foundations of chemical kinetics lecture 26 diffusion
play

Foundations of Chemical Kinetics Lecture 26: Diffusion Marc R. - PowerPoint PPT Presentation

Foundations of Chemical Kinetics Lecture 26: Diffusion Marc R. Roussel Department of Chemistry and Biochemistry Mathematical background: Gradient The gradient operator is defined by x , y , = z


  1. Foundations of Chemical Kinetics Lecture 26: Diffusion Marc R. Roussel Department of Chemistry and Biochemistry

  2. Mathematical background: Gradient ◮ The gradient operator ∇ is defined by � ∂ ∂ x , ∂ ∂ y , ∂ � ∇ = ∂ z ◮ The gradient of a function � ∂ f ∂ x , ∂ f ∂ y , ∂ f � ∇ f = ∂ z is a vector that points in the direction of greatest increase of f .

  3. Mathematical background: Divergence ◮ The divergence of a vector is ∇ · v = ∂ v x ∂ x + ∂ v y ∂ y + ∂ v z ∂ z ◮ A positive value of the divergence at a point indicates that this point is a source of v ; a negative value indicates a sink.

  4. Mathematical background: Laplacian ◮ The Laplacian operator is ∇ · ∇ = ∇ 2 = ∂ 2 ∂ x 2 + ∂ 2 ∂ y 2 + ∂ 2 ∂ z 2 ◮ The Laplacian of a function is ∇ · ∇ f = ∇ 2 f = ∂ 2 f ∂ x 2 + ∂ 2 f ∂ y 2 + ∂ 2 f ∂ z 2

  5. Mathematical background: Differential operators and coordinate systems ◮ The gradient, divergence and Laplacian can be expressed in other coordinate systems. ◮ The formulas are generally non-trivial and should be looked up. ◮ Example: The Laplacian in spherical polar coordinates is ∂ 2 � � � � ∇ 2 = 1 ∂ r 2 ∂ 1 ∂ sin θ∂ 1 + + r 2 sin θ r 2 sin 2 θ r 2 ∂φ 2 ∂ r ∂ r ∂θ ∂θ i.e. ∂ 2 f ∇ 2 f = 1 ∂ � r 2 ∂ f � 1 ∂ � sin θ∂ f � 1 + + r 2 sin θ r 2 sin 2 θ r 2 ∂φ 2 ∂ r ∂ r ∂θ ∂θ

  6. Background: Chemical potential ◮ In thermodynamics, you would have learned the equation dG = V dp − S dT which tells us how G depends on T and p for a closed system. ◮ Recall that the following relations follow from this equation: � � ∂ G ∂ G � � = V = − S � � ∂ p ∂ T � � T p

  7. Background: Chemical potential (continued) ◮ What if we add a very small amount of substance i to the system? The chemical potential µ i is defined as the partial derivative of G with respect to n i : � ∂ G � = µ i � ∂ n i � p , T , n j � = i Note that, if we allow material to be added to the system, ∂ G /∂ p and ∂ G /∂ T must be evaluated holding all the n i constant. ◮ From the definition of the chemical potential, we get � dG = V dp − S dT + µ i dn i i where the sum is over all species in the system.

  8. Background: Chemical potential (continued) � dG = V dp − S dT + µ i dn i i ◮ Since G and n i are extensive variables (depend on the size of the system), µ i must be an intensive variable (doesn’t depend on the size). ◮ µ i can only depend on intensive variables ( p , T , c i , etc.).

  9. Background: Chemical potential (continued) � dG = V dp − S dT + µ i dn i i ◮ Suppose that we start with an empty container and reach the final state by adding material to the system holding p , T and all the c i constant, i.e. add all the system components in the correct, final proportions. Then, � dG = µ i dn i i � n i � � ∴ G = µ i dn i = µ i n i 0 i i

  10. Background: Chemical potential (continued) � n i � � G = µ i dn i = µ i n i 0 i i Interpretation: µ i is the portion of the Gibbs free energy of a system that can be attributed to species i . ◮ That being the case, it should not be very surprising that µ i obeys the equation µ i = µ ◦ i + RT ln a i where µ ◦ i is the chemical potential under standard conditions and a i is the activity of species i . ◮ Recall that a i = γ i c i / c ◦ for solutes.

  11. Flux ◮ Imagine a small (imaginary) surface of area dA immersed in a fluid. ◮ The net number of molecules passing through this surface in a specified direction per unit time divided by the area is the flux, J . ◮ If the surface is oriented perpendicular to the x axis and we count the net number of molecules travelling to the right (i.e. right-travelling minus left-travelling), then we have J x , the x component of the flux vector.

  12. The transport equation J ( x ) J x ( x + ∆ x ) x Area A ∆ x ◮ What is the rate of change of the concentration in the prism of the material whose flux is J x ? ◮ J x is the number of particles flowing through the surface per unit area per unit time, so ∆ n ∆ t = A [ J x ( x ) − J x ( x + ∆ x )] ◮ ∆ n / V is the change in concentration, and V = A ∆ x , so ∆ c ∆ t = [ J x ( x ) − J x ( x + ∆ x )] ∆ x

  13. The transport equation (continued) ∆ c ∆ t = [ J x ( x ) − J x ( x + ∆ x )] ∆ x ◮ If we let ∆ x → 0 and ∆ t → 0, we get ∂ c ∂ t = − ∂ J x ∂ x ◮ We only considered a flux along the x direction. If we consider the fluxes in the other directions, we get the transport equation ∂ c ∂ t = − ∂ J x ∂ x − ∂ J y ∂ y − ∂ J z ∂ z = −∇ · J

  14. Driving force of diffusion ◮ Suppose that we have an inhomogeneous system, so that the composition, and therefore the chemical potential, varies from point to point. ◮ Suppose that we take one molecule of substance i from the vicinity of position x along the x axis, and move it to position x + dx . The corresponding change in free energy is dG = [ µ i ( x + dx ) − µ i ( x )] / L ( µ i is in J mol − 1 , so division by L gives us the chemical potential per molecule.)

  15. Driving force of diffusion (continued) ◮ Because the negative of the free energy change is the maximum work, we have dw = − [ µ i ( x + dx ) − µ i ( x )] / L ◮ From the definition of work, dw = F i , x dx , so F i , x = − 1 µ i ( x + dx ) − µ i ( x ) L dx or, in the limit as dx → 0, F i , x = − 1 ∂µ i L ∂ x

  16. Driving force of diffusion (continued) ◮ For an ideal solute, µ i = µ ◦ i + RT ln( c i ( x ) / c ◦ ) ∴ F i , x = − RT ∂ c i ∂ x = − k B T ∂ c i ∂ x Lc i c i or, in three dimensions F i = − k B T ∇ c i c i ◮ The negative of this virtual force is the opposing force we would have to apply to prevent diffusion from occurring. ◮ Note that the force is directed opposite to the gradient, i.e. diffusion is equivalent to a force pushing the molecules toward regions of low concentration.

  17. Driving force of diffusion (continued) ◮ A constantly applied force would cause molecules to accelerate. ◮ The virtual diffusion force is opposed by drag. ◮ At low speeds, the drag force is F ( d ) = − f i v i where f i is a i frictional (drag) coefficient and v i is the mean drift velocity. ◮ F ( d ) should be equal to the negative of F i , i.e. i = − f i v i = k B T F ( d ) ∇ c i i c i ◮ The flux (molecules per unit area per unit time) is (molecules per unit volume) × (distance travelled per unit time), i.e. J i = c i v i . Thus, J i = − k B T ∇ c i f i

  18. Driving force of diffusion (continued) J i = − k B T ∇ c i f i ◮ The drag coefficient f i depends on the solute and solvent environment. ◮ Define the diffusion coefficient D i = k B T / f i so that J i = − D i ∇ c i which is known as Fick’s first law. ◮ Fick’s first law states that the flux of species i runs in the opposite direction to its gradient.

  19. Driving force of diffusion (continued) ◮ The derivation of Fick’s first law involves a number of assumptions, some obvious, some less so. Accordingly, Fick’s first law has a limited range of applicability (negligible forces between solutes, gradients not too large).

  20. The diffusion equation Transport equation: ∂ c i ∂ t = −∇ · J i Fick’s first law: J i = − D i ∇ c i ◮ Putting these two together, we get the diffusion equation, also known as Fick’s second law: ∂ c i ∂ t = D i ∇ · ∇ c i = D i ∇ 2 c i

  21. The diffusion equation (continued) ◮ In one dimension: ∂ 2 c i ∂ c i ∂ t = D i ∂ x 2 c x

  22. Example: Diffusive spread in one dimension ◮ Suppose that we put a small drop of material (e.g. a dye) at the origin in a very long, narrow tube. ◮ A drop of negligible width can be modeled as a delta function, with the following properties: ◮ δ ( x ) = 0 everywhere except at x = 0. ◮ � a − a δ ( x ) dx = 1 for any a > 0. ◮ Take c ( x , t = 0) = n A δ ( x ), where n is the total number of moles of the material and A is the cross-sectional area of the tube. ◮ Solution of the diffusion equation requires special techniques. Many standard cases have already been solved, and the solutions can simply be looked up. ◮ In this case, the solution is − x 2 n � � c ( x , t ) = exp √ 4 Dt 2 A π Dt

  23. Example: Diffusive spread in one dimension (continued) − x 2 n � � c ( x , t ) = exp √ 4 Dt 2 A π Dt Note: This is a Gaussian with time-dependent variance 2 Dt .

  24. Example: Diffusive spread in one dimension (continued) ◮ Diffusion coefficient of sucrose in water at 20 ◦ C: 5 . 7 × 10 − 10 m 2 s − 1 ◮ Assume a cylindrical pipe with an internal diameter of 5.00 mm, giving A = 1 . 96 × 10 − 5 m 2 , and n = 1 µ mol. 0.7 t = 1 s 10 0.6 100 0.5 c /mol L -1 0.4 0.3 0.2 0.1 0 -1 -0.5 0 0.5 1 x /mm

Download Presentation
Download Policy: The content available on the website is offered to you 'AS IS' for your personal information and use only. It cannot be commercialized, licensed, or distributed on other websites without prior consent from the author. To download a presentation, simply click this link. If you encounter any difficulties during the download process, it's possible that the publisher has removed the file from their server.

Recommend


More recommend