.; Apaydin, N.; Gupta, A.A.; Spentzouris, G.; Ysique, J.R. Clinical anatomy of the mastoid and occipital emissary veins in a large series. Surg. Radiol. Anat. 2008, 31, 139–144. [Google Scholar] [CrossRef] [PubMed]
  101. Kutoglu, T.; Turut, M.; Kocabiyik, N.; Ozan, H.; Yildirim, M. Anatomical analysis of azygos vein system in human cadavers. Rom. J. Morphol. Embryol. = Rev. Roum. Morphol. Embryol. 2012, 53, 1051–1056. [Google Scholar]
  102. Arnautovic, K.; Al-Mefty, O.; Pait, T.; Krisht, A.; Husain, M. The suboccipital cavernous sinus. Neurosurg. Focus 1996, 1, E2. [Google Scholar] [CrossRef]
  103. Zachrisson, H.; Lindenberger, M.; Lanne, T. Diameter and compliance of the greater saphenous vein-the effect of age and glyceryl trinitrates. Faseb J. 2008, 22, 1211.10. [Google Scholar] [CrossRef]
  104. Kölegård, R.; Mekjavic, I.; Eiken, O. Increased distensibility in dependent veins following prolonged bedrest. Eur. J. Appl. Physiol. 2009, 106, 547–554. [Google Scholar] [CrossRef] [PubMed]
  105. Hertzberg, B.; Kliewer, M.; DeLong, D.; Lalouche, K.; Paulson, E.; Frederick, M.; Carroll, B. Sonographic assessment of lower limb vein diameters: Implications for the diagnosis and characterization of deep venous thrombosis. AJR Am. J. Roentgenol. 1997, 168, 1253–1257. [Google Scholar] [CrossRef]
  106. Oguzkurt, L.; Tercan, F.; Pourbagher, M.A.; Kizilkilic, O.; Turkoz, R.; Boyvat, F. Computed tomography findings in 10 cases of iliac vein compression (May–Thurner) syndrome. Eur. J. Radiol. 2005, 55, 421–425. [Google Scholar] [CrossRef]
  107. Das, K.; Begum, S.; Dey, S.; Quddus, M.; Mohiuddin, A. Sonographic Measurement Of Inferior Vena Cava Diameter—A Noninvasive Tool To Detect Acute Blood Loss. Ibrahim Med. Coll. J. 2012, 5, 45–50. [Google Scholar] [CrossRef]
  108. Desser, T.S.; Sze, D.Y.; Jeffrey, R.B. Imaging and Intervention in the Hepatic Veins. Am. J. Roentgenol. 2003, 180, 1583–1591. [Google Scholar] [CrossRef]
  109. Wafae, N.; Hirose, K.; Franco, C.; Wafae, G.; Ruiz, C.; Daher, L.; Person, O. The anatomy of the human thyroid veins and its surgical application. Folia Morphol. 2008, 67, 221–225. [Google Scholar]
  110. Wéber, R.; Gyürki, D.; Paál, G. First blood: An efficient, hybrid one- and zero-dimensional, modular hemodynamic solver. Int. J. Numer. Methods Biomed. Eng. 2023, 39, e3701. [Google Scholar] [CrossRef] [PubMed]
  111. Ferreira, A.; Chen, S.; Simaan, M.; Boston, J.; Antaki, J. A nonlinear state-space model of a combined cardiovascular system and a rotary pump. In Proceedings of the 44th IEEE Conference on Decision and Control, Seville, Spain, 12–15 December 2005; Volume 2005, pp. 897–902. [Google Scholar] [CrossRef]
  112. Kim, H.; Vignon-Clementel, I.; Coogan, J.; Figueroa, C.; Jansen, K.; Taylor, C. Patient-specific modeling of blood flow and pressure in human coronary arteries. Ann. Biomed. Eng. 2010, 38, 3195–3209. [Google Scholar] [CrossRef] [PubMed]
  113. Charlton, P.H.; Harana, J.M.; Vennin, S.; Li, Y.; Chowienczyk, P.; Alastruey, J. Modeling arterial pulse waves in healthy aging: A database for in silico evaluation of hemodynamics and pulse wave indexes. Am. J. Physiol. Heart Circ. Physiol. 2019, 317, H1062–H1085. [Google Scholar] [CrossRef] [PubMed]
  114. Zhang, H.; Fujiwara, N.; Kobayashi, M.; Yamada, S.; Liang, F.; Takagi, S.; Oshima, M. Development of a numerical method for patient-specific cerebral circulation using 1D–0D simulation of the entire cardiovascular system with SPECT data. Ann. Biomed. Eng. 2016, 44, 2351–2363. [Google Scholar] [CrossRef] [PubMed]
  115. Kobayashi, M.; Hoshina, K.; Yamamoto, S.; Nemoto, Y.; Akai, T.; Shigematsu, K.; Watanabe, T.; Ohshima, M. Development of an image-based modeling system to investigate evolutional geometric changes of a stent graft in an abdominal aortic aneurysm. Circ. J. Off. J. Jpn. Circ. Soc. 2015, 79, 1534–1541. [Google Scholar] [CrossRef]
  116. Oliver, J.; Webb, D. Noninvasive assessment of arterial stiffness and risk of atherosclerotic events. Arterioscler. Thromb. Vasc. Biol. 2003, 23, 554–566. [Google Scholar] [CrossRef] [PubMed]
  117. Kashefi, A.; Mahdinia, M.; Firoozabadi, B.; Amirkhosravi, M.; Ahmadi, G.; Saidi, M. Multidimensional modeling of the stenosed carotid artery: A novel CAD approach accompanied by an extensive lumped model. Acta Mech. Sin. 2014, 30, 259–273. [Google Scholar] [CrossRef]
  118. Gijsen, F.; van de Vosse, F.; Janssen, J. The influence of the non-Newtonian properties of blood on the flow in large arteries: Steady flow in a carotid bifurcation model. J. Biomech. 1999, 32, 601–608. [Google Scholar] [CrossRef]
  119. Liu, X.; Gao, Z.; Xiong, H.; Ghista, D.; Ren, L.; Zhang, H.; Wu, W.; Huang, W.; Hau, W.K. Three-dimensional hemodynamics analysis of the circle of Willis in the patient-specific nonintegral arterial structures. Biomech. Model. Mechanobiol. 2016, 15, 1439–1456. [Google Scholar] [CrossRef]
  120. Morbiducci, U.; Gallo, D.; Massai, D.; Consolo, F.; Ponzini, R.; Antiga, L.; Bignardi, C.; Deriu, M.A.; Redaelli, A. Outflow conditions for image-based hemodynamic models of the carotid bifurcation: Implications for indicators of abnormal flow. J. Biomech. Eng. 2010, 132, 091005. [Google Scholar] [CrossRef] [PubMed]
  121. Sun, H.; Li, B.; Zhang, L.; Zhang, Y.; Liu, J.; Huang, S.; Xi, X.; Liu, Y. Numerical study of hemodynamic changes in the Circle of Willis after stenosis of the internal carotid artery. Comput. Methods Programs Biomed. 2024, 243, 107881. [Google Scholar] [CrossRef]
  122. Ho, H.; Sands, G.; Schmid, H.; Mithraratne, K.; Mallinson, G.; Hunter, P. A hybrid 1D and 3D approach to hemodynamics modelling for a patient-specific cerebral vasculature and aneurysm. In Proceedings of the Medical Image Computing and Computer-Assisted Intervention–MICCAI 2009: 12th International Conference, London, UK, 20–24 September 2009; Volume 12, pp. 323–330. [Google Scholar] [CrossRef]
  123. Liang, F.; Liu, X.; Yamaguchi, R.; Liu, H. Sensitivity of flow patterns in aneurysms on the anterior communicating artery to anatomic variations of the cerebral arterial network. J. Biomech. 2016, 49, 3731–3740. [Google Scholar] [CrossRef] [PubMed]
  124. Marzo, A.; Singh, P.; Larrabide, I.; Radaelli, A.; Coley, S.; Gwilliam, M.; Wilkinson, I.D.; Lawford, P.; Reymond, P.; Patel, U.; et al. Computational hemodynamics in cerebral aneurysms: The effects of modeled versus measured boundary conditions. Ann. Biomed. Eng. 2011, 39, 884–896. [Google Scholar] [CrossRef] [PubMed]
  125. Neidlin, M.; Büsen, M.; Brockmann, C.; Wiesmann, M.; Sonntag, S.J.; Steinseifer, U.; Kaufmann, T.A. A numerical framework to investigate hemodynamics during endovascular mechanical recanalization in acute stroke. Int. J. Numer. Methods Biomed. Eng. 2016, 32, e02748. [Google Scholar] [CrossRef] [PubMed]
  126. Enzmann, D.; Ross, M.; Marks, M.; Pelc, N. Blood flow in major cerebral arteries measured by phase-contrast cine MR. AJNR Am. J. Neuroradiol. 1994, 15, 123–129. [Google Scholar]
  127. Wåhlin, A.; Ambarki, K.; Birgander, R.; Wieben, O.; Johnson, K.; Malm, J.; Eklund, A. Measuring pulsatile flow in cerebral arteries using 4d phase-contrast mr imaging. AJNR Am. J. Neuroradiol. 2013, 34, 1740–1745. [Google Scholar] [CrossRef]
  128. Ooij, P.; Zwanenburg, J.; Visser, F.; Majoie, C.; Vanbavel, E.; Hendrikse, J.; Nederveen, A. Quantification and visualization of flow in the circle of willis: Time-resolved three-dimensional phase contrast mri at 7 T compared with 3 T. Magn. Reson. Med. 2013, 69, 868–876. [Google Scholar] [CrossRef]
  129. Passerini, T.; de Luca, M.; Formaggia, L.; Quarteroni, A.; Veneziani, A. A 3D/1D geometrical multiscale model of cerebral vasculature. J. Eng. Math. 2009, 64, 319–330. [Google Scholar] [CrossRef]
  130. Stergiopulos, N.; Westerhof, B.; Westerhof, N. Total arterial inertance as the fourth element of the Windkessel model. Am. J. Physiol. 1999, 276, H81–H88. [Google Scholar] [CrossRef]
  131. Šutalo, I.; Bui, A.; Ahmed, S.; Liffman, K.; Manasseh, R. Modeling of Flow Through The Circle of Willis and Cerebral Vasculature to Assess The Effects of Changes In The Peripheral Small Cerebral Vasculature on The Inflows. Eng. Appl. Comput. Fluid Mech. 2015, 8, 609–622. [Google Scholar] [CrossRef]
  132. Urquiza, S.A.; Blanco, P.J.; Vénere, M.J.; Feijóo, R.A. Multidimensional modelling for the carotid artery blood flow. Comput. Methods Appl. Mech. Eng. 2006, 195, 4002–4017. [Google Scholar] [CrossRef]
  133. Hughes, T.J.R.; Brooks, A.N. Streamline upwind/Petrov-Galerkin formulations for convection dominated flows with particular emphasis on the incompressible Navier-Stokes equations. Comput. Methods Appl. Mech. Eng. 1982, 32, 199–259. [Google Scholar] [CrossRef]
  134. Liang, F.; Oshima, M.; Huang, H.; Liu, H.; Takagi, S. Numerical study of cerebroarterial hemodynamic changes following carotid artery operation: A comparison between multiscale modeling and stand-alone three-dimensional modeling. J. Biomech. Eng. 2015, 137, 101011. [Google Scholar] [CrossRef]
  135. Oshima, M.; Torii, R.; Tokuda, S.; Yamada, S.; Koizumi, A. Patient-specific modeling and multi-scale blood simulation for computational hemodynamic study on the human cerebrovascular system. Curr. Pharm. Biotechnol. 2012, 13, 2153–2165. [Google Scholar] [CrossRef] [PubMed]
  136. Kokalari, I.; Karaja, T.; Guerrisi, M. Review on lumped parameter method for modeling the blood flow in systemic arteries. J. Biomed. Sci. Eng. 2013, 6, 92–99. [Google Scholar] [CrossRef]
  137. Avolio, A.P. Multi-branched model of the human arterial system. Biol. Eng. Comput. 1980, 18, 709–718. [Google Scholar] [CrossRef]
  138. van de Vosse, F.; Stergiopulos, N. Pulse wave propagation in the arterial tree. Annu. Rev. Fluid Mech. 2011, 43, 467–499. [Google Scholar] [CrossRef]
  139. Sethaput, T. Mathematical Model for Hemodynamic and Intracranial Windkessel Mechanism. Ph.D. Thesis, Case Western Reserve University, Cleveland, OH, USA, 2013. [Google Scholar]
  140. Quarteroni, A.; Ragni, S.; Veneziani, A. Coupling between lumped and distributed models for blood flow problems. Comput. Vis. Sci. 2001, 4, 111–124. [Google Scholar] [CrossRef]
  141. Schaaf, B.; Abbrecht, P. Digital computer simulation of human systemic arterial pulse wave transmission: A nonlinear model. J. Biomech. 1972, 5, 345–364. [Google Scholar] [CrossRef]
  142. Stettler, J.C.; Niederer, P.; Anliker, M. Theoretical analysis of arterial hemodynamics including the influence of bifurcations-Part I: Mathematical model and prediction of normal pulse patterns. Ann. Biomed. Eng. 1981, 9, 145–164. [Google Scholar] [CrossRef]
  143. Alastruey, J.; Parker, K.H.; Sherwin, S.J. Arterial pulse wave haemodynamics. In Proceedings of the 11th International Conference on Pressure, Lisbon, Portugal, 24–26 October 2012; pp. 401–443. [Google Scholar]
  144. Pullan, A.; Smith, N.; Hunter, P. An Anatomically Based Model of Transient Coronary Blood Flow in the Heart. SIAM J. Appl. Math. 2002, 62, 990–1018. [Google Scholar] [CrossRef]
  145. Sherwin, S.J.; Franke, V.; Peiró, J.; Parker, K. One-dimensional modelling of a vascular network in space-time variables. J. Eng. Math. 2003, 47, 217–250. [Google Scholar] [CrossRef]
  146. Smith, F.; Ovenden, N.; Franke, P.; Doorly, D. What happens to pressure when a flow enters a side branch? J. Fluid Mech. 2003, 479, 231–258. [Google Scholar] [CrossRef]
  147. Moorhead, K.T.; Doran, C.V.; Chase, J.G.; David, T. Lumped parameter and feedback control models of the auto-regulatory response in the circle of willis. Comput. Methods Biomech. Biomed. Eng. 2004, 7, 121–130. [Google Scholar] [CrossRef] [PubMed]
  148. Yu, L.; Xinmiao, Z.; Yawei, W.; Wentao, F.; Jing, J.; Zhunjun, S.; Wang, B.; Yongjun, W.; Yubo, F. Effects of abnormal vertebral arteries and the circle of Willis on vertebrobasilar dolichoectasia: A multi-scale simulation study. Clin. Biomech. 2022, 101, 105853. [Google Scholar] [CrossRef] [PubMed]
  149. Blanco, P.J.; Müller, L.O.; Watanabe, S.M.; Feijóo, R.A. On the anatomical definition of arterial networks in blood flow simulations: Comparison of detailed and simplified models. Biomech. Model. Mechanobiol. 2020, 19, 1663–1678. [Google Scholar] [CrossRef]
  150. Majka, M.; Gadda, G.; Taibi, A.; Gałazka, M.; Zieliński, P. Earliest effects of sudden occlusions on pressure profiles in selected locations of the human systemic arterial system. Phys. Rev. E 2017, 95, 032414. [Google Scholar] [CrossRef]
  151. Azer, K.; Peskin, C.S. A one-dimensional model of blood flow in arteries with friction and convection based on the Womersley velocity profile. Cardiovasc. Eng. 2007, 7, 51–73. [Google Scholar] [CrossRef]
  152. Xie, X.; Yuanyuan, W.; Zhu, H.; Zhou, J. Computation of hemodynamics in tortuous left coronary artery: A morphological parametric study. J. Biomech. Eng. 2014, 136, 101006. [Google Scholar] [CrossRef]
  153. Perron, S.; Boivin, S.; Hérard, J.M. A finite volume method to solve the 3D Navier–Stokes equations on unstructured collocated meshes. Comput. Fluids 2004, 33, 1305–1333. [Google Scholar] [CrossRef]
  154. Rezaie, H.; Ashrafizadeh, A.; Mojra, A. A patient-specific three-dimensional hemodynamic model of the circle of willis. Cardiovasc. Eng. Technol. 2017, 8, 495–504. [Google Scholar] [CrossRef]
  155. Guangyu, Z.; Yuan, Q.; Yang, J.; Yeo, J. The role of the circle of Willis in internal carotid artery stenosis and anatomical variations: A computational study based on a patient-specific three-dimensional model. Biomed. Eng. Online 2015, 14, 107. [Google Scholar] [CrossRef]
  156. Valen-Sendstad, K.; Bergersen, A.W.; Shimogonya, Y.; Goubergrits, L.; Bruening, J.; Pallares, J.; Cito, S.; Piskin, S.; Pekkan, K.; Geers, A.J.; et al. Real-world variability in the prediction of intracranial aneurysm wall shear stress: The 2015 international aneurysm cfd challenge. Cardiovasc. Eng. Technol. 2018, 9, 544–564. [Google Scholar] [CrossRef]
  157. Jahed, M.; Ghalichi, F.; Farhoudi, M. Fluid-structure interaction of patient-specific circle of willis with aneurysm: Investigation of hemodynamic parameters. Bio-Med. Mater. Eng. 2018, 29, 357–368. [Google Scholar] [CrossRef] [PubMed]
  158. Razaghi, R.; Biglari, H.; Karimi, A. Risk of rupture of the cerebral aneurysm in relation to traumatic brain injury using a patient-specific fluid-structure interaction model. Comput. Methods Programs Biomed. 2019, 176, 9–16. [Google Scholar] [CrossRef]
  159. Rahma, A.G.; Yousef, K.; Abdelhamid, T. Blood flow CFD simulation on a cerebral artery of a stroke patient. SN Appl. Sci. 2022, 4, 261. [Google Scholar] [CrossRef]
  160. Xiang, J.; Siddiqui, A.; Meng, H. The effect of inlet waveforms on computational hemodynamics of patient-specific intracranial aneurysms. J. Biomech. 2014, 47, 3882–3890. [Google Scholar] [CrossRef]
  161. Mut, F.; Ruijters, D.; Babic, D.; Bleise, C.; Lylyk, P.; Cebral, J. Effects of changing physiologic conditions on the in vivo quantification of hemodynamic variables in cerebral aneurysms treated with flow diverting devices. Int. J. Numer. Methods Biomed. Eng. 2014, 30, 135–142. [Google Scholar] [CrossRef]
  162. Saalfeld, S.; Voß, S.; Beuing, O.; Preim, B.; Berg, P. Flow-splitting-based computation of outlet boundary conditions for improved cerebrovascular simulation in multiple intracranial aneurysms. Int. J. Comput. Assist. Radiol. Surg. 2019, 14, 1805–1813. [Google Scholar] [CrossRef]
  163. Berg, P.; Saalfeld, S.; Janiga, G.; Brina, O.; Cancelliere, N.; Machi, P.; Pereira, V. Virtual stenting of intracranial aneurysms: A pilot study for the prediction of treatment success based on hemodynamic simulations. Int. J. Artif. Organs 2018, 41, 698–705. [Google Scholar] [CrossRef]
  164. Sekhane, D. Image-based computational fluid dynamics (CFD) modeling cerebral blood flow in the circle of willis. J. Adv. Res. Phys. 2016, 6, 1–5. [Google Scholar]
  165. Piechna, A.; Cieslicki, K. Reversed robin hood syndrome in the light of nonlinear model of cerebral circulation. Int. J. Appl. Mech. Eng. 2017, 22, 459–464. [Google Scholar] [CrossRef]
  166. Alnæs, M.S.; Isaksen, J.; Mardal, K.A.; Romner, B.; Morgan, M.K.; Ingebrigtsen, T. Computation of hemodynamics in the circle of Willis. Stroke 2007, 38, 2500–2505. [Google Scholar] [CrossRef] [PubMed]
  167. Nam, S.W.; Choi, S.; Cheong, Y.; Kim, Y.H.; Park, H.K. Evaluation of aneurysm-associated wall shear stress related to morphological variations of circle of Willis using a microfluidic device. J. Biomech. 2015, 48, 348–353. [Google Scholar] [CrossRef]
  168. Šutalo, I.D.; Bui, A.; Ahmed, S.; Liffman, K.; Manasseh, R. Modelling of flow through the circle of willis and cerebral vasculature. Model. Med. Biol. VIII 2009, 13, 83–92. [Google Scholar] [CrossRef]
  169. Roloff, C.; Stucht, D.; Beuing, O.; Berg, P. Comparison of intracranial aneurysm flow quantification techniques: Standard PIV vs stereoscopic PIV vs tomographic PIV vs phase-contrast MRI vs CFD. J. Neurointerventional Surg. 2018, 11, 275–282. [Google Scholar] [CrossRef] [PubMed]
  170. Brindise, M.; Rothenberger, S.; Dickerhoff, B.; Schnell, S.; Markl, M.; Saloner, D.; Rayz, V.; Vlachos, P. Multi-modality cerebral aneurysm haemodynamic analysis: In vivo 4D flow MRI, in vitro volumetric particle velocimetry and in silico computational fluid dynamics. J. R. Soc. Interface 2019, 16, 20190465. [Google Scholar] [CrossRef]
  171. Korte, J.; Gaidzik, F.; Larsen, N.; Schütz, E.; Damm, T.; Wodarg, F.; Hövener, J.B.; Jansen, O.; Janiga, G.; Berg, P.; et al. In vitro and in silico assessment of flow modulation after deploying the Contour Neurovascular System in intracranial aneurysm models. J. Neurointerventional Surg. 2023. [Google Scholar] [CrossRef]
  172. Steinman, D.; Milner, J.; Norley, C.; Lownie, S.; Holdsworth, D. Image-Based Computational Simulation of Flow Dynamics in a Giant Intracranial Aneurysm. AJNR Am. J. Neuroradiol. 2003, 24, 559–566. [Google Scholar]
  173. Perktold, K.; Resch, M.; Florian, H. Pulsatile non-newtonian flow characteristics in a three-dimensional human carotid bifurcation model, ASME. J. Biomech. Eng. 1991, 113, 464–475. [Google Scholar] [CrossRef]
  174. Barletta, E.A.; Ricci, R.L.; Di Silva, R.G.; Gaspar, R.H.M.L.; Araújo, J.F.M.; Neves, M.W.F.; de Aquino, J.L.B.; Barba Belsuzarri, T.A. Fusiform aneurysms: A review from its pathogenesis to treatment options. Surg. Neurol. Int. 2018, 9, 189. [Google Scholar] [CrossRef]
  175. Stahl, J.; Marsh, L.; Thormann, M.; Ding, A.; Saalfeld, S.; Behme, D.; Berg, P. Assessment of the flow-diverter efficacy for intracranial aneurysm treatment considering pre- and post-interventional hemodynamics. Comput. Biol. Med. 2023, 156, 106720. [Google Scholar] [CrossRef] [PubMed]
  176. Formaggia, L.; Nobile, F.; Quarteroni, A.; Veneziani, A. Computing and visualization in science multiscale modelling of the circulatory system: A preliminary analysis. Comput. Visual Sci. 1999, 2, 75–83. [Google Scholar] [CrossRef]
  177. Pontrelli, G. A multiscale approach for modelling wave propagation in an arterial segment. Comput. Methods Biomech. Biomed. Eng. 2004, 7, 79–89. [Google Scholar] [CrossRef] [PubMed]
  178. Müller, L.O.; Toro, E.F. A global multiscale mathematical model for the human circulation with emphasis on the venous system. Int. J. Numer. Methods Biomed. Eng. 2014, 30, 681–725. [Google Scholar] [CrossRef]
  179. Lee, K.E.; Ryu, A.J.; Shin, E.S.; Shim, E.B. Physiome approach for the analysis of vascular flow reserve in the heart and brain. Pflugers Arch. Eur. J. Physiol. 2017, 469, 613–628. [Google Scholar] [CrossRef]
  180. Voß, S.; Glaßer, S.; Hoffmann, T.; Beuing, O.; Weigand, S.; Jachau, K.; Preim, B.; Thévenin, D.; Janiga, G.; Berg, P. Fluid-structure simulations of a ruptured intracranial aneurysm: Constant versus patient-specific wall thickness. Comput. Math. Methods Med. 2016, 2016, 9854539. [Google Scholar] [CrossRef]
  181. Tanaka, H.; Fujita, N.; Enoki, T.; Matsumoto, K.; Watanabe, Y.; Murase, K.; Nakamura, H. Relationship between variations in the circle of Willis and flow rates in internal carotid and basilar arteries determined by means of magnetic resonance imaging with semiautomated lumen segmentation: Reference data from 125 healthy volunteers. AJNR Am. J. Neuroradiol. 2006, 27, 1770–1775. [Google Scholar]
  182. Gobin, Y.; Cloughesy, T.; Chow, K.; Duckwiler, G.; Sayre, J.; Milanese, K.; Viñuela, F. Intraarterial chemotherapy for brain tumors by using a spatial dose fractionation algorithm and pulsatile delivery. Radiology 2001, 218, 724–732. [Google Scholar] [CrossRef]
  183. Blanco, P.; Leiva, J.; Buscaglia, G. A black-box decomposition approach for coupling heterogeneous components in hemodynamics simulations. Int. J. Numer. Methods Biomed. Eng. 2013, 29, 408–427. [Google Scholar] [CrossRef] [PubMed]
  184. Chourdakis, G.; Davis, K.; Rodenberg, B.; Schulte, M.; Simonis, F.; Uekermann, B.; Abrams, G.; Bungartz, H.; Cheung Yau, L.; Desai, I.; et al. preCICE v2: A sustainable and user-friendly coupling library. Open Res. Eur. 2022, 2, 51. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Flowchart describing the underlying study selection process for the performed review analysis based on [20,21].
Figure 1. Flowchart describing the underlying study selection process for the performed review analysis based on [20,21].
Bioengineering 11 00072 g001
Figure 2. Sample arterial compartment comprising a 3-element Windkessel model with compliance (C), resistance (R) and inductance (L) [136].
Figure 2. Sample arterial compartment comprising a 3-element Windkessel model with compliance (C), resistance (R) and inductance (L) [136].
Bioengineering 11 00072 g002
Table 1. Summary of articles reviewed in this study dealing with in silico approaches to describe hemodynamics in the neurovascular and cardiovascular system in low-, high-, and multi-scales. PS: Pulmonary System; ICA: Internal Carotid Artery; MCA: Middle Cerebral Artery; ACA: Anterior Communicating Artery; TOF: Time of Flight; MRA: Magnetic Resonance Angiography; MRI: Magnetic Resonance Imaging; CT: Computed Tomography, D: dimension.
Table 1. Summary of articles reviewed in this study dealing with in silico approaches to describe hemodynamics in the neurovascular and cardiovascular system in low-, high-, and multi-scales. PS: Pulmonary System; ICA: Internal Carotid Artery; MCA: Middle Cerebral Artery; ACA: Anterior Communicating Artery; TOF: Time of Flight; MRA: Magnetic Resonance Angiography; MRI: Magnetic Resonance Imaging; CT: Computed Tomography, D: dimension.
ArticleDAnatomyLoopWall ModelSource of InputdataSoftware/SolverDerived ResultsComparison
ABDI et al. (2013) [22]0DCoW, Heart,closedneglectedLiterature [23]MATLABPressure–time curvesExperiments [24,25]
  PS      
BURATTINI et al. (1998) [26]0DHeartopenviscoelasticAnimal studyComputer simulationComplianceWK2, WK3, VW, FPM models
BURKHOFF et al. (1988) [10]0DAortaopenneglectedAnimal studyComputer simulationPressure, ImpedanceAnimal study
CHNAFA et al. (2017) [27]0DICA, MCA,openneglected3D angiogramspyNS [28]Flow3D CFD results
  ACA      
CONNOLLY et al. (2014) [29]0DCoWopenelasticLiterature [23,30,31]MATLAB CVXFlow, Pulse waveformsVasoconstriction to vasodilation
LAL et al. (2017) [32]0DCoW, AortaopenelasticTOF MRA, Kalman filterIn-houseResisistance, ComplianceData assimilation
LI et al. (2002) [33]0DCoW, HeartclosedneglectedDoppler measurementsDelphi 5.0PressureDoppler measurements
MCCONNELL et al. (2017) [34]0DCoWopenrigidLiterature [23,35,36]In-houseFlowLiterature [37]
OLUFSEN et al. (2004) [38]0DCoW, Heart,closedneglectedDoppler measurementsMATLABPressure, VelocityDoppler measurements
  Full body      
POPE et al. (2009) [39]0DCoW, HeartclosedneglectedDoppler measurements,MATLABPressure, VelocityDoppler measurements
     Literature [40,41]   
SADRAIE et al. (2014) [42]0DCoW, HeartopenneglectedLiterature [23,43,44]MATLAB SimulinkPressure wave propagationLiterature [45]
WESTERHOF et al. (1969) [46]0DFull bodyopenneglectedLiterature [47,48,49]n/aImpedance, PressureDoppler measurements
ZHANG et al. (2014) [50]0DCoWopenrigidLiterature [23]n/aFlowDoppler measurements
ZHANG et al. (2014) [51]0DCoWopenrigidLiterature [36]Fourth-order Runge–Kutta algorithmFlowDoppler measurements
ALASTRUEY et al. (2007) [23]1DCoW, AortaopenelasticLiterature [35,36]discontinuous Galerkin schemeFlow/pressure wave propagationLiterature [52]
BESSEMS et al. (2007) [53]1DNoneopenelasticGalerkinGalerkin weighted-residual methodFlow/pressure wave propagationWomersley profiles
     weighted-residuals method   
CASSOT et al. (2000) [54]1DCoWopenviscoelasticLiterature [55]In-houseResistance, PressureNon-linear numerical model
FERRANDEZ et al. (2002) [56]1DCoWopenrigidSimulated pressureFluent, ANSICFlowLiterature [57]
FITCHETT et al. (1991) [11]1DCoW, Heart,closedviscoelasticNumerically assessedn/aPressure, Stroke work,n/a
  Full body    Impulse reponse 
FORMAGGIA et al. (2003) [58]1DEndograftopenelasticPressure waveformsFinite element Taylor–Galerkin schemeFlow, Pulse waveforms2D fluid–structure interaction
HUANG et al. (2015) [59]1DCoW, Full bodyclosedelasticn/aTVD schemeVelocity, Pressure, FlowLiterature [60]
HUANG et al. (2018) [61]1DCoWopenelasticMRI/CT measurementsTHINkSFlow, VelocityLiterature [60]
OLUFSEN et al. (2000) [62]1DFull bodyopenelasticMRI measurementsTwo-step Lax–Wendroff methodAlgorithms, PressureMRI measurements
PARK et al. (2019) [63]1DCoWopenelasticMRI measurementsMethod of weighted residualsPressure, Flow, AreaMRI measurements
      Fourier series  
PERDIKARIS et al. (2015) [64]1DCoW, ArmclosedviscoelasticMRI/PCMRI measurementsDiscontinuous Galerkin solverPressure, Flown/a
REYMOND et al. (2009) [65]1DCoW, Heart,openviscoelasticMRI/Doppler measurements,Witzig–Womersley TheoryPressure, Flow, AreaMRI/Doppler measurements,
  Full body  Tonometry  Tonometry
RYU et al. (2015) [66]1DCoW, HeartopenelasticLiterature [23,35,36,67]Second-order spatial schemeFlow velocity, Activation factor,Simulations
      Two-step Adams–Bashforth temporal schemeActive tension, Flow 
BERG et al. (2014) [68]3DCoWopenrigidMRI measurementsANSYS FluentVelocity fields, Angular similarity index,n/a
       Magnitude similarity index 
CEBRAL et al. (2003) [69]3DCoWopenrigidMRI/PCMRI measurementsIn-house simulationFlow, Resistances,n/a
       Shear parameters (OSI, WSS) 
CHAICHANA et al. (2011) [70]3DCoronary arteriesopenrigidLiterature [71]ANSYS CFX, MATLABWSS, WSS gradient, Velocityn/a
FABBRI et al. (2014) [72]3DCoWopenrigidMRI measurementsANSYS CFX, Particle–fluid interactionParticle trackingn/a
FAN et al. (2009) [73]3DCarotidopenrigidLiterature [74]FLUENT, GAMBITVelocity, WSSn/a
GHAFFARI et al. (2017) [75]3DCoWopenrigidMRI measurements,ANSY FluentMorphology, Wall shear,Statistical analysis
     NOVA scans Vorticity, Pressure, Flow 
GRINBERG et al. (2011) [76]3DCoWopenrigidPC-MRI measurementsHigh-order spectral/hp element solverPressure, Flow distributionn/a
MAMATYUKOV et al. (2018) [77]3DCoWopenelasticDoppler measurementsANSYS CFXVelocityn/a
PISKIN et al. (2015) [78]3DCoWopenrigidLiterature [79]ANSYS FluentVelocity, WSS, PressureLiterature [79]
RAZAVI et al. (2014) [80]3DCoWopenrigidMRI measurementsCOMSOLVelocity, Flow, Shear raten/a
REN et al. (2015) [81]3DCoWopenrigidMRI/PCMRI measurementsANSYS FluentVelocity, Flown/a
ZHANG et al. (2020) [82]3DCoWopenrigidMRI measurementsANSYS CFXVelocity, Shear parametersn/a
ZHU et al. (2018) [83]3DCoWopenrigidMRI/PCMRI measurements,ANSYS CFXPressure drop, WSS, FlowNOVA data
DEVAULT et al. (2008) [84]0D-1DCoWopenviscoelasticMRI measurementsChebyshev collocation methodsVelocity, Pressure, FlowTranscranial Doppler
FORMAGGIA et al. (2006) [85]0D-1DHeart and arteriesopenviscoelasticBlood pressure measurementsSecond-order Taylor–Galerkin FE schemeFlow, Pressuren/a
LIANG et al. (2009) [86]0D-1DCoW, Full bodyclosedelasticLiterature [35,87,88]Two-step Lax–Wendroff methodBlood flow, PressureLiterature [89]
      Fourth-order Runge–Kutta methodStroke volume 
LIANG et al. (2009) [90]0D-1DCoW, Full bodyclosedelasticLiterature [35,87,88]Two-step Lax–Wendroff methodBlood flow, Pressuren/a
      ‘Ghost- point’ methodStroke volume 
LIANG et al. (2011) [37]0D-1DCoW, HeartclosedelasticIn vivo data, Literature [23,65,84]Two-step Lax–Wendroff methodFlow, HyperfusionIn vivo data,
      Fourth-order Runge–Kutta method Literature [23,65,84]
MÜLLER et al. (2014) [60]0D-1DTotalclosedelasticLiterature [37]Dumbser–Enaux–Toro (DET) solverPressure, FlowMRI measurements,
        Literature [91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109]
WÉBER et al. (2023) [110]0D-1DCoW, Heart,openPoynting–Literature [65,111,112]C++Pressure, FlowLiterature [113]
  Full body Thomson    
ZHANG et al. (2016) [114]0D-1DCoW, Heart,closedelasticIn-house software „V-modeler“ [115],In-house simulationFlowLiterature [23,35,65,116]
  Full body  MRI/PC-MRI measurements,  
     SPECT data   
KASHEFI et al. (2014) [117]0D-3DCarotid, Heart,openelasticLiterature [74]Matlab Simulink, ANSYS FluentVelocity distribution,Steady simulation [118]
  Full body    Pressure, Flow 
LIU et al. (2016) [119]0D-3DCoW, AortaopenrigidMRI/Doppler measurementsComsol MultiphysicsVelocity, FlowDoppler measurements
MORBIDUCCI et al. (2010) [120]0D-3DCarotidopenrigidCTA (angiographic images)ANSYS FluentWall shear parametersn/a
SUN et al. (2023) [121]0D-3DCoWclosedrigidCTAANSYS CFX
HO et al. (2009) [122]1D-3DCoWopenrigidCTAANSYS CFXVelocityDoppler measurements
LIANG et al. (2016) [123]1D-3DCoW, IAopenrigidMRI measurementsIn-house code,
Two-Step Lax-Wendroff method
Velocity, Flow, OSI, WSSPIV measurements
MARZO et al. (2011) [124]1D-3DCoWopenelastic3DRA/PC-MRI measurementsANSYS CFXFlow, Shear Parametersn/a
NEIDLIN et al. (2016) [125]1D-3DCoWopenrigidCTAANSYS CFXFlow, VelocityLiterature [126,127,128]
PASSERINI et al. (2009) [129]1D-3DCoW, CarotidopenrigidLiterature [23,30,130]LifeVFlow, Velocityn/a
ŠUTALO et al. (2014) [131]1D-3DCoWopenelasticCTANSYS CFXFlow, Pressure, Velocityn/a
URQUIZA et al. (2006) [132]1D-3DFull bodyopenelasticLiterature [74]SUPG [133]Flow, Velocity, Shear stressLiterature
BLANCO et al. (2010) [17]0D-1D-3DFull body, AortaclosedviscoelasticMRI measurementsGauss–Seidel methodFlow, OSI, Particle trackingn/a
LIANG et al. (2015) [134]0D-1D-3DCoW, Full bodyopenrigidLiterature [86,90]Two-step Lax–Wendroff methodWSS, Flow, Velocity3D simulation
      Fourth-order Runge–Kutta  
OSHIMA et al. (2012) [135]0D-1D-3DCoWopenrigidMRI measurementsIn-houseWSS, Pressure, Flown/a
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Korte, J.; Klopp, E.S.; Berg, P. Multi-Dimensional Modeling of Cerebral Hemodynamics: A Systematic Review. Bioengineering 2024, 11, 72. https://doi.org/10.3390/bioengineering11010072

AMA Style

Korte J, Klopp ES, Berg P. Multi-Dimensional Modeling of Cerebral Hemodynamics: A Systematic Review. Bioengineering. 2024; 11(1):72. https://doi.org/10.3390/bioengineering11010072

Chicago/Turabian Style

Korte, Jana, Ehlar Sophie Klopp, and Philipp Berg. 2024. "Multi-Dimensional Modeling of Cerebral Hemodynamics: A Systematic Review" Bioengineering 11, no. 1: 72. https://doi.org/10.3390/bioengineering11010072

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop