Murray's law

Last updated

In biophysical fluid dynamics, Murray's law is a potential relationship between radii at junctions in a network of fluid-carrying tubular pipes. Its simplest version proposes that whenever a branch of radius splits into two branches of radii and , then the three radii should obey the equation

Contents

If network flow is smooth and leak-free, then systems that obey Murray's law minimize the resistance to flow through the network. For turbulent networks, the law takes the same form but with a different characteristic exponent α.

Murray's law is observed in the vascular and respiratory systems of animals, xylem in plants, and the respiratory system of insects. In principle, Murray's law also applies to biomimetic engineering, but human designs rarely exploit the law.

Murray's law is named after Cecil D. Murray, a physiologist at Bryn Mawr College, who first argued that efficient transport might determine the structure of the human vascular system.

Assumptions

Murray's law assumes material is passively transported by the flow of fluid in a network of tubular pipes, [1] and that the network requires energy to maintain both flow and structural integrity. [2] Variation in the fluid viscosity across scales will affect the Murray's law exponent, but is usually too small to matter. [3]

At least two different conditions are known in which the cube exponent is optimal.

In the first, organisms have free (variable) circulatory volume. Also, maintenance energy is not proportional to the pipe material, but instead the quantity of working fluid. The latter assumption is justified in metabolically active biological fluids, such as blood. [4] It is also justified for metabolically inactive fluids, such as air, as long as the energetic "cost" of the infrastructure scales with the cross-sectional area of each tube; such is the case for all known biological tubules. [5]

In the second, organisms have fixed circulatory volume and pressure, but wish to minimize the resistance to flow through the system. Equivalently, maintenance is negligible and organisms with to maximize the volumetric flow rate. [6]

Although most derivations of Murray's law assume a steady state flow field, the same results apply for flow in tubes that have a moderate to small width, relative to the flow wavelength. [7]

Derivation in laminar, mass-conservative networks with non-negligible upkeep

Poiseuille flow in a cylinder of diameter h; the velocity field at height y is u(y). Flux de Poiseuille.png
Poiseuille flow in a cylinder of diameter h; the velocity field at height y is u(y).

Murray's original derivation uses the first set of assumptions described above. She begins with the Hagen–Poiseuille equation, which states that for fluid of dynamic viscosity μ, flowing laminarly through a cylindrical pipe of radius r and length l, the volumetric flow rate Q associated with a pressure drop Δp is

and the corresponding power consumed is [1]

Said pipe contains volume πlr2. If the maintenance power density is λ, then the total power consumed (from both flow and upkeep) is

Minimizing this quantity depends on precisely which variables the organism is free to manipulate, but the minimum invariably occurs when the two terms are proportional to each other. [8] In that minimal case, the proportionality determines a relationship between Q and r. Canceling common factors and taking a square root,

 

 

 

 

(1)

That is, when using as little energy as possible, the mass flowing through the pipe must be proportional to the cube of the pipe's radius. Since flow is leakless, the total flow rate into a junction must be the total flow rate out:

Substituting ( 1 ) then gives Murray's law with α=3. [9]

Exponents for other networks

If the network does not rely on transported material getting "swept up in the flow", but instead expects it to passively diffuse, then resistance to transport is minimized when α=2: that is,

The same law would apply to a direct-current electrical grid composed of wires of only one material, but varying diameter. [10] For turbulent flow, transport resistance is minimized when α=7/3; that is: [11] [12]

In general, networks intermediate between diffusion and laminar flow are expected to have characteristic exponents between 2 and 3, at least approximately. [13] [14]

Experimental tests

Murray's law has been verified in chicks; dog intestines and lungs; cat mesentery; and human intestines and lung capillaries. [15] [16] Mice genetically engineered to lack the blood-vessel-wall protein elastin have smaller and thinner blood vessels, but still obey Murray's law. [17]

In humans, large vessels, such as the aorta or trachea, do not appear to obey Murray's law, instead obeying a Murray's law with exponent close to 2. [16] But flow in those vessels is also partially turbulent, and so should exhibit an exponent nearer to 7/3 than to 3. [18]

The Murray law in plants and insects. Murray materials in leaf and insect with a parent pore (radius r0) connecting to many children pores (radius ri). .jpg
The Murray law in plants and insects.

Insects do not have a fully-fledged circulatory system, instead relying on passive diffusion through the haemocoel. For those networks, Murray's law predicts constant cross-sectional area, which is observed. [10]

The same arguments that imply Murray's law also imply that the distribution of tubules should exhibit a specific power law scaling with size. Plant xylem is known to exhibit that scaling except in scenarios where the passages double as structural supports. [19] [20]

History

The first phenomenon now recognized as Murray's law is Young's rule for circulatory systems, which states that two identical subcapillaries should combine to form a capillary with radius about 1.26≈32 times larger, and dates to the early 19th century. [21] Bryn Mawr physiologist Cecil D. Murray published the law's modern, general formulation in 1926, [22] [21] but it languished in a disciplinary no-man's-land for the next fifty years: too trivial for physicists and too complicated for biologists. [21] Interest in the law revived in the 1970s. [21]

Applications

In circulatory system governed by Murray's law with α=3, shear stress on vessel walls is roughly constant. Consequently, variations in shear stress are a sign of deviation from Murray's law; Rodbard and Zamir suggest that such variations stimulate homeostatic growth or contraction. [23]

Murray's law rarely applies to engineered materials, because man-made transport routes attempt to reduce flow resistance by minimizing branching and maximizing diameter. [24]

Murray materials

Zheng achieved Murray fractal branching through layering porous zinc oxide nanoparticle sheets. Murray materials.jpg
Zheng achieved Murray fractal branching through layering porous zinc oxide nanoparticle sheets.

Materials that obey Murray's law at the microscale, known as Murray materials, are expected to have favorable flow characteristics, but their construction is difficult, because it requires tight control over pore size typically over a wide range of scales. [14] [25]

Lim et al propose designing microfluidic "labs on a chip" in accord with Murray's law to minimize flow resistance during analysis. Conventional lithography does not support such construction, because it cannot produce channels of varying depth. [26]

Seeking long-lived lithium battery electrodes, Zheng et al constructed Murray materials out of layers of sintered zinc oxide nanoparticles. The evaporation rate of the dissolved zinc oxide solvent controlled the size of the pores in each layer; the network was then just layers of ZnO with different pore sizes placed atop each other. [14]

Because power plant working fluids typically funnel into many small tubules for efficient heat transfer, Murray's law may be appropriate for nuclear reactor design. [27]

Related Research Articles

In the study of heat transfer, Newton's law of cooling is a physical law which states that

The rate of heat loss of a body is directly proportional to the difference in the temperatures between the body and its environment.

In fluid dynamics, the Darcy–Weisbach equation is an empirical equation that relates the head loss, or pressure loss, due to friction along a given length of pipe to the average velocity of the fluid flow for an incompressible fluid. The equation is named after Henry Darcy and Julius Weisbach. Currently, there is no formula more accurate or universally applicable than the Darcy-Weisbach supplemented by the Moody diagram or Colebrook equation.

<span class="mw-page-title-main">Series and parallel circuits</span> Types of electrical circuits

Two-terminal components and electrical networks can be connected in series or parallel. The resulting electrical network will have two terminals, and itself can participate in a series or parallel topology. Whether a two-terminal "object" is an electrical component or an electrical network is a matter of perspective. This article will use "component" to refer to a two-terminal "object" that participates in the series/parallel networks.

Hemodynamics or haemodynamics are the dynamics of blood flow. The circulatory system is controlled by homeostatic mechanisms of autoregulation, just as hydraulic circuits are controlled by control systems. The hemodynamic response continuously monitors and adjusts to conditions in the body and its environment. Hemodynamics explains the physical laws that govern the flow of blood in the blood vessels.

Vascular resistance is the resistance that must be overcome to push blood through the circulatory system and create blood flow. The resistance offered by the systemic circulation is known as the systemic vascular resistance (SVR) or may sometimes be called by the older term total peripheral resistance (TPR), while the resistance offered by the pulmonary circulation is known as the pulmonary vascular resistance (PVR). Systemic vascular resistance is used in calculations of blood pressure, blood flow, and cardiac function. Vasoconstriction increases SVR, whereas vasodilation decreases SVR.

In thermodynamics, the heat transfer coefficient or film coefficient, or film effectiveness, is the proportionality constant between the heat flux and the thermodynamic driving force for the flow of heat. It is used in calculating the heat transfer, typically by convection or phase transition between a fluid and a solid. The heat transfer coefficient has SI units in watts per square meter per kelvin (W/m²K).

<span class="mw-page-title-main">Allometry</span> Study of the relationship of body size to shape, anatomy, physiology, and behavior

Allometry is the study of the relationship of body size to shape, anatomy, physiology and behaviour, first outlined by Otto Snell in 1892, by D'Arcy Thompson in 1917 in On Growth and Form and by Julian Huxley in 1932.

<span class="mw-page-title-main">Continuous stirred-tank reactor</span> Type of chemical reactor

The continuous stirred-tank reactor (CSTR), also known as vat- or backmix reactor, mixed flow reactor (MFR), or a continuous-flow stirred-tank reactor (CFSTR), is a common model for a chemical reactor in chemical engineering and environmental engineering. A CSTR often refers to a model used to estimate the key unit operation variables when using a continuous agitated-tank reactor to reach a specified output. The mathematical model works for all fluids: liquids, gases, and slurries.

Fluid mechanics is that branch of physics concerned with the mechanics of fluids and the forces on them. It has applications in a wide range of disciplines, including mechanical, aerospace, civil, chemical, and biomedical engineering, as well as geophysics, oceanography, meteorology, astrophysics, and biology.

The Womersley number is a dimensionless number in biofluid mechanics and biofluid dynamics. It is a dimensionless expression of the pulsatile flow frequency in relation to viscous effects. It is named after John R. Womersley (1907–1958) for his work with blood flow in arteries. The Womersley number is important in keeping dynamic similarity when scaling an experiment. An example of this is scaling up the vascular system for experimental study. The Womersley number is also important in determining the thickness of the boundary layer to see if entrance effects can be ignored.

In respiratory physiology, airway resistance is the resistance of the respiratory tract to airflow during inhalation and exhalation. Airway resistance can be measured using plethysmography.

The Fåhraeus–Lindqvist effect describes how the viscosity of a fluid, in this case blood, changes with the diameter of the tube it travels through. In particular there is a 'decrease in viscosity as the tube's diameter decreases'. This is because erythrocytes move over to the centre of the vessel, leaving only plasma near the wall of the vessel.

In nonideal fluid dynamics, the Hagen–Poiseuille equation, also known as the Hagen–Poiseuille law, Poiseuille law or Poiseuille equation, is a physical law that gives the pressure drop in an incompressible and Newtonian fluid in laminar flow flowing through a long cylindrical pipe of constant cross section. It can be successfully applied to air flow in lung alveoli, or the flow through a drinking straw or through a hypodermic needle. It was experimentally derived independently by Jean Léonard Marie Poiseuille in 1838 and Gotthilf Heinrich Ludwig Hagen, and published by Poiseuille in 1840–41 and 1846. The theoretical justification of the Poiseuille law was given by George Stokes in 1845.

In fluid dynamics, the Oseen equations describe the flow of a viscous and incompressible fluid at small Reynolds numbers, as formulated by Carl Wilhelm Oseen in 1910. Oseen flow is an improved description of these flows, as compared to Stokes flow, with the (partial) inclusion of convective acceleration.

The convection–diffusion equation is a combination of the diffusion and convection (advection) equations, and describes physical phenomena where particles, energy, or other physical quantities are transferred inside a physical system due to two processes: diffusion and convection. Depending on context, the same equation can be called the advection–diffusion equation, drift–diffusion equation, or (generic) scalar transport equation.

<span class="mw-page-title-main">Diffusion</span> Transport of dissolved species from the highest to the lowest concentration region

Diffusion is the net movement of anything generally from a region of higher concentration to a region of lower concentration. Diffusion is driven by a gradient in Gibbs free energy or chemical potential. It is possible to diffuse "uphill" from a region of lower concentration to a region of higher concentration, as in spinodal decomposition. Diffusion is a stochastic process due to the inherent randomness of the diffusing entity and can be used to model many real-life stochastic scenarios. Therefore, diffusion and the corresponding mathematical models are used in several fields beyond physics, such as statistics, probability theory, information theory, neural networks, finance, and marketing.

<span class="mw-page-title-main">Reynolds number</span> Ratio of inertial to viscous forces acting on a liquid

In fluid mechanics, the Reynolds number is a dimensionless quantity that helps predict fluid flow patterns in different situations by measuring the ratio between inertial and viscous forces. At low Reynolds numbers, flows tend to be dominated by laminar (sheet-like) flow, while at high Reynolds numbers, flows tend to be turbulent. The turbulence results from differences in the fluid's speed and direction, which may sometimes intersect or even move counter to the overall direction of the flow. These eddy currents begin to churn the flow, using up energy in the process, which for liquids increases the chances of cavitation.

In engineering, physics, and chemistry, the study of transport phenomena concerns the exchange of mass, energy, charge, momentum and angular momentum between observed and studied systems. While it draws from fields as diverse as continuum mechanics and thermodynamics, it places a heavy emphasis on the commonalities between the topics covered. Mass, momentum, and heat transport all share a very similar mathematical framework, and the parallels between them are exploited in the study of transport phenomena to draw deep mathematical connections that often provide very useful tools in the analysis of one field that are directly derived from the others.

<span class="mw-page-title-main">Entropy production</span> Development of entropy in a thermodynamic system

Entropy production is the amount of entropy which is produced during heat process to evaluate the efficiency of the process.

<span class="mw-page-title-main">Lumped parameter model for the cardiovascular system</span>

A lumped parameter cardiovascular model is a zero-dimensional mathematical model used to describe the hemodynamics of the cardiovascular system. Given a set of parameters that have a physical meaning, it allows to study the changes in blood pressures or flow rates throughout the cardiovascular system. Modifying the parameters, it is possible to study the effects of a specific disease. For example, arterial hypertension is modeled increasing the arterial resistances of the model.

References

Footnotes

  1. 1 2 Murray 1926a , p. 210: "Poisseuille's equation for the law of flow in cylindrical tubes may be taken as a very fair approximation giving the relationships required."
  2. Murray 1926b , p. 299: "The cost per second, E, of operating a section of blood vessel was taken to be the sum of two work terms: (1) the work done per second against friction by the flow of blood, given in ergs/sec. by the product pf, where p is the fall in pressure in dynes/cm., and f is the rate of blood flow in cc./sec., and (2) the work done per second in maintaining the mere volume of blood contained in the vessel, given also in ergs/sec. by the term bvol, where vol is the volume in cc., and b is the cost in ergs/sec. per cc. of blood."
  3. Sherman 1981, p. 447.
  4. Murray 1926a , p. 210: "Let b, then, be the cost of blood in ergs per second per cubic centimeter of whole blood of average composition (and let B be the cost in Calories per day per cc. of blood). There is, as far as I can see, nothing arbitrary about this step: it is certain that the maintenance of blood requires fuel. (The cost of blood may, however, be a complex account distributed among such factors as the small metabolism of blood itself, the cost of upkeep of all the constituents, perhaps especially of hemoglobin, the cost of the containing vessels, and the burden placed upon the body in general by the mere weight of blood.)
  5. Sherman 1981 , p. 437: "If the volume of vessel tissue is proportional to the square of the internal radius (as when wall thickness is proportional to r), then Murray's derivation can hold for a biological vasculature even when the flowing fluid itself is inert and nonliving, as in the airways of the lungs, where the vessels are filled with air (of no metabolic cost) rather than with blood."
  6. Sherman 1981, pp. 437–439.
  7. Painter, Edén & Bengtsson 2006.
  8. Murray 1926a , pp. 210, 212: equations (4) and (6). Murray 1926b , p. 304: "The primitive meaning of organization is contained in the simplest problem of maxima and minima: if xyz=E, the sum x+y+z, is a minimum when x=y=z=3E.". Sherman 1981 , p. 436: "Under these conditions, only the coefficient b changes by inclusion of the vessel wall; the overall derivation remains the same, and Murray's law still holds."
  9. Murray 1926a , p. 212: equation (7).
  10. 1 2 Sherman 1981, p. 446.
  11. Uylings 1977 , p. 514: equation (15) with the maximal value of j=5.
  12. Williams et al. 2008 , p. 60: equation (5.2).
  13. Uylings 1977 , p. 514: equation (15).
  14. 1 2 3 4 5 Zheng et al. 2017.
  15. Williams et al. 2008, p. 56.
  16. 1 2 Sherman 1981, pp. 440–445.
  17. Sather, Hageman & Wagenseil 2012.
  18. Uylings 1977 , p. 509: "In the upper airways of the lung, turbulent flow exists during inspiration, which changes to laminar flow in the lower airways. The blood flow in the aorta of, for instance, normal rabbits and humans is turbulent under some circumstances."
  19. McCulloh, Sperry & Adler 2003.
  20. McCulloh & Sperry 2006.
  21. 1 2 3 4 Sherman 1981, p. 432.
  22. Murray  1926a , 1926b.
  23. Sherman 1981, p. 449.
  24. Williams et al. 2008 , p. 56: "This is not strictly accurate; Murray's law geometry was derived to obtain minimum system-level power demand and if lowest resistance is required then the largest vessel diameter should be selected at all points." The quote misidentifies the Murray's law optimization  Murray's law does provide lowest resistance, assuming system volume is fixed  but correctly identifies that, when system volume can grow freely, choosing maximal diameters will lower resistance.
  25. Williams et al. 2008.
  26. Lim et al. 2003.
  27. Acharya, Sarmistha (10 April 2017). "Leaf vein structure might help improve battery life". International Business Times . Yahoo! News. Archived from the original on 11 April 2017.

Further reading