fbpx
Wikipedia

Darken's equations

In metallurgy, the Darken equations are used to describe the solid-state diffusion of materials in binary solutions. They were first described by Lawrence Stamper Darken in 1948.[1] The equations apply to cases where a solid solution's two components do not have the same coefficient of diffusion.

The equations

Darken's first equation is:

 

where:

  •   is the marker velocity of inert markers showing the diffusive flux.
  •   and   are the diffusion coefficients of the two components.
  •   and   are the atomic fractions of the two components.
  •   represents the direction in which the diffusion is measured.

It is important to note that this equation only holds in situations where the total concentration remains constant.

Darken's second equation is:

 

where:

  •   is the activity coefficient of the first component.
  •   is the overall diffusivity of the binary solution.

Experimental methods

In deriving the first equation, Darken referenced Simgelskas and Kirkendall's experiment, which tested the mechanisms and rates of diffusion and gave rise to the concept now known as the Kirkendall effect.[2] For the experiment, inert molybdenum wires were placed at the interface between copper and brass components, and the motion of the markers was monitored. The experiment supported the concept that a concentration gradient in a binary alloy would result in the different components having different velocities in the solid solution. The experiment showed that in brass zinc had a faster relative velocity than copper, since the molybdenum wires moved farther into the brass. In establishing the coordinate axes to evaluate the derivation, Darken refers back to Smigelskas and Kirkendall’s experiment which the inert wires were designated as the origin.[1]

In respect to the derivation of the second equation, Darken referenced W. A. Johnson’s experiment on a gold–silver system, which was performed to determine the chemical diffusivity. In this experiment radioactive gold and silver isotopes were used to measure the diffusivity of gold and silver, because it was assumed that the radioactive isotopes have relatively the same mobility as the non-radioactive elements. If the gold–silver solution is assumed to behave ideally, it would be expected the diffusivities would also be equivalent. Therefore, the overall diffusion coefficient of the system would be the average of each components diffusivity; however, this was found not to be true.[1] This finding led Darken to analyze Johnson's experiment and derive the equation for chemical diffusivity of binary solutions.

Darken's first equation

Background

As stated previously, Darken's first equation allows the calculation of the marker velocity   in respect to a binary system where the two components have different diffusion coefficients. For this equation to be applicable, the analyzed system must have a constant concentration and can be modeled by the Boltzmann–Matano solution.

For the derivation, a hypothetical case is considered where two homogeneous binary alloy rods of two different compositions are in contact. The sides are protected, so that all of the diffusion occurs parallel to the length of the rod. In establishing the coordinate axes to evaluate the derivation, Darken sets the x-axis to be fixed at the far ends of the rods, and the origin at the initial position of the interface between the two rods. In addition this choice of a coordinate system allows the derivation to be simplified, whereas Smigelskas and Kirkendall's coordinate system was considered to be the non-optimal choice for this particular calculation as can be seen in the following section. At the initial planar interface between the rods, it is considered that there are infinitely small inert markers placed in a plane which is perpendicular to the length of the rods. Here, inert markers are defined to be a group of particles that are of a different elemental make-up from either of the diffusing components and move in the same fashion. For this derivation, the inert markers are assumed to be following the motion of the crystal lattice. The motion relative to the marker is associated with diffusion,  , while the motion of the markers is associated with advection,  . Fick’s first law, the previous equation stated for diffusion, describes the entirety of the system for only small distances from the origin, since at large distances advection needs to be accounted for. This results in the total rate of transport for the system being influenced by both factors, diffusion and advection.[1]

Derivation

The derivation starts with Fick's first law using a uniform distance axis y as the coordinate system and having the origin fixed to the location of the markers. It is assumed that the markers move relative to the diffusion of one component and into one of the two initial rods, as was chosen in Kirkendall's experiment. In the following equation, which represents Fick's first law for one of the two components, D1 is the diffusion coefficient of component one, and C1 is the concentration of component one:

 

This coordinate system only works for short range from the origin because of the assumption that marker movement is indicative of diffusion alone, which is not true for long distances from the origin as stated before. The coordinate system is transformed using a Galilean transformation, y = x − νt, where x is the new coordinate system that is fixed to the ends of the two rods, ν is the marker velocity measured with respect to the x axis. The variable t, time, is assumed to be constant, so that the partial derivative of C1 with respect to y is equal to the partial of C1 with respect to x. This transformation then yields

 

The above equation, in terms of the variable x, only takes into account diffusion, so the term for the motion of the markers must also be included, since the frame of reference is no longer moving with the marker particles. In the equation below,   is the velocity of the markers.

 

Taking the above equation and then equating it to the accumulation rate in a volume results in the following equation. This result is similar to Fick's second law, but with an additional advection term:

 

The same equation can be written for the other component, designated as component two:

 

Using the assumption that C, the total concentration, is constant,[3] C1 and C2 can be related in the following expression:

 

The above equation can then be used to combine the expressions for   and   to yield

 

Since C is constant, the above equation can be written as

 

The above equation states that   is constant because the derivative of a constant is equal to zero. Therefore, by integrating the above equation it is transforms to  , where   is an integration constant.

At relative infinite distances from the initial interface, the concentration gradients of each of the components and the marker velocity can be assumed to be equal to zero. Based on this condition and the choice for the coordinate axis, where the x axis fixed at the far ends of the rods, I is equal zero.[4] These conditions then allow the equation to be rearranged to give

 

Since C is assumed to be constant,  . Rewriting this equation in terms of atom fraction   and   yields[1]

 

Accompanying derivation

Referring back to the derivation for Darken's first equation,   is written as

 

Inserting this value for   in   gives

 

As stated before,  , which gives

 

Rewriting this equation in terms of atom fraction   and   yields

 

By using   and solving to the form  , it is found that

 

Integrating the above gives the final equation:

 

This equation is only applicable for binary systems that follow the equations of state and the Gibbs–Duhem equation. This equation, as well as Darken's first law,  , gives a complete description of an ideal binary diffusion system.[1] This derivation was the approach taken by Darken in his original 1948, though shorter methods can be used to attain the same result.

Darken's second equation

Background

Darken's second equation relates the chemical diffusion coefficient,  , of a binary system to the atomic fractions of the two components. Similar to the first equation, this equation is applicable when the system does not undergo a volume change. This equation also only applies to multicomponent systems, including binary systems, that obey the equations of state and the Gibbs–Duhem equations.

Derivation

To derive Darken's second equation the gradient in Gibb's chemical potential is analyzed. The gradient in potential energy, denoted by F2, is the force which causes atoms to diffuse.[1] To begin, the flux J is equated to the product of the differential of the gradient and the mobility B, which is defined as the diffusing atom's velocity per unit of applied force.[5] In addition, NA is the Avogadro constant, and C2 is the concentration of diffusing component two. This yields

 

which can be equated to the expression for Fick's first law:

 

so that the expression can be written as

 

After some rearrangement of variables the expression can be written for D2, the diffusivity of component two:

 

Assuming that atomic volume is constant, so C = C1 + C2,

 

Using a definition activity,  , where R is the gas constant, and T is the temperature, to rewrite the equation in terms of activity gives

 

The above equation can be rewritten in terms of the activity coefficient γ, which is defined in terms of activity by the equation  . This yields

 

The same equation can also be written for the diffusivity of component one,  , and combining the equations for D1 and D2 gives the final equation:[1]

 

Applications

Darken’s equations can be applied to almost any scenario involving the diffusion of two different components that have different diffusion coefficients. This holds true except in situations where there is an accompanying volume change in the material because this violates one of Darken’s critical assumptions that atomic volume is constant. More complicated equations than presented must be used in cases where there is convection. One application in which Darken’s equations play an instrumental role is in analyzing the process of diffusion bonding.[6] Diffusion bonding is used widely in manufacturing to connect two materials without using adhesives or welding techniques. Diffusion bonding works because atoms from both materials diffuse into the other material, resulting in a bond that is formed between the two materials. The diffusion of atoms between the two materials is achieved by placing the materials in contact with each other at high pressure and temperature, while not exceeding the melting temperature of either material. Darken’s equations, particularly Darken’s second equation, come into play when determining the diffusion coefficients for the two materials in the diffusion couple. Knowing the diffusion coefficients is necessary for predicting the flux of atoms between the two materials, which can then be used in numerical models of the diffusion bonding process, as, for example, was looked at in the paper by Orhan, Aksoy, and Eroglu when creating a model to determine the amount of time required to create a diffusion bond.[6] In a similar manner, Darken’s equations were used in a paper by Watanabe et al., on the nickel-aluminum system, to verify the interdiffusion coefficients that were calculated for nickel aluminum alloys.[7]

Application of Darken’s first equation has important implications for analyzing the structural integrity of materials. Darken’s first equation,  , can be rewritten in terms of vacancy flux,  .[8] Use of Darken’s equation in this form has important implications for determining the flux of vacancies into a material undergoing diffusion bonding, which, due to the Kirkendall effect, could lead to porosity in the material and have an adverse effect on its strength. This is particularly important in materials such as aluminum nickel superalloys that are used in jet engines, where the structural integrity of the materials is extremely important. Porosity formation, known as Kirkendall porosity, in these nickel-aluminum superalloys have been observed when diffusion bonding has been used.[9][10] It is important then to use Darken’s findings to predict this porosity formation.

See also

References

  1. ^ a b c d e f g h Darken, L. S. "Diffusion, mobility and their interrelation through free energy in binary metallic systems". Trans. AIME 175.1 (1948): 184–194.
  2. ^ Smigelskas, A. D., and E. O. Kirkendall. "Zinc diffusion in alpha brass". Trans. AIME 171 (1947): 130–142.
  3. ^ Sekerka, R. F. "Similarity Solutions for a Binary Diffusion Couple with Diffusivity and Density Dependent on Composition". Progress in Materials Science 49 (2004): 511–536.
  4. ^ Glicksman, Martin E. Diffusion in Solids: Field theory, Solid-Stat Principles, and Applications. New York: John Wiley and Sons, 2000.
  5. ^ Gaskell, David R. An Introduction to: Transport Phenomena in Materials Engineering. 2nd ed. New York; Momentum Press, 2012.
  6. ^ a b Orhan, N , M Aksoy, and M Eroglu. "A new model for diffusion bonding and its application to duplex alloys." Materials Science and Engineering 271.1-2 (1999): 458-468. Science Direct. Web.
  7. ^ Watanabe, M., Z. Horita, T. Sano, and M. Nemoto. "Electron microscopy study of Ni/Ni3Al diffusion-couple interface-II. Diffusivity measurement." Acta Metallurgica et Materialia 42.10 (1994): 3389-3396. Science Direct. Web.
  8. ^ "DoITPoMS - TLP Library Diffusion - Derivation of darken equation".
  9. ^ Karunaratne, M.S.A, P. Carter, and R.C. Reed. "On the diffusion of aluminium and titanium in the Ni-rich Ni–Al–Ti system between 900 and 1200°C." Acta Materialia 49.5 (2001): 861-875. Science Direct. Web.
  10. ^ Janssen, M.M.P.. "Diffusion in the nickel-rich part of the Ni−Al system at 1000° to 1300°C; Ni3Al layer growth, diffusion coefficients, and interface concentrations." Metallurgical Transactions 4.6 (1973): 1623-1633.Springer Link. Web.

darken, equations, this, article, needs, attention, from, expert, physics, specific, problem, check, factual, accuracy, wikiproject, physics, able, help, recruit, expert, july, 2013, this, article, technical, most, readers, understand, please, help, improve, m. This article needs attention from an expert in Physics The specific problem is check factual accuracy WikiProject Physics may be able to help recruit an expert July 2013 This article may be too technical for most readers to understand Please help improve it to make it understandable to non experts without removing the technical details May 2018 Learn how and when to remove this template message In metallurgy the Darken equations are used to describe the solid state diffusion of materials in binary solutions They were first described by Lawrence Stamper Darken in 1948 1 The equations apply to cases where a solid solution s two components do not have the same coefficient of diffusion Contents 1 The equations 2 Experimental methods 3 Darken s first equation 3 1 Background 3 2 Derivation 3 3 Accompanying derivation 4 Darken s second equation 4 1 Background 4 2 Derivation 5 Applications 6 See also 7 ReferencesThe equations EditDarken s first equation is n D 1 D 2 N 1 x D 2 D 1 N 2 x displaystyle nu D 1 D 2 frac partial N 1 partial x D 2 D 1 frac partial N 2 partial x where n displaystyle nu is the marker velocity of inert markers showing the diffusive flux D 1 displaystyle D 1 and D 2 displaystyle D 2 are the diffusion coefficients of the two components N 1 displaystyle N 1 and N 2 displaystyle N 2 are the atomic fractions of the two components x displaystyle x represents the direction in which the diffusion is measured It is important to note that this equation only holds in situations where the total concentration remains constant Darken s second equation is D N 1 D 2 N 2 D 1 1 N 1 ln a 1 ln N 1 displaystyle tilde D N 1 D 2 N 2 D 1 1 N 1 frac partial ln a 1 partial ln N 1 where a 1 displaystyle a 1 is the activity coefficient of the first component D displaystyle tilde D is the overall diffusivity of the binary solution Experimental methods EditIn deriving the first equation Darken referenced Simgelskas and Kirkendall s experiment which tested the mechanisms and rates of diffusion and gave rise to the concept now known as the Kirkendall effect 2 For the experiment inert molybdenum wires were placed at the interface between copper and brass components and the motion of the markers was monitored The experiment supported the concept that a concentration gradient in a binary alloy would result in the different components having different velocities in the solid solution The experiment showed that in brass zinc had a faster relative velocity than copper since the molybdenum wires moved farther into the brass In establishing the coordinate axes to evaluate the derivation Darken refers back to Smigelskas and Kirkendall s experiment which the inert wires were designated as the origin 1 In respect to the derivation of the second equation Darken referenced W A Johnson s experiment on a gold silver system which was performed to determine the chemical diffusivity In this experiment radioactive gold and silver isotopes were used to measure the diffusivity of gold and silver because it was assumed that the radioactive isotopes have relatively the same mobility as the non radioactive elements If the gold silver solution is assumed to behave ideally it would be expected the diffusivities would also be equivalent Therefore the overall diffusion coefficient of the system would be the average of each components diffusivity however this was found not to be true 1 This finding led Darken to analyze Johnson s experiment and derive the equation for chemical diffusivity of binary solutions Darken s first equation EditBackground Edit As stated previously Darken s first equation allows the calculation of the marker velocity n displaystyle nu in respect to a binary system where the two components have different diffusion coefficients For this equation to be applicable the analyzed system must have a constant concentration and can be modeled by the Boltzmann Matano solution For the derivation a hypothetical case is considered where two homogeneous binary alloy rods of two different compositions are in contact The sides are protected so that all of the diffusion occurs parallel to the length of the rod In establishing the coordinate axes to evaluate the derivation Darken sets the x axis to be fixed at the far ends of the rods and the origin at the initial position of the interface between the two rods In addition this choice of a coordinate system allows the derivation to be simplified whereas Smigelskas and Kirkendall s coordinate system was considered to be the non optimal choice for this particular calculation as can be seen in the following section At the initial planar interface between the rods it is considered that there are infinitely small inert markers placed in a plane which is perpendicular to the length of the rods Here inert markers are defined to be a group of particles that are of a different elemental make up from either of the diffusing components and move in the same fashion For this derivation the inert markers are assumed to be following the motion of the crystal lattice The motion relative to the marker is associated with diffusion D 1 C 1 y displaystyle D 1 tfrac partial C 1 partial y while the motion of the markers is associated with advection C 1 n displaystyle C 1 nu Fick s first law the previous equation stated for diffusion describes the entirety of the system for only small distances from the origin since at large distances advection needs to be accounted for This results in the total rate of transport for the system being influenced by both factors diffusion and advection 1 Derivation Edit The derivation starts with Fick s first law using a uniform distance axis y as the coordinate system and having the origin fixed to the location of the markers It is assumed that the markers move relative to the diffusion of one component and into one of the two initial rods as was chosen in Kirkendall s experiment In the following equation which represents Fick s first law for one of the two components D1 is the diffusion coefficient of component one and C1 is the concentration of component one D 1 C 1 y displaystyle D 1 frac partial C 1 partial y This coordinate system only works for short range from the origin because of the assumption that marker movement is indicative of diffusion alone which is not true for long distances from the origin as stated before The coordinate system is transformed using a Galilean transformation y x nt where x is the new coordinate system that is fixed to the ends of the two rods n is the marker velocity measured with respect to the x axis The variable t time is assumed to be constant so that the partial derivative of C1 with respect to y is equal to the partial of C1 with respect to x This transformation then yields D 1 C 1 x displaystyle D 1 frac partial C 1 partial x The above equation in terms of the variable x only takes into account diffusion so the term for the motion of the markers must also be included since the frame of reference is no longer moving with the marker particles In the equation below n displaystyle nu is the velocity of the markers D 1 C 1 x C 1 n displaystyle left D 1 frac partial C 1 partial x C 1 nu right Taking the above equation and then equating it to the accumulation rate in a volume results in the following equation This result is similar to Fick s second law but with an additional advection term C 1 t x D 1 C 1 x C 1 n displaystyle frac partial C 1 partial t frac partial partial x left D 1 frac partial C 1 partial x C 1 nu right The same equation can be written for the other component designated as component two C 2 t x D 2 C 2 x C 2 n displaystyle frac partial C 2 partial t frac partial partial x left D 2 frac partial C 2 partial x C 2 nu right Using the assumption that C the total concentration is constant 3 C1 and C2 can be related in the following expression C C 1 C 2 displaystyle C C 1 C 2 The above equation can then be used to combine the expressions for C 1 t displaystyle tfrac partial C 1 partial t and C 2 t displaystyle tfrac partial C 2 partial t to yield C t x D 1 C 1 x D 2 C 2 x C n displaystyle frac partial C partial t frac partial partial x left D 1 frac partial C 1 partial x D 2 frac partial C 2 partial x C nu right Since C is constant the above equation can be written as 0 x D 1 C 1 x D 2 C 2 x C n displaystyle 0 frac partial partial x left D 1 frac partial C 1 partial x D 2 frac partial C 2 partial x C nu right The above equation states that D 1 C 1 x D 2 C 2 x C n displaystyle textstyle D 1 frac partial C 1 partial x D 2 frac partial C 2 partial x C nu is constant because the derivative of a constant is equal to zero Therefore by integrating the above equation it is transforms to D 1 C 1 x D 2 C 2 x C n I displaystyle textstyle D 1 frac partial C 1 partial x D 2 frac partial C 2 partial x C nu I where I displaystyle I is an integration constant At relative infinite distances from the initial interface the concentration gradients of each of the components and the marker velocity can be assumed to be equal to zero Based on this condition and the choice for the coordinate axis where the x axis fixed at the far ends of the rods I is equal zero 4 These conditions then allow the equation to be rearranged to give n 1 C D 1 C 1 x D 2 C 2 x displaystyle nu frac 1 C left D 1 frac partial C 1 partial x D 2 frac partial C 2 partial x right Since C is assumed to be constant C 1 x C 2 x displaystyle textstyle frac partial C 1 partial x frac partial C 2 partial x Rewriting this equation in terms of atom fraction N 1 C 1 C displaystyle N 1 tfrac C 1 C and N 2 C 2 C displaystyle N 2 tfrac C 2 C yields 1 n D 1 D 2 N 1 x D 2 D 1 N 2 x displaystyle nu D 1 D 2 frac partial N 1 partial x D 2 D 1 frac partial N 2 partial x Accompanying derivation Edit Referring back to the derivation for Darken s first equation n displaystyle nu is written as n 1 C D 1 C 1 x D 2 C 2 x displaystyle nu frac 1 C left D 1 frac partial C 1 partial x D 2 frac partial C 2 partial x right Inserting this value for n displaystyle nu in C t x D 1 C 1 x C 1 n displaystyle textstyle frac partial C partial t frac partial partial x left D 1 frac partial C 1 partial x C 1 nu right gives C 1 t x D 1 C 1 x C 1 C D 1 C 1 x D 2 C 2 x displaystyle frac partial C 1 partial t frac partial partial x left D 1 frac partial C 1 partial x frac C 1 C left D 1 frac partial C 1 partial x D 2 frac partial C 2 partial x right right As stated before C 1 x C 2 x displaystyle textstyle frac partial C 1 partial x frac partial C 2 partial x which gives C 1 t x C 1 C 2 C D 1 C 1 x C 1 C D 1 C 1 x D 2 C 1 x displaystyle frac partial C 1 partial t frac partial partial x left frac C 1 C 2 C D 1 frac partial C 1 partial x frac C 1 C left D 1 frac partial C 1 partial x D 2 frac partial C 1 partial x right right Rewriting this equation in terms of atom fraction N 1 C 1 C displaystyle N 1 tfrac C 1 C and N 2 C 2 C displaystyle N 2 tfrac C 2 C yields N 1 t x N 2 D 1 N 1 D 2 N 1 x displaystyle frac partial N 1 partial t frac partial partial x left N 2 D 1 N 1 D 2 frac partial N 1 partial x right By using l x t 1 2 displaystyle lambda equiv tfrac x t 1 2 and solving to the form N 1 f l displaystyle N 1 f lambda it is found that 1 2 l d N 1 d N 2 D 1 N 1 D 2 d N 1 d l displaystyle frac 1 2 lambda dN 1 d N 2 D 1 N 1 D 2 frac dN 1 d lambda Integrating the above gives the final equation D D 1 N 2 D 2 N 1 displaystyle D D 1 N 2 D 2 N 1 This equation is only applicable for binary systems that follow the equations of state and the Gibbs Duhem equation This equation as well as Darken s first law n D 2 D 1 N 2 x displaystyle nu D 2 D 1 tfrac partial N 2 partial x gives a complete description of an ideal binary diffusion system 1 This derivation was the approach taken by Darken in his original 1948 though shorter methods can be used to attain the same result Darken s second equation EditBackground Edit Darken s second equation relates the chemical diffusion coefficient D displaystyle tilde D of a binary system to the atomic fractions of the two components Similar to the first equation this equation is applicable when the system does not undergo a volume change This equation also only applies to multicomponent systems including binary systems that obey the equations of state and the Gibbs Duhem equations Derivation Edit To derive Darken s second equation the gradient in Gibb s chemical potential is analyzed The gradient in potential energy denoted by F2 is the force which causes atoms to diffuse 1 To begin the flux J is equated to the product of the differential of the gradient and the mobility B which is defined as the diffusing atom s velocity per unit of applied force 5 In addition NA is the Avogadro constant and C2 is the concentration of diffusing component two This yields J 1 N A d F 2 d x B 2 C 2 displaystyle J frac 1 N A frac dF 2 dx B 2 C 2 which can be equated to the expression for Fick s first law D 2 d C 2 d x displaystyle D 2 frac dC 2 dx so that the expression can be written as D 2 d C 2 d x 1 N A d F 2 d x B 2 C 2 displaystyle D 2 frac dC 2 dx frac 1 N text A frac dF 2 dx B 2 C 2 After some rearrangement of variables the expression can be written for D2 the diffusivity of component two D 2 d F 2 d C 2 B 2 C 2 N A displaystyle D 2 frac dF 2 dC 2 frac B 2 C 2 N text A Assuming that atomic volume is constant so C C1 C2 1 N A d F 2 d N 2 B 2 N 2 displaystyle frac 1 N text A frac dF 2 dN 2 B 2 N 2 Using a definition activity d F 2 R T d ln a 2 displaystyle dF 2 RT d ln a 2 where R is the gas constant and T is the temperature to rewrite the equation in terms of activity gives D 2 k T B 2 d ln a 2 d ln N 2 displaystyle D 2 kTB 2 frac d ln a 2 d ln N 2 The above equation can be rewritten in terms of the activity coefficient g which is defined in terms of activity by the equation g 2 a 2 N 2 displaystyle gamma 2 a 2 N 2 This yields D 2 k T B 2 1 N 2 d ln g 2 d ln N 2 displaystyle D 2 kTB 2 left 1 N 2 frac d ln gamma 2 d ln N 2 right The same equation can also be written for the diffusivity of component one D 1 k T B 1 1 N 1 d ln g 1 d ln N 1 displaystyle D 1 kTB 1 left 1 N 1 tfrac d ln gamma 1 d ln N 1 right and combining the equations for D1 and D2 gives the final equation 1 D N 1 D 2 N 2 D 1 ln a 1 ln N 1 displaystyle tilde D N 1 D 2 N 2 D 1 frac partial ln a 1 partial ln N 1 Applications EditDarken s equations can be applied to almost any scenario involving the diffusion of two different components that have different diffusion coefficients This holds true except in situations where there is an accompanying volume change in the material because this violates one of Darken s critical assumptions that atomic volume is constant More complicated equations than presented must be used in cases where there is convection One application in which Darken s equations play an instrumental role is in analyzing the process of diffusion bonding 6 Diffusion bonding is used widely in manufacturing to connect two materials without using adhesives or welding techniques Diffusion bonding works because atoms from both materials diffuse into the other material resulting in a bond that is formed between the two materials The diffusion of atoms between the two materials is achieved by placing the materials in contact with each other at high pressure and temperature while not exceeding the melting temperature of either material Darken s equations particularly Darken s second equation come into play when determining the diffusion coefficients for the two materials in the diffusion couple Knowing the diffusion coefficients is necessary for predicting the flux of atoms between the two materials which can then be used in numerical models of the diffusion bonding process as for example was looked at in the paper by Orhan Aksoy and Eroglu when creating a model to determine the amount of time required to create a diffusion bond 6 In a similar manner Darken s equations were used in a paper by Watanabe et al on the nickel aluminum system to verify the interdiffusion coefficients that were calculated for nickel aluminum alloys 7 Application of Darken s first equation has important implications for analyzing the structural integrity of materials Darken s first equation v D 2 D 1 N 2 x displaystyle textstyle v D 2 D 1 frac partial N 2 partial x can be rewritten in terms of vacancy flux J v D 2 D 1 N 2 x displaystyle textstyle J v D 2 D 1 frac partial N 2 partial x 8 Use of Darken s equation in this form has important implications for determining the flux of vacancies into a material undergoing diffusion bonding which due to the Kirkendall effect could lead to porosity in the material and have an adverse effect on its strength This is particularly important in materials such as aluminum nickel superalloys that are used in jet engines where the structural integrity of the materials is extremely important Porosity formation known as Kirkendall porosity in these nickel aluminum superalloys have been observed when diffusion bonding has been used 9 10 It is important then to use Darken s findings to predict this porosity formation See also EditGibbs Duhem equation Ternary and multicomponent solutions and mixturesReferences Edit a b c d e f g h Darken L S Diffusion mobility and their interrelation through free energy in binary metallic systems Trans AIME 175 1 1948 184 194 Smigelskas A D and E O Kirkendall Zinc diffusion in alpha brass Trans AIME 171 1947 130 142 Sekerka R F Similarity Solutions for a Binary Diffusion Couple with Diffusivity and Density Dependent on Composition Progress in Materials Science 49 2004 511 536 Glicksman Martin E Diffusion in Solids Field theory Solid Stat Principles and Applications New York John Wiley and Sons 2000 Gaskell David R An Introduction to Transport Phenomena in Materials Engineering 2nd ed New York Momentum Press 2012 a b Orhan N M Aksoy and M Eroglu A new model for diffusion bonding and its application to duplex alloys Materials Science and Engineering 271 1 2 1999 458 468 Science Direct Web Watanabe M Z Horita T Sano and M Nemoto Electron microscopy study of Ni Ni3Al diffusion couple interface II Diffusivity measurement Acta Metallurgica et Materialia 42 10 1994 3389 3396 Science Direct Web DoITPoMS TLP Library Diffusion Derivation of darken equation Karunaratne M S A P Carter and R C Reed On the diffusion of aluminium and titanium in the Ni rich Ni Al Ti system between 900 and 1200 C Acta Materialia 49 5 2001 861 875 Science Direct Web Janssen M M P Diffusion in the nickel rich part of the Ni Al system at 1000 to 1300 C Ni3Al layer growth diffusion coefficients and interface concentrations Metallurgical Transactions 4 6 1973 1623 1633 Springer Link Web Retrieved from https en wikipedia org w index php title Darken 27s equations amp oldid 1119570652, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.