Murray's law

The name of the pictureThe name of the pictureThe name of the pictureClash Royale CLAN TAG#URR8PPP



Murray's law predicts the thickness of branches in transport networks, such that the cost for transport and maintenance of the transport medium is minimized. This law is observed in the vascular and respiratory systems of animals, xylem in plants, and the respiratory system of insects.[1][2][3][4] Its simplest version states that if a branch of radius rdisplaystyle rr splits into two branches of radii r1displaystyle r_1r_1 and r2displaystyle r_2r_2, then r3=r13+r23displaystyle r^3=r_1^3+r_2^3displaystyle r^3=r_1^3+r_2^3. Like Hagen-Poiseuille Law and Fick's Law, which were also formulated from a biological context, Murray’s law is a basic physical principle for transfer networks.[2][4]


Murray's law is also a powerful biomimetics design tool in engineering. It has been applied in the design of self-healing materials, batteries, photocatalysts, and gas sensors.[citation needed] However, since its discovery, little attention has been paid to exploit this law for designing advanced materials, reactors and industrial processes for maximizing mass or energy transfer to improve material performance and process efficiency.[3]


The original version of Murray's law is only applicable to mass-conservative transport in the network. There are generalizations for non-conservative networks, which describe effects such as chemical reactions and diffusion through the walls.[1]




Contents





  • 1 Murray's law in mass-conservative networks

    • 1.1 Derivation

      • 1.1.1 Laminar Flow


      • 1.1.2 Diffusion




  • 2 The generalized Murray’s law


  • 3 Murray materials


  • 4 References




Murray's law in mass-conservative networks


Murray's original analysis[5][6] is based on the assumption that the radii inside a lumen-based system are such that the work for transport and upkeep is minimized. Larger vessels lower the expended energy for transport, but increase the overall volume of blood in the system; blood being a living fluid and hence requiring metabolic support. Murray's law is therefore an optimisation exercise to balance these factors.


For ndisplaystyle nn child branches splitting from a common parent branch, the law states that:


r3=r13+r23+r33+...+rn3displaystyle r^3=r_1^3+r_2^3+r_3^3+...+r_n^3displaystyle r^3=r_1^3+r_2^3+r_3^3+...+r_n^3


where rdisplaystyle rr is the radius of the parent branch, and r1,…,rndisplaystyle r_1,ldots ,r_nr_1,ldots ,r_n are the radii of the child branches. This law is only valid for laminar flow, since its derivation uses the Hagen-Poiseuille equation as a measure for the transport work (see below). Williams et al. deduced the formula for turbulent flow:[3]


r(7/3)=r1(7/3)+r2(7/3)+r3(7/3)+...+rn(7/3)displaystyle r^(7/3)=r_1^(7/3)+r_2^(7/3)+r_3^(7/3)+...+r_n^(7/3)displaystyle r^(7/3)=r_1^(7/3)+r_2^(7/3)+r_3^(7/3)+...+r_n^(7/3)



Derivation


As stated above, the underlying assumption of the Murray's law is that the power (energy per time) for transport and upkeep is minimal in a natural transport system. Hence we seek to minimize P=Pt+Pmdisplaystyle P=P_t+P_mdisplaystyle P=P_t+P_m, where Ptdisplaystyle P_tP_t is the power required for transport and Pmdisplaystyle P_mP_m the power required to maintain the transport medium (e.g. blood).



Laminar Flow


We first minimize the power of transport and upkeep in a single channel of the system, i.e. ignore bifurcations. Together with the assumption of mass-conservation this will yield the law. Let Qdisplaystyle QQ be the laminar flow rate in this channel, which is assumed to be fixed. The power for transport in a laminar flow is Pt=QΔpdisplaystyle P_t=QDelta pdisplaystyle P_t=QDelta p, where Δpdisplaystyle Delta pDelta p is the pressure difference between the entry and exit of a tube of radius rdisplaystyle rr and length ldisplaystyle ll. The Hagen-Poiseuille Law for laminar flow states that Q=πr48μlΔptextstyle Q=frac pi ,r^48,mu ,lDelta ptextstyle Q=frac pi ,r^48,mu ,lDelta p and hence Δp=8μlπr4Qtextstyle Delta p=frac 8,mu ,lpi ,r^4Qtextstyle Delta p=frac 8,mu ,lpi ,r^4Q , where μdisplaystyle mu mu is the dynamic viscosity of the fluid. Substituting this into the equation for Ptdisplaystyle P_tP_t, we get



Pt=8μlπr4Q2.displaystyle P_t=frac 8,mu ,lpi ,r^4Q^2.

displaystyle P_t=frac 8,mu ,lpi ,r^4Q^2.
We furthermore make the assumption that the power necessary for maintenance is proportional to the volume of the cylinder V=πlr2displaystyle V=pi ,l,r^2displaystyle V=pi ,l,r^2 :

Pm=λV=λπlr2displaystyle P_m=lambda ,V=lambda ,pi ,l,r^2

displaystyle P_m=lambda ,V=lambda ,pi ,l,r^2
where λ>0displaystyle lambda >0lambda >0 is a constant number (metabolic factor). This yields

P=Pt+Pm=8μlπr4Q2+λπlr2.displaystyle P=P_t+P_m=frac 8,mu ,lpi ,r^4Q^2+lambda ,pi ,l,r^2.

displaystyle P=P_t+P_m=frac 8,mu ,lpi ,r^4Q^2+lambda ,pi ,l,r^2.
Since we seek the radius with minimal power, we calculate the stationary point with respect to r

ddrP=0⇔−32μlπr5Q2+2λπlr=0⇔Q2=π216λμr6⇐Q=π4λμr3.displaystyle operatorname d over operatorname d !rP=0;Leftrightarrow ;-frac 32,mu ,lpi ,r^5Q^2+2,lambda ,pi ,l,r=0;Leftrightarrow ;Q^2=frac pi ^216frac lambda mu r^6;Leftarrow ;Q=frac pi 4sqrt frac lambda mu r^3.

displaystyle operatorname d over operatorname d !rP=0;Leftrightarrow ;-frac 32,mu ,lpi ,r^5Q^2+2,lambda ,pi ,l,r=0;Leftrightarrow ;Q^2=frac pi ^216frac lambda mu r^6;Leftarrow ;Q=frac pi 4sqrt frac lambda mu r^3.
With this relationship between the flow rate Qdisplaystyle QQ and radius rdisplaystyle rr on an arbitrary channel, we return to the bifurcation: Since mass is conserved, the flow rate of the parent branch Qdisplaystyle QQ is the sum of the flow rates in the children branches Q1,…,Qndisplaystyle Q_1,ldots ,Q_ndisplaystyle Q_1,ldots ,Q_n. Since in each of these branches the power is minimized, the relationship above holds and hence, as claimed:

Q=∑i=1NQi⇒π4λμr3=∑i=1Nπ4λμri3⇒r3=∑i=1Nri3.displaystyle Q=sum _i=1^NQ_i;Rightarrow ;frac pi 4sqrt frac lambda mu ,r^3=sum _i=1^Nfrac pi 4sqrt frac lambda mu ,r_i^3;Rightarrow ;r^3=sum _i=1^Nr_i^3.

displaystyle Q=sum _i=1^NQ_i;Rightarrow ;frac pi 4sqrt frac lambda mu ,r^3=sum _i=1^Nfrac pi 4sqrt frac lambda mu ,r_i^3;Rightarrow ;r^3=sum _i=1^Nr_i^3.



Diffusion


The power for transport spent by diffusion is given by


Pt=QΔC=(πDr2ΔCl)ΔCdisplaystyle P_t=Q,Delta C=left(pi ,D,r^2frac Delta Clright)Delta Cdisplaystyle P_t=Q,Delta C=left(pi ,D,r^2frac Delta Clright)Delta C


where the flow rate is given by Fick's Law, whose Ddisplaystyle DD is the diffusivity constant and ΔCdisplaystyle Delta CDelta C is the difference of concentration between the ends of the cylinder. Similarly to the case of laminar flow, the minimisation of the objective function results in


ΔCl=λD⇒Q=πDr2λD⇔Q=kr2displaystyle frac Delta Cl=sqrt frac lambda DRightarrow Q=pi ,D,r^2sqrt frac lambda DLeftrightarrow Q=k,r^2displaystyle frac Delta Cl=sqrt frac lambda DRightarrow Q=pi ,D,r^2sqrt frac lambda DLeftrightarrow Q=k,r^2


Hence,


Qp=∑i=1NQi⇒krp2=∑i=1Nkri2⇔rp2=∑i=1Nri2displaystyle Q_p=sum _i=1^NQ_iRightarrow k,r_p^2=sum _i=1^Nk,r_i^2Leftrightarrow r_p^2=sum _i=1^Nr_i^2displaystyle Q_p=sum _i=1^NQ_iRightarrow k,r_p^2=sum _i=1^Nk,r_i^2Leftrightarrow r_p^2=sum _i=1^Nr_i^2



The generalized Murray’s law



However, the special Murray’s law is only applicable to flow processes involving no mass variations. Significant theoretical advances need to be made for more broadly applicable in the fields of chemistry, applied materials, and industrial reactions.




Murray networks. (Zheng, CC-BY-SA)


The generalized Murray's law deduced by Zheng et al. can be applicable for optimizing mass transfer involving mass variations and chemical reactions involving flow processes, molecule or ion diffusion, etc.[1]


For connecting a parent pipe with radius of r0 to many children pipes with radius of ri , the formula of generalized Murray's law is: roa=11−X∑i=1Nriadisplaystyle r_o^a=1 over 1-Xsum _i=1^Nr_i^adisplaystyle r_o^a=1 over 1-Xsum _i=1^Nr_i^a, where the X is the ratio of mass variation during mass transfer in the parent pore, the exponent α is dependent on the type of the transfer. For laminar flow α =3; for turbulent flow α =7/3; for molecule or ionic diffusion α =2; etc.


I t is applicable to an enormous range of porous materials and has a broad scope in functional ceramics and nano-metals for energy and environmental applications.



Murray materials


The generalized Murray’s law defines the basic geometric features for porous materials with optimum transfer properties. The generalized Murray’s law can be used to design and optimize the structures of an enormous range of porous materials. This concept has led to materials, termed as the Murray materials, whose pore-sizes are multiscale and are designed with diameter-ratios obeying the generalized Murray’s law.[1]




A diagram of Murray materials with macro-meso-micropores built by nanopaticles as building blocks. (Zheng, CC-BY-SA)


As lithium battery electrodes, the Murray materials can reduce the stresses in these electrodes during the charge/discharge processes, improving their structural stability and resulting in a longer life time for energy storage devices. This material could also be used for boosting the performance of a gas sensor and a photocatalysis process that broke down a dye using light.




Murray materials in leaf and insect. (Zheng, CC-BY-SA)


To achieve substances or energy transfer with extremely high efficiency, evolution by natural selection has endowed many classes of organisms with Murray materials, in which the pore-sizes regularly decrease across multiple scales and finally terminate in size-invariant units. For example, in plant stems and leaf veins, the sum of the radii cubed remains constant across every branch point to maximize the flow conductance, which is proportional to the rate of photosynthesis. For insects relying upon gas diffusion for breathing, the sum of radii squared of tracheal pores remains constant along the diffusion pathway, to maximize gases diffusion. From plants, animals and materials to industrial processes, the introduction of Murray material concept to industrial reactions can revolutionize the design of reactors with highly enhanced efficiency, minimum energy, time, and raw material consumption for a sustainable future.



References




  1. ^ abcd Zheng, Xianfeng; Shen, Guofang; Wang, Chao; Li, Yu; Dunphy, Darren; Hasan, Tawfique; Brinker, C. Jeffrey; Su, Bao-Lian (2017-04-06). "Bio-inspired Murray materials for mass transfer and activity". Nature Communications. 8. doi:10.1038/ncomms14921. ISSN 2041-1723..mw-parser-output cite.citationfont-style:inherit.mw-parser-output qquotes:"""""""'""'".mw-parser-output code.cs1-codecolor:inherit;background:inherit;border:inherit;padding:inherit.mw-parser-output .cs1-lock-free abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-lock-limited a,.mw-parser-output .cs1-lock-registration abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-lock-subscription abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registrationcolor:#555.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration spanborder-bottom:1px dotted;cursor:help.mw-parser-output .cs1-hidden-errordisplay:none;font-size:100%.mw-parser-output .cs1-visible-errorfont-size:100%.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-formatfont-size:95%.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-leftpadding-left:0.2em.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-rightpadding-right:0.2em


  2. ^ ab Sherman, Thomas F. (1981). "On connecting large vessels to small. The meaning of Murray's law" (pdf). The Journal of General Physiology. 78 (4): a 431–453. doi:10.1085/jgp.78.4.431. PMC 2228620. PMID 7288393.


  3. ^ abc Williams, Hugo R.; Trask, Richard S.; Weaver, Paul M.; Bond, Ian P. (2008). "Minimum mass vascular networks in multifunctional materials". Journal of the Royal Society Interface. 5 (18): 55–65. doi:10.1098/rsif.2007.1022. PMC 2605499. PMID 17426011.


  4. ^ ab McCulloh, Katherine A.; John S. Sperry; Frederick R. Adler (2003). "Water transport in plants obeys Murray's law". Nature. 421 (6926): 939–942. doi:10.1038/nature01444. PMID 12607000.


  5. ^ Murray, Cecil D. (1926). "The Physiological Principle of Minimum Work: I. The Vascular System and the Cost of Blood Volume". Proceedings of the National Academy of Sciences of the United States of America. 12 (3): 207–214. doi:10.1073/pnas.12.3.207. PMC 1084489. PMID 16576980.


  6. ^ Murray, Cecil D. (1926). "The Physiological Principle of Minimum Work: II. Oxygen Exchange in Capillaries". Proceedings of the National Academy of Sciences of the United States of America. 12 (5): 299–304. doi:10.1073/pnas.12.5.299. PMC 1084544. PMID 16587082.








Popular posts from this blog

倭马亚王朝

Gabbro

托萊多 (西班牙)