Skip to main content

Hagen–Poiseuille equation








Hagen–Poiseuille equation


From Wikipedia, the free encyclopedia
  (Redirected from Poiseuille's law)

Jump to navigation
Jump to search







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[1] and Gotthilf Heinrich Ludwig Hagen,[2] and published by Poiseuille in 1840–41 and 1846.[1]


The assumptions of the equation are that the fluid is incompressible and Newtonian; the flow is laminar through a pipe of constant circular cross-section that is substantially longer than its diameter; and there is no acceleration of fluid in the pipe. For velocities and pipe diameters above a threshold, actual fluid flow is not laminar but turbulent, leading to larger pressure drops than calculated by the Hagen–Poiseuille equation.




Contents





  • 1 Equation


  • 2 Relation to Darcy–Weisbach


  • 3 Derivation

    • 3.1 Liquid flow through a pipe


    • 3.2 Viscosity


    • 3.3 Faster lamina


    • 3.4 Slower lamina


    • 3.5 Putting it all together


    • 3.6 Poiseuille's law


    • 3.7 Startup of Poiseuille flow in a pipe[8]



  • 4 Poiseuille flow in annular section[9]


  • 5 Plane Poiseuille flow


  • 6 Poiseuille flow through some non-circular cross-sections[10]


  • 7 Poiseuille flow through arbitrary cross-section


  • 8 Poiseuille's equation for compressible fluids


  • 9 Electrical circuits analogy


  • 10 Medical applications – intravenous access and fluid delivery


  • 11 See also


  • 12 Notes


  • 13 References


  • 14 External links




Equation[edit]


In standard fluid-kinetics notation:[3][4][5]


ΔP=8μLQπR4displaystyle Delta P=frac 8mu LQpi R^4displaystyle Delta P=frac 8mu LQpi R^4

where:



ΔP is the pressure difference between the two ends,


L is the length of pipe,


μ is the dynamic viscosity,


Q is the volumetric flow rate,


R is the pipe radius.

The equation does not hold close to the pipe entrance.[6]:3


The equation fails in the limit of low viscosity, wide and/or short pipe. Low viscosity or a wide pipe may result in turbulent flow, making it necessary to use more complex models, such as Darcy–Weisbach equation. If the pipe is too short, the Hagen–Poiseuille equation may result in unphysically high flow rates; the flow is bounded by Bernoulli's principle, under less restrictive conditions, by


Qmax=πR22ΔPρ.displaystyle Q_max =pi R^2sqrt frac 2Delta Prho .displaystyle Q_max =pi R^2sqrt frac 2Delta Prho .


Relation to Darcy–Weisbach[edit]


Normally, Hagen-Poiseuille flow implies not just the relation for the pressure drop, above, but also the full solution for the laminar flow profile, which is parabolic. However, the result for the pressure drop can be extended to turbulent flow by inferring an effective turbulent viscosity in the case of turbulent flow, even though the flow profile in turbulent flow is strictly speaking not actually parabolic. In both cases, laminar or turbulent, the pressure drop is related to the stress at the wall, which determines the so-called friction factor. The wall stress can be determined phenomenologically by the Darcy–Weisbach equation in the field of hydraulics, given a relationship for the friction factor in terms of the Reynolds number. In the case of laminar flow, for a circular cross section:


Λ=64Re,Re=ρvdμ,displaystyle Lambda =frac 64mathrm Re ,quad mathrm Re =frac rho vdmu ,displaystyle Lambda =frac 64mathrm Re ,quad mathrm Re =frac rho vdmu ,

where Re is the Reynolds number, ρ is the fluid density, v is the mean flow velocity, which is half the maximal flow velocity in the case of laminar flow. It proves more useful to define the Reynolds number in terms of the mean flow velocity because this quantity remains well defined even in the case of turbulent flow, whereas the maximal flow velocity may not be, or in any case, it may be difficult to infer. In this form the law approximates the Darcy friction factor, the energy (head) loss factor, friction loss factor or Darcy (friction) factor Λ in the laminar flow at very low velocities in cylindrical tube. The theoretical derivation of a slightly different form of the law was made independently by Wiedman in 1856 and Neumann and E. Hagenbach in 1858 (1859, 1860). Hagenbach was the first who called this law the Poiseuille's law.


The law is also very important in hemorheology and hemodynamics, both fields of physiology.[7]


Poiseuille's law was later in 1891 extended to turbulent flow by L. R. Wilberforce, based on Hagenbach's work.



Derivation[edit]



The Hagen–Poiseuille equation can be derived from the Navier–Stokes equations. Although more lengthy than directly using the Navier–Stokes equations, an alternative method of deriving the Hagen–Poiseuille equation is as follows.



Liquid flow through a pipe[edit]





a) A tube showing the imaginary lamina. b) A cross section of the tube shows the lamina moving at different speeds. Those closest to the edge of the tube are moving slowly while those near the center are moving quickly.


Assume the liquid exhibits laminar flow. Laminar flow in a round pipe prescribes that there are a bunch of circular layers (lamina) of liquid, each having a velocity determined only by their radial distance from the center of the tube. Also assume the center is moving fastest while the liquid touching the walls of the tube is stationary (due to the no-slip condition).


To figure out the motion of the liquid, all forces acting on each lamina must be known:


  1. The pressure force pushing the liquid through the tube is the change in pressure multiplied by the area: F = −A ΔP. This force is in the direction of the motion of the liquid. The negative sign comes from the conventional way we define ΔP = PendPtop < 0.


  2. Viscosity effects will pull from the faster lamina immediately closer to the center of the tube.


  3. Viscosity effects will drag from the slower lamina immediately closer to the walls of the tube.


Viscosity[edit]




Two fluids moving past each other in the x direction. The liquid on top is moving faster and will be pulled in the negative direction by the bottom liquid while the bottom liquid will be pulled in the positive direction by the top liquid.


When two layers of liquid in contact with each other move at different speeds, there will be a shear force between them. This force is proportional to the area of contact A, the velocity gradient perpendicular to the direction of flow Δvx/Δy, and a proportionality constant (viscosity) and is given by


Fviscosity, top=−μAΔvxΔy.displaystyle F_textviscosity, top=-mu Afrac Delta v_xDelta y.displaystyle F_textviscosity, top=-mu Afrac Delta v_xDelta y.

The negative sign is in there because we are concerned with the faster moving liquid (top in figure), which is being slowed by the slower liquid (bottom in figure). By Newton's third law of motion, the force on the slower liquid is equal and opposite (no negative sign) to the force on the faster liquid. This equation assumes that the area of contact is so large that we can ignore any effects from the edges and that the fluids behave as Newtonian fluids.



Faster lamina[edit]


Assume that we are figuring out the force on the lamina with radius r. From the equation above, we need to know the area of contact and the velocity gradient. Think of the lamina as a ring of radius r, thickness dr, and length Δx. The area of contact between the lamina and the faster one is simply the area of the inside of the cylinder: A = 2πr Δx. We don't know the exact form for the velocity of the liquid within the tube yet, but we do know (from our assumption above) that it is dependent on the radius. Therefore, the velocity gradient is the change of the velocity with respect to the change in the radius at the intersection of these two laminae. That intersection is at a radius of r. So, considering that this force will be positive with respect to the movement of the liquid (but the derivative of the velocity is negative), the final form of the equation becomes


Fviscosity, fast=−2πrμΔxdvdr|rdisplaystyle F_textviscosity, fast=-2pi rmu ,Delta x,left.frac dvdrrightdisplaystyle F_textviscosity, fast=-2pi rmu ,Delta x,left.frac dvdrright

where the vertical bar and subscript r following the derivative indicates that it should be taken at a radius of r.



Slower lamina[edit]


Next let's find the force of drag from the slower lamina. We need to calculate the same values that we did for the force from the faster lamina. In this case, the area of contact is at r + dr instead of r. Also, we need to remember that this force opposes the direction of movement of the liquid and will therefore be negative (and that the derivative of the velocity is negative).


Fviscosity, slow=2π(r+dr)μΔxdvdr|r+drdisplaystyle F_textviscosity, slow=2pi (r+dr)mu ,Delta xleft.frac dvdrrightdisplaystyle F_textviscosity, slow=2pi (r+dr)mu ,Delta xleft.frac dvdrright


Putting it all together[edit]


To find the solution for the flow of a laminar layer through a tube, we need to make one last assumption. There is no acceleration of liquid in the pipe, and by Newton's first law, there is no net force. If there is no net force then we can add all of the forces together to get zero


0=Fpressure+Fviscosity, fast+Fviscosity, slowdisplaystyle 0=F_textpressure+F_textviscosity, fast+F_textviscosity, slowdisplaystyle 0=F_textpressure+F_textviscosity, fast+F_textviscosity, slow

or


0=ΔP2πrdr−2πrμΔxdvdr|r+2π(r+dr)μΔxdvdr|r+dr._r+2pi (r+dr)mu ,Delta x,left.frac dvdrrightvert _r+dr._r+2pi (r+dr)mu ,Delta x,left.frac dvdrrightvert _r+dr.

First, to get everything happening at the same point, use the first two terms of a Taylor series expansion of the velocity gradient:


dvdr|r+dr=dvdr|r+d2vdr2|rdr._r+dr=left.frac dvdrright_r+dr=left.frac dvdrright

The expression is valid for all laminae. Grouping like terms and dropping the vertical bar since all derivatives are assumed to be at radius r,


0=ΔP2πrdr+2πμdrΔxdvdr+2πrμdrΔxd2vdr2+2πμ(dr)2Δxd2vdr2.displaystyle 0=Delta P2pi r,dr+2pi mu ,dr,Delta xfrac dvdr+2pi rmu ,dr,Delta xfrac d^2vdr^2+2pi mu (dr)^2,Delta xfrac d^2vdr^2.displaystyle 0=Delta P2pi r,dr+2pi mu ,dr,Delta xfrac dvdr+2pi rmu ,dr,Delta xfrac d^2vdr^2+2pi mu (dr)^2,Delta xfrac d^2vdr^2.

Finally, put this expression in the form of a differential equation, dropping the term quadratic in dr.


1μΔPΔx=d2vdr2+1rdvdrdisplaystyle frac 1mu frac Delta PDelta x=frac d^2vdr^2+frac 1rfrac dvdrdisplaystyle frac 1mu frac Delta PDelta x=frac d^2vdr^2+frac 1rfrac dvdr

It can be seen that both sides of the equations are negative: there is a drop of pressure along the tube (left side) and both first and second derivatives of the velocity are negative (velocity has a maximum value at the center of the tube, where r = 0). Using the product rule, the equation may be rearranged to:


1μΔPΔx=1rddr(rdvdr).displaystyle frac 1mu frac Delta PDelta x=frac 1rfrac ddrleft(rfrac dvdrright).displaystyle frac 1mu frac Delta PDelta x=frac 1rfrac ddrleft(rfrac dvdrright).

The right-hand side is the radial term of the Laplace operator 2, so this differential equation is a special case of the Poisson equation. It is subject to the following boundary conditions:






v(r)=0 at r=Rdisplaystyle v(r)=0quad mbox at r=Rdisplaystyle v(r)=0quad mbox at r=R
 — "no-slip" boundary condition at the wall
dvdr=0 at r=0displaystyle frac dvdr=0quad mbox at r=0displaystyle frac dvdr=0quad mbox at r=0
 — axial symmetry.

Axial symmetry means that the velocity v(r) is maximum at the center of the tube, therefore the first derivative dv/dr is zero at r = 0.


The differential equation can be integrated to:


v(r)=14μr2ΔPΔx+Aln⁡(r)+B.displaystyle v(r)=frac 14mu r^2frac Delta PDelta x+Aln(r)+B.displaystyle v(r)=frac 14mu r^2frac Delta PDelta x+Aln(r)+B.

To find A and B, we use the boundary conditions.


First, the symmetry boundary condition indicates:


dvdr=12μrΔPΔx+Ar=0 at r=0.displaystyle frac dvdr=frac 12mu rfrac Delta PDelta x+frac Ar=0quad mbox at r=0.displaystyle frac dvdr=frac 12mu rfrac Delta PDelta x+frac Ar=0quad mbox at r=0.

A solution possible only if A = 0. Next the no-slip boundary condition is applied to the remaining equation:


v(R)=14μR2ΔPΔx+B=0displaystyle v(R)=frac 14mu R^2frac Delta PDelta x+B=0displaystyle v(R)=frac 14mu R^2frac Delta PDelta x+B=0

so therefore


B=−14μR2ΔPΔx.displaystyle B=-frac 14mu R^2frac Delta PDelta x.displaystyle B=-frac 14mu R^2frac Delta PDelta x.

Now we have a formula for the velocity of liquid moving through the tube as a function of the distance from the center of the tube


v=−14μΔPΔx(R2−r2)displaystyle v=-frac 14mu frac Delta PDelta x(R^2-r^2)displaystyle v=-frac 14mu frac Delta PDelta x(R^2-r^2)

or, at the center of the tube where the liquid is moving fastest (r = 0) with R being the radius of the tube,


vmax=−14μΔPΔxR2.displaystyle v_max=-frac 14mu frac Delta PDelta xR^2.displaystyle v_max=-frac 14mu frac Delta PDelta xR^2.


Poiseuille's law[edit]


To get the total volume that flows through the tube, we need to add up the contributions from each lamina. To calculate the flow through each lamina, we multiply the velocity (from above) and the area of the lamina.


Q(r)dr=−14μΔPΔx(R2−r2)2πrdr=−π2μΔPΔx(rR2−r3)drdisplaystyle Q(r),dr=-frac 14mu frac Delta PDelta x(R^2-r^2)2pi r,dr=-frac pi 2mu frac Delta PDelta x(rR^2-r^3),drdisplaystyle Q(r),dr=-frac 14mu frac Delta PDelta x(R^2-r^2)2pi r,dr=-frac pi 2mu frac Delta PDelta x(rR^2-r^3),dr

Finally, we integrate over all lamina via the radius variable r.


Q=−π2μΔPΔx∫0R(rR2−r3)dr=−ΔPπR48μΔxdisplaystyle Q=-frac pi 2mu frac Delta PDelta xint _0^R(rR^2-r^3),dr=-frac Delta Ppi R^48mu Delta xdisplaystyle Q=-frac pi 2mu frac Delta PDelta xint _0^R(rR^2-r^3),dr=-frac Delta Ppi R^48mu Delta x


Startup of Poiseuille flow in a pipe[8][edit]


When a constant pressure gradient G=−dp/dx=constantdisplaystyle G=-dp/dx=textconstantdisplaystyle G=-dp/dx=textconstant is applied between two ends of a long pipe, the flow will not immediately obtain Poiseuille profile, rather it develops through time and reaches the Poiseuille profile at steady state. The Navier-Stokes equations reduce to


∂u∂t=Gρ+ν(∂2u∂r2+1r∂u∂r)displaystyle frac partial upartial t=frac Grho +nu left(frac partial ^2upartial r^2+frac 1rfrac partial upartial rright)displaystyle frac partial upartial t=frac Grho +nu left(frac partial ^2upartial r^2+frac 1rfrac partial upartial rright)

with initial and boundary conditions,


u(r,0)=0,u(R,t)=0.displaystyle u(r,0)=0,quad u(R,t)=0.displaystyle u(r,0)=0,quad u(R,t)=0.

The velocity distribution is given by


u(r,t)=G4μ(R2−r2)−2GR2μ∑n=1∞1λn3Jo(λnr/R)J1(λn)e−λn2νtR2,Jo(λn)=0displaystyle u(r,t)=frac G4mu (R^2-r^2)-frac 2GR^2mu sum _n=1^infty frac 1lambda _n^3frac J_o(lambda _nr/R)J_1(lambda _n)e^-lambda _n^2frac nu tR^2,quad J_o(lambda _n)=0displaystyle u(r,t)=frac G4mu (R^2-r^2)-frac 2GR^2mu sum _n=1^infty frac 1lambda _n^3frac J_o(lambda _nr/R)J_1(lambda _n)e^-lambda _n^2frac nu tR^2,quad J_o(lambda _n)=0

where Jo(λnr/R)displaystyle J_o(lambda _nr/R)displaystyle J_o(lambda _nr/R) is the Bessel function of the first kind of order zero and λndisplaystyle lambda _nlambda _n are the positive roots of this function and J1(λn)displaystyle J_1(lambda _n)displaystyle J_1(lambda _n) is the Bessel function of the first kind of order one. As t→∞displaystyle trightarrow infty trightarrow infty , Poiseuille solution is recovered.



Poiseuille flow in annular section[9][edit]




Poiseuille flow in annular section


If R1displaystyle R_1R_1 is the inner cylinder radii and R2displaystyle R_2R_2 is the outer cylinder radii, with applied pressure gradient between the two ends G=−dp/dx=constantdisplaystyle G=-dp/dx=textconstantdisplaystyle G=-dp/dx=textconstant, the velocity distribution and the volume flux through the annular pipe are


u(r)=G4μ(R12−r2)+G4μ(R22−R12)ln⁡(r/R1)ln⁡(R2/R1),Q=Gπ8μ[R24−R14−(R22−R12)2ln⁡R2/R1].displaystyle beginalignedu(r)&=frac G4mu (R_1^2-r^2)+frac G4mu (R_2^2-R_1^2)frac ln(r/R_1)ln(R_2/R_1),\Q&=frac Gpi 8mu left[R_2^4-R_1^4-frac (R_2^2-R_1^2)^2ln R_2/R_1right].endaligneddisplaystyle beginalignedu(r)&=frac G4mu (R_1^2-r^2)+frac G4mu (R_2^2-R_1^2)frac ln(r/R_1)ln(R_2/R_1),\Q&=frac Gpi 8mu left[R_2^4-R_1^4-frac (R_2^2-R_1^2)^2ln R_2/R_1right].endaligned

When R2=R, R1=0displaystyle R_2=R, R_1=0displaystyle R_2=R, R_1=0, the original problem is recovered.



Plane Poiseuille flow[edit]




Plane Poiseuille flow


Plane Poiseuille flow is flow created between two infinitely long parallel plates, separated by a distance hdisplaystyle hh with a constant pressure gradient G=−dp/dx=constantdisplaystyle G=-dp/dx=textconstantdisplaystyle G=-dp/dx=textconstant is applied in the direction of flow. The flow is essentially unidirectional because of infinite length. The Navier-Stokes equations reduce to


d2udy2=−Gμdisplaystyle frac d^2udy^2=-frac Gmu displaystyle frac d^2udy^2=-frac Gmu

with no-slip condition on both walls


u(0)=0,u(h)=0displaystyle u(0)=0,quad u(h)=0displaystyle u(0)=0,quad u(h)=0

Therefore, the velocity distribution and the volume flow rate per unit length are


u(y)=G2μy(h−y),Q=Gh312μ.displaystyle u(y)=frac G2mu y(h-y),quad Q=frac Gh^312mu .displaystyle u(y)=frac G2mu y(h-y),quad Q=frac Gh^312mu .


Poiseuille flow through some non-circular cross-sections[10][edit]


Joseph Boussinesq[11] derived the velocity profile and volume flow rate in 1868 for rectangular channel and tubes of equilateral triangular cross-section and for elliptical cross-section. Joseph Proudman[12] derived the same for isosceles triangles in 1914. Let G=−dp/dx=constantdisplaystyle G=-dp/dx=textconstantdisplaystyle G=-dp/dx=textconstant be the constant pressure gradient acting in direction parallel to the motion.


The velocity and the volume flow rate in a rectangular channel of height 0≤y≤hdisplaystyle 0leq yleq hdisplaystyle 0leq yleq h and width 0≤z≤ldisplaystyle 0leq zleq ldisplaystyle 0leq zleq l are


u(y,z)=G2μy(h−y)−4Gh2μπ3∑n=1∞1(2n−1)3sinh⁡(βnz)+sinh⁡(βn(l−z))sinh⁡(βnl)sin⁡(βny),βn=(2n−1)πh,Q=Gh3l12μ−16Gh4π5μ∑n=1∞1(2n−1)5cosh⁡(βnl)−1sinh⁡(βnl).displaystyle beginalignedu(y,z)&=frac G2mu y(h-y)-frac 4Gh^2mu pi ^3sum _n=1^infty frac 1(2n-1)^3frac sinh(beta _nz)+sinh(beta _n(l-z))sinh(beta _nl)sin(beta _ny),quad beta _n=frac (2n-1)pi h,\Q&=frac Gh^3l12mu -frac 16Gh^4pi ^5mu sum _n=1^infty frac 1(2n-1)^5frac cosh(beta _nl)-1sinh(beta _nl).endaligneddisplaystyle beginalignedu(y,z)&=frac G2mu y(h-y)-frac 4Gh^2mu pi ^3sum _n=1^infty frac 1(2n-1)^3frac sinh(beta _nz)+sinh(beta _n(l-z))sinh(beta _nl)sin(beta _ny),quad beta _n=frac (2n-1)pi h,\Q&=frac Gh^3l12mu -frac 16Gh^4pi ^5mu sum _n=1^infty frac 1(2n-1)^5frac cosh(beta _nl)-1sinh(beta _nl).endaligned

The velocity and the volume flow rate of tube with equilateral triangular cross-section of side length 2h/3displaystyle 2h/sqrt 3displaystyle 2h/sqrt 3 are


u(y,z)=−G4μh(y−h)(y2−3z2),Q=Gh4603μ.displaystyle beginalignedu(y,z)&=-frac G4mu h(y-h)(y^2-3z^2),\Q&=frac Gh^460sqrt 3mu .endaligneddisplaystyle beginalignedu(y,z)&=-frac G4mu h(y-h)(y^2-3z^2),\Q&=frac Gh^460sqrt 3mu .endaligned

The velocity and the volume flow rate in the right-angled isosceles triangle y=π, y±z=0displaystyle y=pi , ypm z=0displaystyle y=pi , ypm z=0 are


u(y,z)=G2μ(y+z)(π−y)−Gπμ∑n=1∞1βn3sinh⁡(2πβn)sinh⁡[βn(2π−y+z)]sin⁡[βn(y+z)]−sinh⁡[βn(y+z)]sin⁡[βn(y−z)],βn=n−12,Q=Gπ412μ−G2πμ∑n=1∞1βn5[coth⁡(2πβn)+csc⁡(2πβn)].displaystyle beginalignedu(y,z)&=frac G2mu (y+z)(pi -y)-frac Gpi mu sum _n=1^infty frac 1beta _n^3sinh(2pi beta _n)sinh[beta _n(2pi -y+z)]sin[beta _n(y+z)]-sinh[beta _n(y+z)]sin[beta _n(y-z)],quad beta _n=n-frac 12,\Q&=frac Gpi ^412mu -frac G2pi mu sum _n=1^infty frac 1beta _n^5[coth(2pi beta _n)+csc(2pi beta _n)].endaligneddisplaystyle beginalignedu(y,z)&=frac G2mu (y+z)(pi -y)-frac Gpi mu sum _n=1^infty frac 1beta _n^3sinh(2pi beta _n)sinh[beta _n(2pi -y+z)]sin[beta _n(y+z)]-sinh[beta _n(y+z)]sin[beta _n(y-z)],quad beta _n=n-frac 12,\Q&=frac Gpi ^412mu -frac G2pi mu sum _n=1^infty frac 1beta _n^5[coth(2pi beta _n)+csc(2pi beta _n)].endaligned

The velocity distribution for tubes of elliptical cross-section with semi-axis adisplaystyle aa and bdisplaystyle bb is[8]


u(y,z)=G2μ(1a2+1b2)(1−y2a2−z2b2),Q=πGa3b34μ(a2+b2).displaystyle beginalignedu(y,z)&=frac G2mu left(frac 1a^2+frac 1b^2right)left(1-frac y^2a^2-frac z^2b^2right),\Q&=frac pi Ga^3b^34mu (a^2+b^2).endaligneddisplaystyle beginalignedu(y,z)&=frac G2mu left(frac 1a^2+frac 1b^2right)left(1-frac y^2a^2-frac z^2b^2right),\Q&=frac pi Ga^3b^34mu (a^2+b^2).endaligned

Here, when a=bdisplaystyle a=ba=b, Poiseuille flow for circular pipe is recovered and when a→∞displaystyle arightarrow infty arightarrowinfty, plane Poiseuille flow is recovered.



Poiseuille flow through arbitrary cross-section[edit]


The flow through arbitrary cross-section u(y,z)displaystyle u(y,z)displaystyle u(y,z) satisfies the condition that u=0displaystyle u=0u=0 on the walls. The governing equation reduces to[13]


∂2u∂y2+∂2u∂z2=−Gμ.displaystyle frac partial ^2upartial y^2+frac partial ^2upartial z^2=-frac Gmu .displaystyle frac partial ^2upartial y^2+frac partial ^2upartial z^2=-frac Gmu .

If we introduce a new dependent variable as


U=u+G4μ(y2+z2),displaystyle U=u+frac G4mu (y^2+z^2),displaystyle U=u+frac G4mu (y^2+z^2),

then it is easy to see that the problem reduces to that integrating a Laplace equation


∂2U∂y2+∂2U∂z2=0displaystyle frac partial ^2Upartial y^2+frac partial ^2Upartial z^2=0displaystyle frac partial ^2Upartial y^2+frac partial ^2Upartial z^2=0

satisfying the condition


U=G4μ(y2+z2)displaystyle U=frac G4mu (y^2+z^2)displaystyle U=frac G4mu (y^2+z^2)

on the wall.



Poiseuille's equation for compressible fluids[edit]


For a compressible fluid in a tube the volumetric flow rate and the linear velocity are not constant along the tube. The flow is usually expressed at outlet pressure. As fluid is compressed or expands, work is done and the fluid is heated or cooled. This means that the flow rate depends on the heat transfer to and from the fluid. For an ideal gas in the isothermal case, where the temperature of the fluid is permitted to equilibrate with its surroundings, and when the pressure difference between ends of the pipe is small, the volumetric flow rate at the pipe outlet is given by


Q=dVdt=vπR2=πR4(Pi−Po)8μL×Pi+Po2Po=πR416μL(Pi2−Po2Po)displaystyle Q=frac dVdt=vpi R^2=frac pi R^4left(P_mathrm i -P_mathrm o right)8mu Ltimes frac P_mathrm i +P_mathrm o 2P_mathrm o =frac pi R^416mu Lleft(frac P_mathrm i ^2-P_mathrm o ^2P_mathrm o right)displaystyle Q=frac dVdt=vpi R^2=frac pi R^4left(P_mathrm i -P_mathrm o right)8mu Ltimes frac P_mathrm i +P_mathrm o 2P_mathrm o =frac pi R^416mu Lleft(frac P_mathrm i ^2-P_mathrm o ^2P_mathrm o right)

where:



Pi is inlet pressure


Po is outlet pressure


L is the length of tube


μdisplaystyle mu mu is the viscosity


R is the radius


V is the volume of the fluid at outlet pressure


v is the velocity of the fluid at outlet pressure

This equation can be seen as Poiseuille's law with an extra correction factor Pi + Po/2Po expressing the average pressure relative to the outlet pressure.



Electrical circuits analogy[edit]




Electricity was originally understood to be a kind of fluid. This hydraulic analogy is still conceptually useful for understanding circuits. This analogy is also used to study the frequency response of fluid-mechanical networks using circuit tools, in which case the fluid network is termed a hydraulic circuit.


Poiseuille's law corresponds to Ohm's law for electrical circuits, V = IR.


Since the net force acting on the fluid is equal to


ΔF=SΔP,displaystyle Delta F=SDelta P,displaystyle Delta F=SDelta P,

where S = πr2, i.e. ΔF = πr2 ΔP, then from Poiseuille's law


ΔP=8μLQπr4displaystyle Delta P=frac 8mu LQpi r^4Delta P=frac 8mu LQpi r^4

it follows that



ΔF=8μLQr2displaystyle Delta F=frac 8mu LQr^2Delta F=frac 8mu LQr^2.

For electrical circuits, let n be the concentration of free charged particles (in m−3) and let q* be the charge of each particle (in coulombs). (For electrons, q* = e = 6981160000000000000♠1.6×10−19 C.)


Then nQ is the number of particles in the volume Q, and nQq* is their total charge. This is the charge that flows through the cross section per unit time, i.e. the current I. Therefore, I = nQq*. Consequently, Q = I/nq*, and


ΔF=8μLInr2q∗.displaystyle Delta F=frac 8mu LInr^2q^*.displaystyle Delta F=frac 8mu LInr^2q^*.

But ΔF = Eq, where q is the total charge in the volume of the tube. The volume of the tube is equal to πr2L, so the number of charged particles in this volume is equal to nπr2L, and their total charge is


q=nπr2Lq∗.displaystyle q=npi r^2Lq^*.displaystyle q=npi r^2Lq^*.

Now,


E=ΔFq=8μIn2πr4(q∗)2.displaystyle E=frac Delta Fq=frac 8mu In^2pi r^4(q^*)^2.displaystyle E=frac Delta Fq=frac 8mu In^2pi r^4(q^*)^2.

Since the voltage V = EL, we get


V=8μLIn2πr4(q∗)2.displaystyle V=frac 8mu LIn^2pi r^4(q^*)^2.displaystyle V=frac 8mu LIn^2pi r^4(q^*)^2.

This is exactly Ohm's law, where the resistance R = V/I is described by the formula



R=8μLn2πr4(q∗)2displaystyle R=frac 8mu Ln^2pi r^4(q^*)^2displaystyle R=frac 8mu Ln^2pi r^4(q^*)^2.

It follows that the resistance R is proportional to the length L of the resistor, which is true. However, it also follows that
the resistance R is inversely proportional to the fourth power of the radius r, i.e. the resistance R is inversely proportional to the second power of the cross section area S = πr2 of the resistor, which is wrong according to the electrical analogy.


The correct relation is


R=ρLS,displaystyle R=frac rho LS,displaystyle R=frac rho LS,

where ρ is the specific resistance; i.e. the resistance R is inversely proportional to the cross section area S of the resistor.[14]


The reason why Poiseuille's law leads to a wrong formula for the resistance R is the difference between the fluid flow and the electric current. Electron gas is inviscid, so its velocity does not depend on the distance to the walls of the conductor. The resistance is due to the interaction between the flowing electrons and the atoms of the conductor. Therefore, Poiseuille's law and the hydraulic analogy are useful only within certain limits when applied to electricity.


Both Ohm's law and Poiseuille's law illustrate transport phenomena.



Medical applications – intravenous access and fluid delivery[edit]




The Hagen–Poiseuille equation is useful in determining the flow rate of intravenous fluids that may be achieved using various sizes of peripheral and central cannulas. The equation states that flow rate is proportional to the radius to the fourth power, meaning that a small increase in the internal diameter of the cannula yields a significant increase in flow rate of IV fluids. The radius of IV cannulas is typically measured in "gauge", which is inversely proportional to the radius. Peripheral IV cannulas are typically available as (from large to small) 14G, 16G, 18G, 20G, 22G. As an example, the flow of a 14G cannula is typically twice that of a 16G, and ten times that of a 20G. It also states that flow is inversely proportional to length, meaning that longer lines have lower flow rates. This is important to remember as in an emergency, many clinicians favor shorter, larger catheters compared to longer, narrower catheters. While of less clinical importance, the change in pressure can be used to speed up flow rate by pressurizing the bag of fluid, squeezing the bag, or hanging the bag higher from the level of the cannula. It is also useful to understand that viscous fluids will flow slower (e.g. in blood transfusion).



See also[edit]


  • Couette flow

  • Darcy's law

  • Pulse

  • Wave

  • Hydraulic circuit


Notes[edit]




  1. ^ ab Sutera, Salvatore P.; Skalak, Richard (1993). "The History of Poiseuille's Law". Annual Review of Fluid Mechanics. 25: 1–19. Bibcode:1993AnRFM..25....1S. doi:10.1146/annurev.fl.25.010193.000245..mw-parser-output cite.citationfont-style:inherit.mw-parser-output .citation qquotes:"""""""'""'".mw-parser-output .citation .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 .citation .cs1-lock-limited a,.mw-parser-output .citation .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 .citation .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-ws-icon abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center.mw-parser-output code.cs1-codecolor:inherit;background:inherit;border:inherit;padding:inherit.mw-parser-output .cs1-hidden-errordisplay:none;font-size:100%.mw-parser-output .cs1-visible-errorfont-size:100%.mw-parser-output .cs1-maintdisplay:none;color:#33aa33;margin-left:0.3em.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. ^ István Szabó, ;;Geschichte der mechanischen Prinzipien und ihrer wichtigsten Anwendungen, Basel: Birkhäuser Verlag, 1979.


  3. ^ Kirby, B. J. (2010). Micro- and Nanoscale Fluid Mechanics: Transport in Microfluidic Devices. Cambridge University Press. ISBN 978-0-521-11903-0.


  4. ^ Bruus, H. (2007). Theoretical Microfluidics.


  5. ^ "Poiseuille and his law" (PDF). pdfs.semanticscholar.org.


  6. ^ Vogel, Steven (1981). Life in Moving Fluids: The Physical Biology of Flow. PWS Kent Publishers. ISBN 0871507498.


  7. ^ Determinants of blood vessel resistance.


  8. ^ ab Batchelor, George Keith. An introduction to fluid dynamics. Cambridge university press, 2000.


  9. ^ Rosenhead, Louis, ed. Laminar boundary layers. Clarendon Press, 1963.


  10. ^ Drazin, Philip G., and Norman Riley. The Navier–Stokes equations: a classification of flows and exact solutions. No. 334. Cambridge University Press, 2006.


  11. ^ Boussinesq, Joseph. "Mémoire sur l’influence des Frottements dans les Mouvements Réguliers des Fluids." J. Math. Pures Appl 13.2 (1868): 377-424.


  12. ^ Proudman, J. "IV. Notes on the motion of viscous liquids in channels." The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 28.163 (1914): 30-36.


  13. ^ Samuel Newby Curle and H.J. Davies. Modern Fluid Dynamics. Volume 1, Incompressible Flow, Van Nostrand Reinhold Inc (1971)


  14. ^ Fütterer, C. et al. "Injection and flow control system for microchannels" Lab-on-a-Chip (2004): 351-356.




References[edit]



  • Sutera, S. P.; Skalak, R. (1993). "The history of Poiseuille's law". Annual Review of Fluid Mechanics. 25: 1–19. Bibcode:1993AnRFM..25....1S. doi:10.1146/annurev.fl.25.010193.000245.


  • Pfitzner, J (1976). "Poiseuille and his law". Anaesthesia. 31 (2) (published Mar 1976). pp. 273–5. doi:10.1111/j.1365-2044.1976.tb11804.x. PMID 779509.


  • Bennett, C. O.; Myers, J. E. (1962). "Momentum, Heat, and Mass Transfer". McGraw-Hill.


External links[edit]


  • Poiseuille's law for power-law non-Newtonian fluid

  • Poiseuille's law in a slightly tapered tube

  • Hagen–Poiseuille equation calculator



Retrieved from "https://en.wikipedia.org/w/index.php?title=Hagen–Poiseuille_equation&oldid=880186289"





Navigation menu


























(window.RLQ=window.RLQ||).push(function()mw.config.set("wgPageParseReport":"limitreport":"cputime":"0.480","walltime":"0.812","ppvisitednodes":"value":2226,"limit":1000000,"ppgeneratednodes":"value":0,"limit":1500000,"postexpandincludesize":"value":62808,"limit":2097152,"templateargumentsize":"value":2820,"limit":2097152,"expansiondepth":"value":9,"limit":40,"expensivefunctioncount":"value":3,"limit":500,"unstrip-depth":"value":1,"limit":20,"unstrip-size":"value":26774,"limit":5000000,"entityaccesscount":"value":2,"limit":400,"timingprofile":["100.00% 390.641 1 -total"," 39.44% 154.057 1 Template:Reflist"," 30.98% 121.035 3 Template:Cite_journal"," 16.40% 64.064 31 Template:Math"," 14.54% 56.786 1 Template:Continuum_mechanics"," 13.60% 53.127 1 Template:Sidebar_with_collapsible_lists"," 12.62% 49.315 2 Template:Unreferenced_section"," 11.38% 44.439 2 Template:Unreferenced"," 10.81% 42.239 1 Template:Val"," 10.22% 39.937 2 Template:Ambox"],"scribunto":"limitreport-timeusage":"value":"0.173","limit":"10.000","limitreport-memusage":"value":5711265,"limit":52428800,"cachereport":"origin":"mw1238","timestamp":"20190227050201","ttl":2592000,"transientcontent":false););"@context":"https://schema.org","@type":"Article","name":"Hagenu2013Poiseuille equation","url":"https://en.wikipedia.org/wiki/Hagen%E2%80%93Poiseuille_equation","sameAs":"http://www.wikidata.org/entity/Q869854","mainEntity":"http://www.wikidata.org/entity/Q869854","author":"@type":"Organization","name":"Contributors to Wikimedia projects","publisher":"@type":"Organization","name":"Wikimedia Foundation, Inc.","logo":"@type":"ImageObject","url":"https://www.wikimedia.org/static/images/wmf-hor-googpub.png","datePublished":"2008-04-20T03:58:17Z","dateModified":"2019-01-25T22:19:16Z"(window.RLQ=window.RLQ||).push(function()mw.config.set("wgBackendResponseTime":105,"wgHostname":"mw1257"););

Popular posts from this blog

𛂒𛀶,𛀽𛀑𛂀𛃧𛂓𛀙𛃆𛃑𛃷𛂟𛁡𛀢𛀟𛁤𛂽𛁕𛁪𛂟𛂯,𛁞𛂧𛀴𛁄𛁠𛁼𛂿𛀤 𛂘,𛁺𛂾𛃭𛃭𛃵𛀺,𛂣𛃍𛂖𛃶 𛀸𛃀𛂖𛁶𛁏𛁚 𛂢𛂞 𛁰𛂆𛀔,𛁸𛀽𛁓𛃋𛂇𛃧𛀧𛃣𛂐𛃇,𛂂𛃻𛃲𛁬𛃞𛀧𛃃𛀅 𛂭𛁠𛁡𛃇𛀷𛃓𛁥,𛁙𛁘𛁞𛃸𛁸𛃣𛁜,𛂛,𛃿,𛁯𛂘𛂌𛃛𛁱𛃌𛂈𛂇 𛁊𛃲,𛀕𛃴𛀜 𛀶𛂆𛀶𛃟𛂉𛀣,𛂐𛁞𛁾 𛁷𛂑𛁳𛂯𛀬𛃅,𛃶𛁼

Edmonton

Crossroads (UK TV series)