Ultrasonography Search

CLOSE


Lee: Hemodynamics in Doppler ultrasonography

Abstract

The cardiovascular system operates through complex hemodynamic processes involving pulsatile blood flow, varying viscosity, and the branching architecture of vessels. Interactions between blood flow and the vascular wall, which are characterized by shear and normal stress, along with wall stiffness, are crucial for maintaining vascular health. Doppler ultrasonography is a highly valuable noninvasive tool for assessing these hemodynamic parameters, enabling the measurement of key indices such as blood flow velocity, flow patterns, wall shear stress, and wall stiffness. This paper emphasizes the clinical significance of these indices and methods of measuring them using Doppler ultrasonography while addressing potential challenges. Accurate interpretation of these measurements is vital for reliable cardiovascular diagnostics and effective clinical decision-making.

Introduction

The cardiovascular system ensures adequate blood flow according to the organism’s metabolic demands. Cardiac contractions drive blood from the heart into the aorta and back through the vena cava. Each step in this pathway exhibits unique hemodynamic characteristics. Imaging methods for in vivo hemodynamic evaluation include ultrasonography and magnetic resonance imaging (MRI). Ultrasonography uses the Doppler principle to visualize and analyze moving blood cells qualitatively and quantitatively. Doppler ultrasonography (DUS) is a highly valuable tool for the noninvasive assessment of blood flow, although ultrasound penetration may be limited in some body regions. This article examines in vivo blood flow hemodynamics through ultrasonographic features.

Blood Flow Physiology

Hemodynamic Function

Blood flow starts in a pulsatile pattern due to the periodic contractions of the left ventricle and the one-way valve action at the ventriculo-aortic junction. The aorta and arteries dampen this pulsation to minimize pressure damage in organs. Their thicker, elastic walls expand and recoil, absorbing the pulsatile energy. However, this pulsatile nature means that simple observations of a vessel's luminal patency may not reflect actual blood flow. Structures with stiffer walls, such as bypass grafts, may even exhibit reversed flow during diastole due to this pulsation [1].
Whole blood, unlike plasma, is a non-Newtonian fluid. Newtonian fluids have constant viscosity, meaning that their resistance to flow increases linearly with increasing shear rate (the difference in velocity between adjacent layers) [2]. In contrast, the viscosity of whole blood depends on several factors, including hematocrit, the aggregation of red blood cells (RBCs), and their deformability. At higher shear rates (caused by smaller vessels or higher flow velocity), RBC aggregation decreases, and the cells deform more, leading to decreased viscosity. This is referred to as shear thinning [3]. Conversely, in slow flow conditions, RBCs aggregate into rouleaux, increasing their resistance to flow and leading to greater viscosity.
Blood vessels have a branching (arteries) or converging (veins) tree-like structure, inherently leading to turbulent flow. Turbulent flow is characterized by chaotic and unpredictable velocity and pressure patterns. It occurs when the kinetic energy of the fluid overcomes the dampening effect of viscosity. Laminar flow, with smooth and predictable layers, is more likely in viscous fluids. The Reynolds number, which reflects the ratio of inertia to viscosity, is used to estimate turbulence in continuous flow. For pulsatile blood flow, the Womersley number is used to account for pulsation [4-6].
Similar to other colloidal suspensions, blood flow exhibits the Fåhræus-Lindqvist effect. As friction at the vessel wall increases due to smaller vessel size, RBCs migrate towards the center, leaving a cell-free zone near the wall. This reduces friction and potentially lowers the overall viscosity of blood in smaller vessels [7].
Among the characteristics of blood flow, the most unique feature of whole blood is its inconsistent dynamic viscosity.

Vascular Wall Function

The vascular system exhibits distinctive features in terms of both blood flow and wall function. The vascular wall is dynamic and capable of significantly influencing blood flow. Shear stress on the vascular wall is crucial for both wall structure and blood flow. Shear stress refers to the frictional force on the vascular endothelium in the direction of blood flow. It is calculated as the product of the wall shear rate and blood viscosity. The wall shear rate describes the velocity gradient between the outermost blood flow layer and the adjacent vessel wall. Increased friction at the vessel wall, caused by faster flow or a smaller vessel diameter, leads to higher wall shear stress. Elevated wall shear stress dissipates flow energy, stimulates endothelial cells, induces vasodilation, and may alter cell arrangement [8].
The vascular wall is subjected to both shear stress and normal stress, which comprises three components: circumferential stress (loop stress, tangential stress, wall stress, or wall tension), radial stress, and axial (longitudinal) stress. Normal stress arises primarily from the intraluminal pressure exerted by the contained blood mass. In contrast to shear stress, which is generated by the kinetic energy of blood flow, normal stress originates from static pressure. Radial stress acts perpendicular to the wall of a cylindrical blood vessel and influences its dilation. Longitudinal stress acts along the long axis of the vessel. In vessels such as the descending aorta, when a person is in a standing position, longitudinal stress can increase pressure distally due to gravity. This may explain the higher prevalence of aneurysms in the infrarenal abdominal aorta than in the thoracic descending aorta. Circumferential stress (wall tension) plays a critical role in vascular rupture, particularly in aneurysms. While circumferential stress primarily threatens wall integrity, all normal stress components influence each other and contribute to the overall wall tension. Wall tension can be calculated using the Young-Laplace equation, which relates it to a ratio between pressure and wall thickness [9,10].
Wall stiffness is another important aspect of wall function, and it can be expressed using parameters such as the tangential modulus (derived from the stress-strain curve), compliance, and wall tension [11]. Compliance—a classic parameter—refers to the rate of change of volume with respect to pressure changes. In vascular structures, however, the pressure-volume (or area) relationship is non-linear. As intraluminal pressure increases, the vessel lumen expands in a sigmoid curve. Compliance and the tangential modulus exhibit corresponding changes along this curve, with initial increases followed by decreases. Additionally, vascular dilation can decrease compliance by causing an upward shift in the pressure-volume curve. Conversely, vascular wall hypertrophy can delay the peak point of the compliance curve [12].
Since vascular wall compliance can be influenced by intraluminal pressure, vessel size, and wall properties, more consistent parameters have been introduced to represent wall stiffness. These include the elastic modulus, stiffness index, arterial distensibility index, and arterial wall stiffness index [13-16].

Physics and Techniques of DUS

Ultrasonography offers various imaging modes that provide different types of information. In M-mode (motion mode), a single ultrasound beam projects into the body, displaying a one-dimensional representation of movement over time. This mode is useful for highly precise evaluations of the motion of structures such as arterial walls and cardiac valves (Fig. 1A). B-mode (brightness mode) uses a linear array transducer to transmit a one-dimensional ultrasound beam, generating a linear signal map that is assembled into a two-dimensional image to evaluate the anatomical features of tissues and organs (Fig. 1B).
The Doppler mode provides information about the velocity of moving reflectors, such as blood cells, displaying a time-velocity spectrum that allows for evaluating quantitative velocities and blood flow patterns. The duplex mode combines B-mode and Doppler modes, enabling real-time visualization of color-coded velocity information superimposed on the B-mode image. In practice, a combination of the duplex image and the time-velocity spectrum is commonly used for a comprehensive analysis (Fig. 1C). Power Doppler mode, a variation of duplex mode, overlays an amplitude signal from moving reflectors on a B-mode background image. This signal reflects the amount of moving blood cells without providing specific velocity information, making it advantageous for detecting low-velocity flow with minimal noise artifacts [17].
DUS measures blood flow velocity using the Doppler principle, which states that the frequency of a reflected wave changes based on the relative motion between the source ultrasound beam and the reflector—typically, a blood cell in the bloodstream. The Doppler equation relates the measured frequency shift (Δf) to the input ultrasound frequency (f0), reflector velocity (v), angle of the ultrasound beam (θ), and the speed of sound in tissue (c):
f=2*f0*v*cosθ/2
Using the measured frequency shift, the velocity of the moving blood cell can be calculated [18]. In a duplex image, velocity information is overlaid onto the B-mode image. The flow direction is typically indicated by color, with red and blue representing incoming and outgoing blood flow, respectively. Color saturation and signal intensity reflect blood flow velocity and density, respectively.
Several parameters need adjustment for optimal Doppler image acquisition. (1) Although convex probes offer software-based beam deviation correction, linear probes are generally preferred due to their simpler beam geometry, which provides more accurate results. Higher-frequency probes offer better spatial resolution but shallower penetration depth. The probe frequency should be chosen based on target depth and desired detail. (2) The imaging depth is adjusted to encompass the target structure while excluding unnecessary deep structures, allowing the use of a higher frequency if the probe has automatic frequency adjustment. (3) In duplex mode, the focal depth is set on the target structure for optimal image quality. In Doppler spectrum imaging, focal depth control is unnecessary because the sampling volume coincides with the focal depth. (4) The Doppler angle refers to the angle between the incident ultrasound beam and the body surface. Doppler window steering is used to adjust the ultrasound beam direction for better signal acquisition. In velocity spectrum mode, the transverse line of the crosshair is aligned with the target blood flow direction for proper angular correction. A Doppler angle of 0° provides the best frequency shift detection, but this angle is difficult to achieve in vivo. A Doppler angle of 30°-60° is recommended for practical purposes. (5) The sampling volume is positioned on the target blood flow, and its size is adjusted. A small sampling volume at the center of the flow is preferred for evaluating blood flow patterns or peak velocity, while a larger sampling volume covering the entire blood flow width is preferable for flow rate measurements. (6) The velocity scales on the y-axis on the color map and the velocity spectrum are adjusted appropriately. A lower scale offers better signal resolution in pulsed-wave Doppler mode but can introduce aliasing artifacts. The scale should be reduced until just before aliasing artifacts appear. (7) The baseline is adjusted to display the entire height of the pulsatile spectrum. If aliasing artifacts persist after adjusting the baseline and scale, a lower-frequency probe should be used.
The velocity spectrum in Doppler mode represents the velocity of all reflectors at each time point on the x-axis. The highest value at each point generates a time-peak velocity curve, allowing measurement of peak systolic and end-diastolic velocities. The mean value of all plots at each time point creates a time-mean velocity curve, and averaging these mean velocities over a specific time interval allows calculation of the time-averaged mean velocity. Multiplying this by the cross-sectional area of the blood flow provides the mean flow rate of the corresponding blood vessel.
Laminar flow, where blood cells travel in a well-defined stream, shows densely clustered plots at each time point in the velocity spectrum. Turbulent flow exhibits more diverse velocity vectors, resulting in more dispersed plots; this pattern is referred to as "spectral broadening." Severe turbulence can obscure the spectral window, which is a signal-free area at the bottom of the velocity spectrum (Figs. 1D, 2B) [19].
Continuous wave Doppler (CWD) transmits a continuous ultrasound beam and receives reflected ultrasound waves simultaneously, which is useful for measuring high-velocity flow without aliasing artifacts, such as in stenotic jets within the heart and aorta. However, CWD cannot focus on a specific depth and simultaneously acquires all signals within the beam path, potentially leading to biased quantitative measurements [20].
DUS is the gold standard for evaluating blood flow due to its ability to acquire velocity directly in real time. MRI using phase-contrast sequences offers an alternative method for blood flow assessment. Compared to Doppler, MRI has advantages, including the absence of limitations in imaging plane and evaluation window and less operator variability [21]. However, MR phase-contrast imaging shows a more blunted pulsatile flow pattern and underestimates peak systolic and end-diastolic velocities, resulting in significantly underestimated resistance or pulsatility indices compared to DUS [18].

Measurement Indices of DUS

Blood Flow Velocity

DUS is used to evaluate blood flow by directly measuring velocity, the primary hemodynamic parameter. Other hemodynamic parameters derive from this velocity data. In duplex mode, velocity appears as a color-coded signal on a grayscale B-mode image. In the time-velocity spectrum, velocity is shown by the spectrum's height at each time point. Spectral waveform analysis provides values such as peak systolic velocity, end-diastolic velocity, and time-averaged mean velocity.
Blood vessels exhibit variations in flow velocity and pulsatile patterns due to physiological differences (Fig. 1C, D). According to Bernoulli's principle, flow velocity increases as a vessel's cross-sectional area decreases. This principle enables the use of DUS to assess vascular stenosis by analyzing velocity changes. For example, carotid artery stenosis can be graded based on peak systolic and end-diastolic velocity measurements in the internal and common carotid arteries (Table 1, Fig. 2C) [22]. Doppler stenosis grading correlates well with computed tomography angiography measurements [23]. Duplex mode can also estimate stenosis severity based on the geometry of the blood flow, although it may underestimate severity due to limitations in capturing marginal flow signals. Despite these limitations, DUS remains the standard for velocity-based stenosis evaluation of the carotid artery [24]. Carotid DUS screening has also been validated for follow-up in high-risk patients with atherosclerosis to predict cerebrovascular events [25].
The clinical applications of DUS extend beyond carotid artery assessment. Park et al. [26] used DUS-based central retinal artery velocimetry to identify ocular ischemic disease in patients with sudden vision loss. A mean velocity threshold of 4.3 cm/s offered high sensitivity (89%) and specificity (95%) for predicting significant vision loss. Espahbodi et al. [27] evaluated lumbar arterial blood flow in patients with lower back pain, finding higher peak systolic flow velocity in patients than in controls. This suggests that DUS is useful for assessing the hemodynamic changes associated with lower back pain. These examples highlight the diverse applications of DUS in assessing blood flow velocity across various vascular territories.

Blood Flow Rate

The flow rate measures the volume of blood flowing through a vessel per unit time. Using DUS, the flow rate is calculated by multiplying the time-averaged mean velocity by the cross-sectional area of the blood vessel and a conversion factor (typically 60 seconds for mL/min). This technique is widely used to assess blood flow in various parts of the body (Fig. 2).
Dark and Singer [28] validated transesophageal DUS for measuring cardiac output against the pulmonary artery thermo-dilution technique. Jung et al. [29] measured flow rates in four major cerebral arteries, finding significantly reduced total cerebral blood flow volume (TCBFV) in patients with acute ischemic stroke and decreased cerebral vascular reserve compared to healthy individuals. Kalfaoglu et al. [30] compared TCBFV between chronic smokers and non-smokers, revealing significantly lower TCBFV in smokers, suggesting that smoking exerts a negative impact on cerebral blood flow. Araz Server et al. [31] linked reduced vertebral artery flow to positional vertigo, finding that patients with undiagnosed positional vertigo had lower flow rates depending on head position. Mutlu et al. [32] found significantly lower peak systolic velocity and flow rate in patients with idiopathic sudden sensorineural hearing loss (ISSNHL), implying that impaired vertebral artery flow may play a role in this condition.
DUS is also valuable for monitoring arteriovenous malformations (AVMs), which are birth defects where arteries and veins connect abnormally, bypassing capillaries. DUS is used to diagnose AVM by assessing both blood flow patterns and the flow rate, with features of increased peak systolic velocity, low-impedance pattern, and slanted linear downslope (saw-tooth appearance). Monitoring changes in the flow rate and peak systolic velocity in feeding arteries is a method for tracking AVM progression or improvement [33]. Mohammadkarim et al. [34] used DUS to monitor common carotid artery flow rate changes after external radiotherapy for cancer treatment; they found a significant decrease in flow rate, suggesting the potential of DUS for predicting late radiotherapy-related side effects.
These examples highlight the versatility of DUS in evaluating the blood flow rate across various clinical scenarios.

Blood Flow Patterns

DUS is highly useful for analyzing blood flow patterns by capturing velocity variations over time. Arterial flow is pulsatile due to the heart's intermittent pumping, and it is influenced by inflow properties and downstream resistance. In high-resistance environments, the systolic peak velocity is much higher than the diastolic velocity. In complex pulsatile flow, "impedance" (the complex-valued generalization of resistance) describes the flow resistance. In this context, flow with a larger systolic-diastolic velocity difference is described as "high-impedance flow" [35]. Venous flow is typically continuous with some pulsatile variations due to respiratory and cardiac motion, especially in larger veins closer to the chest. Capillary flow shows a steady laminar pattern.
Arteries supplying the brain have lower impedance and a more continuous flow pattern than those supplying other organs. Arteries supplying skeletal muscles show a high-impedance flow pattern with biphasic or triphasic waveforms at rest. Arteries supplying visceral organs have an intermediate-impedance flow pattern [19].
Upstream of stenosis—that is, in a pre-stenotic segment—the flow pattern is high-impedance. Blood flow through the stenotic channel creates a high-velocity jet with turbulence. Downstream, the flow pattern exhibits lower impedance and becomes more continuous (Fig. 3). In cases of severe stenosis, post-stenotic flow can exhibit a weak and delayed pulsatile pattern, known as "pulsus tardus et parvus" [36].
DUS assesses relative impedance using flow velocity data. The resistance index (RI) is defined as the ratio between the systolic-diastolic velocity difference and peak systolic velocity. This parameter represents downstream impedance, but may underestimate it in large arteries with high diastolic velocity [37]. The pulsatility index (PI) normalizes velocity differences by vessel size, calculated as the ratio between the systolic-diastolic velocity difference and time-averaged mean velocity, and is also a practical parameter (Fig. 3D) [38].
Zeller et al. [39] reported that a bilateral renal artery RI difference exceeding 0.05 was highly specific for identifying significant stenosis in one renal artery. Mutlu et al. [32] found increased RI in the vertebral artery, suggesting flow disturbance in patients with ISSNHL. Mohammadkarim et al. [34] observed significant increases in the RI and PI of the common carotid artery after radiotherapy for cancer treatment.
Due to the pulsatile blood flow, DUS can evaluate both the upstream and downstream environments of blood flow by analyzing the flow patterns.

Flow Energy

The arterial system's complexity, with varying vessel diameters and branching patterns, requires efficient blood flow management to deliver oxygen and nutrients to organs. Blood flow energy combines the intraluminal pressure gradient and flow rate, which are analogous to electrical voltage and current.
Significant stenosis reduces blood flow energy due to increased friction within the narrowed segment and upstream flow diversion. The pressure gradient across the stenosis is a crucial parameter for assessing blood supply adequacy. Although DUS measures velocity rather than absolute pressure, various methods have been used to estimate pressure gradients based on velocity data.
Bernoulli's principle states that the total energy (kinetic, potential, and intrinsic) of a moving fluid remains constant. Intrinsic energy refers to the static pressure exerted by blood within the vessel, kinetic energy represents blood flow movement, and potential energy relates to the position of blood relative to gravity. Dynamic pressure is defined as the kinetic energy per unit volume of blood [2,40]. Some DUS protocols estimate pressure gradients using a simplified version of the Bernoulli equation. This method assumes negligible stenotic flow acceleration, viscous friction, and zero inlet flow velocity [41]. While suitable for high-velocity jets from near-complete flow stoppage (e.g., in heart valve regurgitation), its accuracy for moderate arterial stenosis with continuous inflow is debated [42,43]. In vitro studies have shown that this method overestimates pressure gradients compared to direct pressure measurements, possibly due to neglecting post-stenotic pressure recovery [43]. DeGroff et al. [44] also found significant discrepancies between Doppler-derived and catheter-measured pressure gradients across modified Blalock-Taussig shunts.
More complex fluid dynamic models such as the Hagen-Poiseuille and Navier-Stokes equations have been explored for estimating pressure gradients. However, their requirements (e.g. a, Newtonian fluid and straight laminar flow) are not always met in vivo, and successful results have not been consistently achieved [45,46].
DUS provides valuable information on blood flow velocity, but has limitations in directly measuring pressure gradients. Simplified methods exist, but their accuracy can be limited. Further research is needed to develop reliable Doppler-based techniques for pressure gradient estimation in blood vessels.

Wall Shear Stress

Wall shear stress triggers endothelial cells to initiate vasodilation and tighten intercellular junctions. Oscillating wall shear stress widens intercellular gaps, potentially promoting atherosclerotic plaque development, particularly in regions like arterial bifurcation shoulders [47].
Measuring wall shear stress in vivo requires acquiring the dynamic viscosity and shear rate of blood flow. Dynamic viscosity is often assumed based on hematocrit, which correlates with blood shear yield in vitro [48]. Using duplex DUS images, Wang et al. [49] calculated the shear rate by measuring velocity differences between pixels near the vascular wall. Assuming constant viscosity, they derived wall shear stress, which was validated against computational fluid dynamics (CFD) simulations for spatial wall shear stress distribution along the arterial wall.
DUS is a convenient method for calculating wall shear stress using the formula "4×viscosity×maximum velocity/vessel diameter," assuming a parabolic (Poiseuille) velocity profile (Fig. 3D) [50]. However, this method's reliability is limited in curved, branching arteries due to non-axisymmetric and turbulent flow profiles [51]. Gates et al. [52] highlighted inaccuracies in DUS-based wall shear stress measurements. Mynard et al. [51] recommended adjustments using CFD techniques.
Despite challenges in accurately measuring the shear rate, simplified techniques for translating wall shear stress to clinical settings are under development but remain a matter of debate. Clinical utility and the significance of wall shear stress values and their circumferential distribution remain unclear. However, DUS-based wall shear stress techniques continue to intrigue clinicians due to their theoretical representation of wall stress via blood flow kinetic energy.

Wall Stiffness

Arterial wall stiffness results from the balance between collagen and elastin proportions and the smooth muscle tone in the tunica media. Smooth muscle tone can be influenced by nervous activity, hormones, vasoactive substances such as nitric oxide from endothelial cells, and drugs. Normally, arterial wall stiffness varies based on the physical properties of arteries at different sites. Under abnormal circumstances, arterial wall stiffness increases with aging, hypertension, and certain medications. These influences also vary depending on the location of the arteries. Increased arterial stiffness reduces the damping effect of pulsatile arterial flow produced by cardiac motion, causing arteriolar and capillary flow to become rougher in the downstream end-organ parenchyma and contributing to the development of arterial hypertension [8].
In the clinical field, various indices have been proposed to evaluate arterial wall stiffness, aiming to overcome the limitations of the primary parameter known as compliance, which varies with intraluminal pressure changes. Since blood pressure and arterial luminal diameter are input variables, DUS is a convenient modality for measuring arterial wall stiffness (Table 2). Although no single index has proven superior to others, a commonly used index in clinical practice and research is arterial distensibility. This parameter is defined as the ratio between arterial luminal diameter change and the product of pulse pressure and diastolic luminal diameter [14].
Ultrasonography is a convenient modality for measuring arterial wall stiffness by assessing the luminal diameter alongside blood pressure information acquired from brachial cuff manometry. Additionally, various indices have been developed to normalize different arterial sizes and blood pressures. However, these indices still have limitations, such as indirectly measured pulse pressure, low spatial resolution to detect subtle diameter changes, and variations in the physical properties of arteries based on their location (Fig. 1E).

Summary

Understanding blood flow in the human body is complex due to dynamic blood flow and vascular wall physiology. This understanding is crucial for the effective clinical use of ultrasonography. Ultrasonography offers real-time imaging and direct evaluation of blood flow velocity, providing duplex velocity information on B-mode images and time-velocity spectra for quantitative analysis. However, most hemodynamic indices derived from DUS depend on velocity data. Errors in velocimetry can lead to systematic errors in the results, underscoring the need for meticulous parameter settings to ensure reliable data acquisition during DUS.

Conflict of Interest

No potential conflict of interest relevant to this article was reported.

References

1. Leong CM, Nackman GB, Wei T. Flow patterns through vascular graft models with and without cuffs. PLoS One 2018;13:e0193304.
crossref
2. Batchelor GK. An introduction to fluid dynamics. Cambridge Mathematical Library series. Cambridge: Cambridge University Press, 2000.

3. Baskurt OK, Meiselman HJ. Blood rheology and hemodynamics. Semin Thromb Hemost 2003;29:435–450.
crossref
4. Eames I, Flor JB. New developments in understanding interfacial processes in turbulent flows. Philos Trans A Math Phys Eng Sci 2011;369:702–705.
crossref pdf
5. Falkovich G. Fluid mechanics. 2nd ed. Cambridge: Cambridge University Press, 2018.

6. Womersley JR. Method for the calculation of velocity, rate of flow and viscous drag in arteries when the pressure gradient is known. J Physiol 1955;127:553–563.
crossref
7. Maeda N. Erythrocyte rheology in microcirculation. Jpn J Physiol 1996;46:1–14.
crossref
8. Bank AJ, Kaiser DR. Smooth muscle relaxation: effects on arterial compliance, distensibility, elastic modulus, and pulse wave velocity. Hypertension 1998;32:356–359.
crossref
9. Harvey JF. Theory and design of modern pressure vessels. 2nd ed. New York: Van Nostrand Reinhold, 1974.

10. van der Vliet JA, Boll AP. Abdominal aortic aneurysm. Lancet 1997;349:863–866.
crossref
11. Di Martino ES, Bohra A, Vande Geest JP, Gupta N, Makaroun MS, Vorp DA. Biomechanical properties of ruptured versus electively repaired abdominal aortic aneurysm wall tissue. J Vasc Surg 2006;43:570–576.
crossref
12. Drzewiecki G, Field S, Moubarak I, Li JK. Vessel growth and collapsible pressure-area relationship. Am J Physiol 1997;273:H2030–H2043.
crossref
13. Selzer RH, Mack WJ, Lee PL, Kwong-Fu H, Hodis HN. Improved common carotid elasticity and intima-media thickness measurements from computer analysis of sequential ultrasound frames. Atherosclerosis 2001;154:185–193.
crossref
14. O'Rourke MF, Staessen JA, Vlachopoulos C, Duprez D, Plante GE. Clinical applications of arterial stiffness: definitions and reference values. Am J Hypertens 2002;15:426–444.
crossref
15. Nollen GJ, Westerhof BE, Groenink M, Osnabrugge A, van der Wall EE, Mulder BJ. Aortic pressure-area relation in Marfan patients with and without beta blocking agents: a new non-invasive approach. Heart 2004;90:314–318.
crossref
16. Seo J, Choi D, Rienmueller R, Lim JG, Chang Y, Lee J. Suggestion for a new image-based aortic wall stiffness evaluation technique: arterial wall stiffness index. Int J Cardiovasc Imaging 2009;25 Suppl 1:83–94.
crossref pdf
17. Martinoli C, Pretolesi F, Crespi G, Bianchi S, Gandolfo N, Valle M, et al. Power Doppler sonography: clinical applications. Eur J Radiol 1998;27 Suppl 2:S133–S140.
crossref
18. Jiang J, Strother C, Johnson K, Baker S, Consigny D, Wieben O, et al. Comparison of blood velocity measurements between ultrasound Doppler and accelerated phase-contrast MR angiography in small arteries with disturbed flow. Phys Med Biol 2011;56:1755–1773.
crossref
19. Chavhan GB, Parra DA, Mann A, Navarro OM. Normal Doppler spectral waveforms of major pediatric vessels: specific patterns. Radiographics 2008;28:691–706.
crossref
20. Bolger AF, Eidenvall L, Ask P, Loyd D, Wranne B. Understanding continuous-wave Doppler signal intensity as a measure of regurgitant severity. J Am Soc Echocardiogr 1997;10:613–622.
crossref
21. Fraser KH, Poelma C, Zhou B, Bazigou E, Tang MX, Weinberg PD. Ultrasound imaging velocimetry with interleaved images for improved pulsatile arterial flow measurements: a new correction method, experimental and in vivo validation. J R Soc Interface 2017;14:20160761.
crossref pdf
22. Nicolaides AN, Shifrin EG, Bradbury A, Dhanjil S, Griffin M, Belcaro G, et al. Angiographic and duplex grading of internal carotid stenosis: can we overcome the confusion? J Endovasc Surg 1996;3:158–165.
crossref pdf
23. Rustempasic N, Gengo M. Assesment of carotid stenosis with CT angiography and color Doppler ultrasonography. Med Arch 2019;73:321–325.
crossref
24. Maroufi SF, Rafiee Alavi SN, Abbasi MH, Famouri A, Mahya N, Armaghan S, et al. Comparison of Doppler ultrasound and digital subtraction angiography in extracranial stenosis. Ann Med Surg (Lond) 2022;74:103202.
crossref
25. Schulte-Altedorneburg G, Droste DW, Felszeghy S, Csiba L, Popa V, Hegedus K, et al. Detection of carotid artery stenosis by in vivo duplex ultrasound: correlation with planimetric measurements of the corresponding postmortem specimens. Stroke 2002;33:2402–2407.
crossref
26. Park J, Lee HJ, Lee J. The feasibility of central retinal artery Doppler ultrasonography in patients with sudden visual disturbance that's caused by retinal vascular occlusion. J Korean Soc Radiol 2009;60:153–157.
crossref
27. Espahbodi S, Dore CJ, Humphries KN, Hughes SP. Color Doppler ultrasonography of lumbar artery blood flow in patients with low back pain. Spine (Phila Pa 1976) 2013;38:E230–E236.
crossref
28. Dark PM, Singer M. The validity of trans-esophageal Doppler ultrasonography as a measure of cardiac output in critically ill adults. Intensive Care Med 2004;30:2060–2066.
crossref pdf
29. Jung HY, Lee HJ, Kim HJ, Baik SK, Lee J, Kim YS, et al. Doppler ultrasonographic parameters for predicting cerebral vascular reserve in patients with acute ischemic stroke. J Korean Soc Ultrasound Med 2006;25:35–40.

30. Kalfaoglu ME, Hizal M, Kiyan A, Gurel K. The effects of chronic smoking on total cerebral blood volume measured by carotid and vertebral artery doppler ultrasonography. J Clin Ultrasound 2017;45:561–565.
crossref pdf
31. Araz Server E, Edizer DT, Yigit O, Yasak AG, Erdim C. Relationship between vertebral artery blood flow in different head positions and vertigo. Acta Otolaryngol 2018;138:1–5.
crossref
32. Mutlu A, Cam I, Dasli S, Topdag M. Doppler ultrasonography can be useful to determine the etiology of idiopathic sudden sensorineural hearing loss. Auris Nasus Larynx 2018;45:456–460.
crossref
33. Harders A, Bien S, Eggert HR, Laborde G, Merland JJ, Rufenacht D. Haemodynamic changes in arteriovenous malformations induced by superselective embolization: transcranial Doppler evaluation. Neurol Res 1988;10:239–245.
crossref
34. Mohammadkarim A, Mokhtari-Dizaji M, Kazemian A, Saberi H. Hemodynamic analysis of radiation-induced damage in common carotid arteries by using color Doppler ultrasonography. Ultrasonography 2018;37:43–49.
crossref pdf
35. Gatti PL, Ferrari V. Applied structural and mechanical vibrations: theory, methods, and measuring instrumentation. New York: E & FN Spon, 1999.

36. Li JC, Yuan Y, Qin W, Wang L, Dai Q, Qi ZH, et al. Evaluation of the tardus-parvus pattern in patients with atherosclerotic and nonatherosclerotic renal artery stenosis. J Ultrasound Med 2007;26:419–426.
crossref pdf
37. Keats TE, Sistrom C. Atlas de medidas radiologicas [Atlas of radiologic measurement]. Madrid: Harcourt, 2002.

38. Tortora GJ, Derrickson BH. Principles of anatomy and physiology. 13th ed. Hoboken, NJ: Wiley, 2012.

39. Zeller T, Bonvini RF, Sixt S. Color-coded duplex ultrasound for diagnosis of renal artery stenosis and as follow-up examination after revascularization. Catheter Cardiovasc Interv 2008;71:995–999.
crossref
40. Clancy LJ. Aerodynamics. London: Pitman, 1975.

41. Baumgartner H, Hung J, Bermejo J, Chambers JB, Evangelista A, Griffin BP, et al. Echocardiographic assessment of valve stenosis: EAE/ASE recommendations for clinical practice. J Am Soc Echocardiogr 2009;22:1–23.

42. Ahmed SN, Syed FM, Porembka DT. Echocardiographic evaluation of hemodynamic parameters. Crit Care Med 2007;35:S323–S329.
crossref
43. Hatle L, Brubakk A, Tromsdal A, Angelsen B. Noninvasive assessment of pressure drop in mitral stenosis by Doppler ultrasound. Br Heart J 1978;40:131–140.
crossref
44. DeGroff CG, Shandas R, Kwon J, Valdes-Cruz L. Accuracy of the Bernoulli equation for estimation of pressure gradient across stenotic Blalock-Taussig shunts: an in vitro and numerical study. Pediatr Cardiol 2000;21:439–447.
crossref pdf
45. Mynard JP, Steinman DA. Effect of velocity profile skewing on blood velocity and volume flow waveforms derived from maximum Doppler spectral velocity. Ultrasound Med Biol 2013;39:870–881.
crossref
46. McLean D. Understanding aerodynamics: arguing from the real physics. Chichester: Wiley, 2013.

47. Shaaban AM, Duerinckx AJ. Wall shear stress and early atherosclerosis: a review. AJR Am J Roentgenol 2000;174:1657–1665.

48. Picart C, Piau JM, Galliard H, Carpentier P. Human blood shear yield stress and its hematocrit dependence. J Rheol 1998;42:1–12.
crossref
49. Wang C, Chen M, Liu SL, Liu Y, Jin JM, Zhang YH. Spatial distribution of wall shear stress in common carotid artery by color Doppler flow imaging. J Digit Imaging 2013;26:466–471.
crossref pdf
50. Gnasso A, Carallo C, Irace C, De Franceschi MS, Mattioli PL, Motti C, et al. Association between wall shear stress and flow-mediated vasodilation in healthy men. Atherosclerosis 2001;156:171–176.
crossref
51. Mynard JP, Wasserman BA, Steinman DA. Errors in the estimation of wall shear stress by maximum Doppler velocity. Atherosclerosis 2013;227:259–266.
crossref
52. Gates PE, Gurung A, Mazzaro L, Aizawa K, Elyas S, Strain WD, et al. Measurement of wall shear stress exerted by flowing blood in the human carotid artery: ultrasound Doppler velocimetry and echo particle image velocimetry. Ultrasound Med Biol 2018;44:1392–1401.
crossref

A 76-year-old woman who underwent carotid Doppler ultrasonography (DUS) due to high cardiovascular risk.

A. The left common carotid artery (CCA) wall motion is quantitatively measured using M-mode. B. In the B-mode image, a partially calcified stable atherosclerotic plaque (crosshairs) is noted in the left carotid bulb without significant geometric stenosis. C. A duplex and Doppler spectral waveform shows normal patterned flow without abnormal velocities. D. The spectral waveform acquired from the left external carotid artery (ECA) shows higher resistance index (RI) and a clearer spectral window (*) than the internal carotid artery (ICA). As a facial muscular artery, the ECA shows a different flow pattern from cerebral arteries. E. DUSbased flowmetry shows no evidence of a decreased flow rate in the bilateral carotid and vertebral arteries. No stenosis is noted in the bilateral ICAs. The left CCA shows increased classic compliance, whereas other stiffness parameters are not increased (red square). This discrepancy is likely be due to the increased pulse pressure. AWSI, arterial wall stiffness index; BMI, body mass index; BP, blood pressure; BSA, bovine serum albumin; EDV, end-diastolic velocity; LCCA, left common carotid artery; LECA, left external carotid artery; LICA, left internal carotid artery; LICAd, distal level of left internal carotid artery; LICAi, imaginary diameter of left internal carotid artery; LT, left; LVA, left vertebral artery; PI, pulsatility index; PROX, proximal; PSV, peak systolic velocity; RCCA, right common carotid artery; RECA, right external carotid artery; RICA, right internal carotid artery; RICAd, distal level of right internal carotid artery; RICAi, imaginary diameter of right internal carotid artery; RVA, right vertebral artery; TAMV, timeaverage mean velocity; VA, vertebral artery; WSS, wall shear stress.
usg-24126f1.jpg
Fig. 1.

An 84-year-old woman who complained of left visual disturbance and underwent Doppler ultrasonography in the orbit and neck.

A. In B-mode, a calcified atherosclerotic plaque with an irregular surface is noted in the left carotid bulb and proximal internal carotid artery (ICA). Geometrically significant stenosis is noted (arrow). B. A duplex and Doppler spectral waveform shows increased peak systolic velocity and a spectral window effaced by turbulence (*). C. An in-house reporting form discloses geometric and hemodynamic stenosis in the left ICA (red square). The left ICA shows decreased flow rate, but the right ICA and left vertebral artery show compensatory hyperflow maintaining normal total cerebral blood flow volume (green square). D. The left central retinal artery shows attenuated post-stenotic pattern flow (continuous low-impedance slow flow), suggesting upstream flow disturbance. AWSI, arterial wall stiffness index; BMI, body mass index; BP, blood pressure; BSA, bovine serum albumin; CCA, common carotid artery; ECA, external carotid artery; EDV, end-diastolic velocity; HR, heart rate; LCCA, left common carotid artery; LECA, left external carotid artery; LICA, left internal carotid artery; LICAd, distal level of left internal carotid artery; LICAi, imaginary diameter of left internal carotid artery; LT, left; LVA, left vertebral artery; PI, pulsatility index; PROX, proximal; PSV, peak systolic velocity; RCCA, right common carotid artery; RECA, right external carotid artery; RI, resistance index; RICA, right internal carotid artery; RICAd, distal level of right internal carotid artery; RICAi, imaginary diameter of right internal carotid artery; RVA, right vertebral artery; TAMV, time-average mean velocity; VA, vertebral artery; WSS, wall shear stress.
usg-24126f2.jpg
Fig. 2.

A 79-year-old man who underwent Doppler ultrasonography to evaluate cerebral symptoms.

A. The right common carotid artery (CCA) shows a large difference between peak systolic velocity (PSV) and end-diastolic velocity (EDV). The resistance index (RI) is increased to 0.92, suggesting downstream flow disturbance. B. The right proximal internal carotid artery (ICA) shows increased flow velocity and an effaced spectral window, suggesting a turbulent stenotic jet flow. C. The right distal extracranial ICA shows high velocity but a lower impedance pattern than stenotic flow, suggesting a post-stenotic flow pattern. D. The report shows hemodynamic stenosis and an increased RI and pulsatility index (PI) of the right ICA (red square). AWSI, arterial wall stiffness index; BMI, body mass index; BSA, bovine serum albumin; BP, blood pressure; EDV, end-diastolic velocity; HR, heart rate; LCCA, left common carotid artery; LECA, left external carotid artery; LICA, left internal carotid artery; LICAd, distal level of left internal carotid artery; LICAi, imaginary diameter of left internal carotid artery; LVA, left vertebral arte; PROX, proximal; PSV, peak systolic velocity; RCCA, right common carotid artery; RECA, right external carotid artery; RI, resistance index; RICA, right internal carotid artery; RICAd, distal level of right internal carotid artery; RICAi, imaginary diameter of right internal carotid artery; RT, right; RVA, right vertebral artery; TAMV, time-average mean velocity; VA, vertebral artery; WSS, wall shear stress.
usg-24126f3.jpg
Fig. 3.
Table 1.
Carotid arterial stenosis evaluation based on Doppler ultrasonographic velocimetry [22]
NASCET stenosis grade PSV EDV PSVICA/PSVCCA PSVICA/EDVCCA EDVICA/EDVCCA
1-9% <120 <40 <1.5 <7 <2.6
10-49% <150 <80 <2.0 <10 <2.6
50-59% <250 <130 <3.2 <10 <2.6
60-69% <250 >130 <4.0 <15 <5.5
70-79% >250 >130 >4.0 <25 <5.5
50-59% >250 >130 >4.0 >25 >5.5
90-99% Trickle flow
100% No flow

NASCET, North American Symptomatic Carotid Endarterectomy Trial; PSV, peak systolic velocity; EDV, end-diastolic velocity; ICA, internal carotid artery; CCA, common carotid artery.

Table 2.
Various indices of arterial wall stiffness [14]
Indices Definition Formula
Arterial distensibility Relative diameter (or area) change for a pressure increment; the inverse of elastic modulus ΔD/(ΔP·ΔD) (/mmHg)
Arterial compliance Absolute diameter (or area) change for a given pressure step at fixed vessel length ΔD/ΔP (cm/mmHg) or (cm2/mmHg)
Volume elastic modulus Pressure step required for (theoretical) 100% increase in volume where there is no change in length ΔP/(ΔV/V) (mmHg)=ΔP/(ΔD/D) (mmHg)
Elastic modulus The pressure step required for (theoretical) 100% stretch from resting diameter at fixed vessel length (ΔP·D/ΔD) (mmHg)
Young’s modulus Elastic modulus per unit area; the pressure step per square centimeter required for (theoretical) 100% stretch from resting length ΔP·D/(ΔD·h) (mmHg/cm)
Stiffness index Ratio of logarithm (systolic/diastolic pressures) to (relative change in diameter) ß=In (Ps/Pd)/[(Ds-Dd)/Dd]
TOOLS
METRICS
0
Crossref
0
Scopus
608
View
109
Download
Editorial Office
A-304 Mapo Trapalace, 53 Mapo-daero, Mapo-gu, Seoul 04158, Korea
TEL : +82-2-763-5627   FAX : +82-2-763-6909   E-mail : office@ultrasound.or.kr
About |  Browse Articles |  Current Issue |  For Authors and Reviewers
Copyright © Korean Society of Ultrasound in Medicine.                 Developed in M2PI
Zoom in Close layer