throbber
Journal of The Electrochemical Society, 153 共6兲 K15-K22 共2006兲
`0013-4651/2006/153共6兲/K15/8/$20.00 © The Electrochemical Society
`Hydrodynamics of Slurry Flow in Chemical Mechanical
`Polishing
`A Review
`Elon J. Terrell and C. Fred Higgs IIIz
`
`Department of Mechanical Engineering, Carnegie Mellon University, Pittsburgh,
`Pennsylvania 15213-3890, USA
`
`K15
`
`Chemical mechanical polishing 共CMP兲 is a process that is commonly used to planarize wafer surfaces during fabrication. Although
`the complex interactions between the wafer, pad, and slurry make the CMP process difficult to predict, it has been postulated that
`the motion of the slurry fluid at the wafer–pad interface has an important effect on the wafer surface wear distribution. This paper
`thus serves as a review of past studies of the hydrodynamics of slurry flow during chemical mechanical polishing. The reviewed
`studies include theoretical and numerical models as well as experimental measurements.
`关DOI: 10.1149/1.2188329兴 All rights reserved.
`© 2006 The Electrochemical Society.
`
`Manuscript submitted August 30, 2005; revised manuscript received January 23, 2006. Available electronically April 19, 2006.
`
`in partial contact and have used a hybrid contact mechanics/fluid
`mechanics approach toward analyzing CMP. Finally, a set of studies
`have analyzed the CMP process solely using fluid mechanics, as-
`suming that the wafer and pad surfaces are completely separated by
`slurry. From their review, Nanz and Camilletti have indicated the
`importance of the slurry flowfield to the CMP process as well as the
`need for more in-depth understanding of slurry flow at the wafer–
`pad interface. Therefore this paper serves as a review of past studies
`in slurry hydrodynamics during CMP. These studies include film
`thickness and hydrodynamic pressure modeling, numerical fluid
`flow modeling, and experimental investigations.
`
`CMP Hydrodynamic Modeling
`In the modeling studies that are discussed in this paper, slurry
`hydrodynamic analysis was used to find expressions for a number of
`parameters, including the slurry pressure field, film thickness distri-
`bution, and shear rate.
`
`Slurry Film Thickness and Hydrodynamic Pressure Modeling
`Assuming that a thin slurry film separates the wafer and pad
`surfaces during CMP, the film thickness and pressure distribution of
`the slurry can be related using the Reynolds equation, shown in 1D
`form as follows
`
`关2兴
`
`冉h3dp
`
`冊 = 6␮U
`
`d d
`
`dh
`dx
`dx
`x
`where p is the hydrodynamic pressure, h is the local film thickness,
`␮ is the dynamic viscosity of the slurry, U is the relative velocity of
`the bottom surface, and x is the downstream distance. Analysis of
`the slurry pressure distribution is of importance in CMP hydrody-
`namic studies because it gives insight into how much the wafer and
`pad surfaces are being pushed away from each other 共positive pres-
`sure兲 or sucked toward each other 共negative pressure兲 along the
`length of the interface.
`A study by Sundararajan et al.2 involved the derivation of 2D
`wafer-scale lubrication and mass-transport models for an assumed
`hydrodynamic slurry interface during CMP. They started the model
`using the one-dimensional, steady-state Reynolds equation 共Eq. 2兲,
`assuming that the film thickness had a convex shape due to bending
`of the wafer. It is important to note the difference between a convex
`and concave wafer according to CMP terminology. As shown in
`Fig. 2, a wafer is termed convex if it is bent toward the pad in the
`middle and concave if it is bent away from the pad in the middle.
`Certain constant parameters in the film thickness expression were
`solved by assuming two constraints: 共i兲 that the integral of the pres-
`sure distribution across the length of the wafer is equal to the ap-
`plied load, and 共ii兲 that the movement of the forces around the center
`of the wafer is zero. After solving for the film thickness and pressure
`distributions, the authors calculated the slurry velocity distribution
`
`Chemical mechanical polishing 共CMP兲 is a manufacturing pro-
`cess that is used to planarize the surfaces of small-scale devices such
`as integrated circuits and hard disk read/write heads during fabrica-
`tion. CMP has emerged as a critical fabrication step due to the
`demand for faster and more complex small-scale devices with mul-
`tilevel interconnects. During CMP, the wafer containing the device
`is mounted face-down onto a rotating carrier and pressed against a
`rotating polishing pad that is flooded with chemically reactive slurry
`containing abrasive nanoparticles, as shown in Fig. 1. The mechani-
`cal and chemical interactions between the wafer, pad, and slurry
`cause the surface of the wafer to wear to atomically smooth levels.
`Although CMP is widely used in industry, much of the physics
`behind CMP is not known because of the complex phenomena at
`the wafer–pad interface. These complexities include 共i兲 the slurry
`flowfield and film thickness distribution in the wafer–pad interface,
`共ii兲 the material wear effects caused by interactions between contact-
`ing wafer and pad asperities, 共iii兲 the effects of wafer and pad
`surface roughness on the slurry flowfield, 共iv兲 the material wear
`effects caused by the nanoparticles, and 共v兲 the effect of the
`nanoparticles on the rheology of the slurry. The lack of detailed
`knowledge of these effects has reduced CMP optimization into a
`mostly empirical process. Thus, CMP modeling and experimentation
`has become critical for understanding CMP and minimizing the
`amount of trial-and-error schemes that are currently necessary for
`CMP optimization.
`A number of experimental and theoretical studies have been
`conducted in order to analyze different aspects of the CMP process.
`A great deal of CMP research involves analysis of the material
`removal rate 共MRR兲. A generalized expression for the wafer surface
`material removal rate is given by Preston’s wear equation, as
`follows
`
`MRR =
`
`kPU
`H
`
`关1兴
`
`where k is the nondimensional Preston’s wear coefficient, P is the
`wafer downforce, V is the relative velocity of the wafer–pad inter-
`face, and H is the hardness of the wafer surface. Preston’s wear
`equation has commonly been used as an approximation for global
`MRR.
`A number of more sophisticated wafer surface wear models have
`been developed to account for various physical phenomena that take
`place during CMP. Nanz and Camilletti1 provided a critical review
`of CMP models up until 1995. Several studies have taken a contact
`mechanics approach toward CMP analysis, assuming that the wafer
`and pad surfaces are in direct sliding contact during the CMP pro-
`cess. Additional studies have assumed that the wafer and pad are
`
`z E-mail: higgs@andrew.cmu.edu
`
`Downloaded on 2016-02-25 to IP
`
`104.129.194.122
`
`
`
`ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see
`
`) unless CC License in place (see abstract).
`
`Raytheon2030-0001
`
`Sony Corp. v. Raytheon Co.
`IPR2015-01201
`
`

`
`
`
`K16K16
`
`Journal of The Electrochemical Society, 153 共6兲 K15-K22 共2006兲
`
`Figure 3. Diagram of 3D pad–surface model from from Nishioka et al.3
`Adapted with permission, © 1999 IEEE.
`
`Figure 1. Diagram of the CMP process.
`
`across the wafer using the expression for Couette/Poiseuelle flow,
`given as follows
`
`冊 −
`
`y h
`
`u共x,y兲 = U冉1 −
`
`h2
`dp
`2␮
`dx
`where U is the velocity of the pad surface, h is the film thickness,
`and p is the hydrodynamic pressure. From the results of their model,
`the authors found that certain conditions caused the slurry flow to
`have an unstable separation region. The results of their lubrication
`study were combined with mass-transport theory in order to predict
`the material removal rate distribution over the surface of the wafer.
`Nishioka et al.3 presented an analytical model for the slurry film
`thickness and wafer–pad coefficient of friction during CMP, ac-
`counting for pad surface roughness. The pad was modeled as a mov-
`ing 3D sinusoidal surface, while the wafer was modeled as a flat,
`stationary surface. The expression for the pad surface is given as
`follows
`
`关3兴
`
`冊
`
`y h
`
`冉1 −
`
`y h
`
`wafer surface was assumed to be fixed for simplification purposes.
`The slurry film thickness was estimated by assuming a contact stress
`distribution across the wafer and then solving for film thickness
`using the Greenwood and Williamson contact stress model.6 This
`resulted in a film thickness distribution that was smallest at the
`edges of the wafer and largest in the middle. From the film thickness
`distribution the authors used a finite-differencing algorithm on the
`1D Reynolds equation 共Eq. 2兲 to determine the predicted pressure
`distribution across the wafer. Their analysis showed that the slurry
`pressure distribution is subambient at certain locations along the
`length of the wafer, implying that the wafer is “sucked down” at
`those locations. Validation experiments were conducted using a
`commercial benchtop polisher that was fitted with preconditioned
`polishing pads. The pressure distribution was measured by outfitting
`the simulated wafer surface with a series of pressure taps whose data
`was acquired using an electronic pressure transducer. For these ex-
`periments, water was substituted for slurry as the interfacial fluid in
`order to prevent the possibility of the particles interfering with the
`pressure taps. The results of the validation experimentation also
`showed a region of subambient pressure and corresponded well with
`the predicted results.
`A study by Higgs III et al.7 expanded on the studies by Shan et
`al.4,5 by performing a 2D analysis of the entire wafer instead of a
`line of constant pad radius. This study, like the previous study, in-
`volved a combination of experimental pressure measurements and
`mathematical modeling. The mathematical model was created using
`the polar form of the Reynolds Equation, given as follows
`⳵p
`⳵p
`⳵r
`⳵␪
`r
`where r and ␪ are radial and tangential coordinates along the wafer–
`pad interface, respectively, and ␻ is the rotational speed of the pad.
`The film thickness, h, was found by assuming a given contact stress
`distribution and then finding the attack angle of the wafer by bal-
`ancing forces and moments about the pivot point. As a result of this
`study, the authors found yet again that a significant portion of the
`pressure distribution was subambient, as shown in Fig. 6. The vali-
`dation experiments were conducted using a tabletop polisher using
`
`冊 = 6␮共r␻兲 ⳵h
`
`⳵␪
`
`关5兴
`
`冉h3
`
`⳵ ⳵
`1 r
`
`␪
`
`冊 +
`
`冉rh3
`
`⳵ ⳵
`
`关4兴
`
`h共x,y兲 = h0 + RP cos
`
`cos
`
`2␲y
`2␲x
`␭y
`␭x
`where h0 is the mean line of the pad, RP is the peak roughness, and
`␭X and ␭Y are the wavelengths of the pad in the x and y directions,
`respectively. A diagram of the resultant sinusoidal surface is shown
`in Fig. 3. The parameters needed to define the sinusoidal surface
`共such as wavelength and roughness amplitude兲 were determined by
`performing roughness analysis on an actual pad. From this model,
`expressions for the mean hydrodynamic pressure and mean shear
`stress were derived as functions of the minimum film thickness.
`Their prediction of the pressure variation with film thickness is
`shown in Fig. 4. In order to validate their model, they compared the
`predicted coefficient of friction of their model to the measured co-
`efficient of friction from experimental CMP tests.
`Studies by Shan et al.4,5 have been used to analyze the slurry
`hydrodynamic pressure distribution across the wafer during CMP.
`Their analysis was conducted using a combination of mathematical
`modeling and validation experiments. The modeling aspect of their
`study involved the use of the one-dimensional Reynolds equation
`taken over a constant-pad-velocity line as shown in Fig. 5. The
`
`Figure 2. Diagram showing the difference between a convex and a concave
`wafer.
`
`Figure 4. Variation of mean pressure with minimum film thickness from
`Nishioka et al.3 Adapted with permission, © 1999 IEEE.
`
`Downloaded on 2016-02-25 to IP
`
`104.129.194.122
`
`
`
`ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see
`
`) unless CC License in place (see abstract).
`
`Raytheon2030-0002
`
`

`
`Journal of The Electrochemical Society, 153 共6兲 K15-K22 共2006兲
`
`
`
`K17K17
`
`be a Newtonian, particle-free fluid, which provided the following
`governing equation
`
`−12␮ⵜ · 关共hw − s兲3ⵜp兴 +
`
`
`
`ⵜ · 共Uជw + Uជp兲
`
`
`
`共hw − s兲
`2
`
`
`
`
`· 共Uជp − Uជw兲 + Vw − Vp = 0
`
`关7兴
`
`共ⵜhw + ⵜs兲
`2
`where ␮ is the dynamic viscosity of the fluid, hw and s are the wafer
`and pad surface topographies, respectively, p is the hydrodynamic
`pressure, and Uw and Up are the velocities of the wafer and pad,
`respectively. The governing equation was solved using an iterative
`finite difference scheme by being subjected to a balance of forces
`and moments around the pivot point. From this analysis, the authors
`were able to calculate the slurry pressure field, flowfield, and film-
`thickness distribution. The pressure field from their study is shown
`in Fig. 8, showing concentric isobars that are greater than ambient
`everywhere in the flowfield. Figure 9 shows vector plots of the
`interfacial flowfield at different vertical “slice” locations inside the
`wafer–pad gap. As Fig. 9 shows, the slurry flowfield appears to
`closely follow the motion of the pad near the pad 共z* = 0兲, then
`transitions into following the motion of the wafer as vertical location
`of interest increases. In the vertical location directly next to the
`wafer 共z* = 1兲, the slurry flow approximately follows the motion as
`the wafer.
`Jeng and Tsai15 presented a CMP model which combines hydro-
`dynamic lubrication theory with granular flow analysis in order to
`account for the motion of the slurry with abrasive nanoparticles. The
`hydrodynamic aspect of this model was based on the macroscopic
`Navier–Stokes equations, while the granular flow aspect of this
`model was based on microscopic molecular theory from a separate
`study.16 These two approaches were combined and simplified into a
`set of governing equations which described particle-fluid motion.
`Their resultant model predicted that the material removal rate in-
`creases proportionately with particle size. They later expanded upon
`their previous model by accounting for pad roughness effects using
`a combination of flow factors from separate studies.17-20 The studies
`by Jeng and Tsai focused a significant amount of attention on the
`effect of the abrasive particles in CMP, which is an important aspect
`of CMP that is often neglected in literature. However, their studies
`assumed that
`the slurry was completely composed of particles,
`which is an exaggerated assumption because slurry is composed
`primarily of fluid and contains only a trace amount 共3–5 wt %兲 of
`particles.21,22
`Chen and Fang23 presented a mathematical model which predicts
`the slurry film thickness and pressure distribution during CMP. The
`wafer was assumed to have a convex shape while the pad was as-
`sumed to be completely flat and horizontal. The resultant pressure
`distribution was found by deriving the polar form of the Reynolds
`equation and solving it by expanding the pressure distribution into
`
`Figure 5. Diagram of constant pad velocity line that was analyzed in the 1D
`CMP hydrodynamic studies.
`
`+
`
`water as a substitute for slurry in a method similar to that of Shan et
`al.4,5 The predicted results matched up well with experimental data.
`More works related to the modeling of subambient hydrodynamic
`slurry pressure can be found in the literature8-12 but are omitted from
`this paper for the sake of brevity.
`Cho et al.13 presented a 2D mathematical model which predicts
`the slurry film thickness and pressure distribution during CMP. For
`this analysis the authors used the polar form of the Reynolds equa-
`tion. They assumed that both the wafer and pad surfaces were com-
`pletely flat, although the wafer was allowed to tilt from its center
`pivot point in order to balance forces and moments. The authors also
`assumed that the slurry was a particle-free, incompressible, Newton-
`ian fluid for their analysis. By coupling the tilt of the wafer 共film
`thickness distribution兲 with the balance of forces and moments on
`the wafer 共pressure distribution兲, the authors were able to solve for
`both using the numerical Newton–Raphson method.
`Thakurta et al.14 developed a model for slurry film thickness and
`velocity distribution during CMP and compared the predicted results
`with the results of experiment. Their model, outlined in Fig. 7, ac-
`counted for the porosity and deflection of the pad surface. The the-
`oretical model was created by first assuming a parabolic shape for
`the convex wafer as follows
`
`h共x,y兲 = h0 + SX冉 x
`
`冊 + SY冉 y
`
`冊 + ␦0冉 x2 + y2
`
`冊
`
`a2
`a
`a
`where h0 is the centerline height of the wafer, ␦0 is the wafer dome
`height, a is the radius of the wafer, and SX and SY are the horizontal
`and vertical slopes of the wafer due to its equilibrium angle on the
`gimbal. Additionally, the pad surface topography, given by s共x,y兲,
`accounted for elastic deformation of the pad, which was specified to
`be directly proportional to the slurry hydrodynamic pressure. The
`pad porosity was also accounted for by assuming that a certain
`amount of slurry seeps into the pad depending on the hydrodynamic
`pressure distribution. Lubrication approximations were then used to
`simplify the polar Navier–Stokes equations, assuming the slurry to
`
`关6兴
`
`Figure 6. Predicted 2D fluid pressure from Higgs III et al.7: 共a兲 tangential and 共b兲 radial.
`
`Downloaded on 2016-02-25 to IP
`
`104.129.194.122
`
`
`
`ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see
`
`) unless CC License in place (see abstract).
`
`Raytheon2030-0003
`
`

`
`
`
`K18K18
`
`Journal of The Electrochemical Society, 153 共6兲 K15-K22 共2006兲
`
`Figure 7. Diagram of modeling domain used in Thakurta et al.14 Reproduced
`by permission of The Electrochemical Society, Inc.
`
`Stoum–Liouville eigenfunctions. As a result of their work they
`found that the pressure distribution takes on a half-parabolic shape
`across the radial direction from the center of the wafer. All predicted
`pressures from this study were greater than atmospheric.
`
`Slurry shear rate modeling.— It is possible that the shear rate of
`the slurry is of great importance to CMP hydrodynamics. Runnels
`and Eyman24 have postulated that the wafer surface wear rate is
`directly proportional to the shear rate of the slurry in the hydrody-
`namic lubrication regime according to the following equation
`关8兴
`MRR = K␴␶
`where MRR is the material removal rate, K is the Preston coefficient,
`␴ is the normal stress, and ␶ is the shear stress of the slurry fluid.
`Equation 8 is proposed only for hydrodynamic lubrication. If the
`minimum film thickness is on the order of or less than the average
`surface roughness, then the effect of solid contact must be taken into
`account.14 Several solid contact CMP models are currently available
`in literature, if solid–solid contact is to be assumed. Details of the
`solid contact models are outside the scope of this paper.
`A study by Sohn et al.25 involved the derivation of expressions
`for the shear rate for slurry flow at the wafer–pad interface. Assum-
`ing that the slurry behaved as a particle-free, Newtonian fluid, the
`authors used a simplified version of the incompressible Navier–
`Stokes equations to model the flowfield. For this analysis it was
`
`Figure 8. Pressure distribution under rotating wafer from Thakurta et al.14
`Reproduced by permission of The Electrochemical Society, Inc.
`
`Figure 9. Slurry flowfield at different vertical locations inside the wafer–pad
`gap from Thakurta et al.14 Reproduced by permission of The Electrochemical
`Society, Inc.
`
`assumed that both the wafer and pad surfaces were rotating, no-slip
`walls. Additionally, the wafer and the pad were assumed to be per-
`fectly parallel to each other, which resulted in a constant slurry
`hydrodynamic pressure. The authors simplified the Navier–Stokes
`equations by assuming that the Reynolds number and aspect ratio
`are both negligibly small. From this analysis the authors were able
`to derive closed-form expressions for the slurry velocity field and
`shear rate. They found that when the wafer rotational speed is
`greater than the speed of the pad, the shear rate is greatest at the
`edges of the wafer. However, when the wafer rotational speed is the
`same as that of the pad, the shear rate is uniform throughout the
`wafer.
`
`discrep-
`Discussion of hydrodynamic modeling studies.— The
`ancy between the predicted results of each of these models appears
`to be rooted in the different assumptions of the wafer and pad sur-
`face geometries. The studies by Sundararajan et al.,2 Thakurta et
`al.,14 Jeng and Tsai,15,17 and Chen and Fang23 assumed that the
`wafer surface was slightly convex, which resulted in a pressure dis-
`tribution that was greater than ambient everywhere in the domain. In
`contrast, the studies by Shan et al.4,5 and Higgs III et al.7 incorpo-
`rated contact stress models into their analysis and thus ended with a
`film thickness that was smallest at the edges of the wafer and largest
`in the middle. As a result, the former group of authors predicted a
`pressure distribution that was greater than ambient everywhere in
`the domain, while the latter group predicted a region of subambient
`pressure. In order to determine the correct pressure distribution, one
`must analyze the film thickness distribution and account for wafer
`
`Downloaded on 2016-02-25 to IP
`
`104.129.194.122
`
`
`
`ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see
`
`) unless CC License in place (see abstract).
`
`Raytheon2030-0004
`
`

`
`Journal of The Electrochemical Society, 153 共6兲 K15-K22 共2006兲
`
`
`
`K19K19
`
`Figure 10. Slurry flow domain and diagram of a real wafer–pad interface from Muldowney28 using Fluent 6.1, a commercial CFD solver. Reprinted with
`permission from Materials Research Society.
`
`bending and pad deflection. Numerous studies have been conducted
`which analyze the deflection of the wafer and pad during CMP,
`although a detailed discussion of these studies is omitted from this
`paper for the sake of brevity.
`Additionally, we must take note of some of the simplifications
`that are used in these models. The assumption of the slurry as a
`Newtonian fluid is the most widely used simplification in each of
`these models, with the exception of the studies by Jeng and Tsai,15,17
`who assumed that the slurry was composed completely of small
`“granular” particles.
`
`Numerical Studies in CMP Hydrodynamics
`A few studies have been conducted which use computational
`fluid dynamics 共CFD兲 to analyze the slurry flowfield in CMP. CFD
`solvers are advantageous for this analysis due to their ability to input
`complex flow domains and solve transport equations for multiphase
`flow and chemical reactions.
`One of the first CMP numerical studies was conducted by Run-
`nels and Eyman,24 who created a numerical model to predict the
`slurry film thickness and hydrodynamic pressure. Assuming that hy-
`drodynamic lubrication takes place between the wafer and pad sur-
`faces, they imposed a sample slurry flowfield domain into a numeri-
`cal code and solved it using a Galerkin finite element scheme. Both
`the wafer and pad surfaces were modeled as being rigid, smooth,
`no-slip walls. The pad was assumed to be flat while the wafer was
`designed to be convex with a specified radius of curvature. The
`wafer and pad walls were bounded by an additional wall which
`joined the two. This bounding wall was given a stress-free boundary
`condition in order to allow fluid to enter and exit the domain freely.
`The slurry film thickness was found by balancing the hydrodynamic
`forces with the applied load and pivoting the wafer such that the
`moment around the pivot point was zero. From the results of these
`simulations, the authors were able to find the amount of load that
`can be supported by the wafer as well as the minimum film thick-
`ness of slurry between the wafer and the pad.
`Fu and Chou26 used CFX-3D, a commercial numerical solver, to
`solve for the slurry flowfield in a CMP domain between the wafer
`and the pad. For their simulation, both the wafer and pad were
`modeled as being perfectly rigid, flat, and smooth, while the slurry
`was modeled as being a Newtonian, incompressible, particle-free
`fluid. The wafer–pad gap was fixed at a given input value, either 20
`or 40 ␮m. Both the wafer and pad walls were modeled as having
`no-slip boundary conditions, while the remaining surfaces were
`modeled as stress-free boundaries in order to allow the slurry to
`freely enter and exit the computational domain. The resultant slurry
`shear stress distribution from the simulation was used to estimate the
`material removal rate from the wafer surface.
`
`Yao et al.27 used the software package Fidap, a commercial nu-
`merical solver, to model the slurry flow pattern between two moving
`surfaces at different locations at the wafer/pad interface. They chose
`not to model the entire wafer/pad domain but rather modeled various
`geometries along the tangent of the wafer in order to conserve com-
`putational resources. Each of the geometries was square and fea-
`tured different scales of roughness. The boundary conditions for
`each of the geometries were dependent on the rotational movements
`of the pad and wafer. The slurry flow was modeled using the incom-
`pressible Navier–Stokes equations, assuming that the slurry exhib-
`ited nonNewtonian behavior due to the abrasive nanoparticles. They
`assumed that the material removal rate was directly proportional to
`the slurry shear stress as postulated in Runnels and Eyman, and then
`used Fidap’s time interval updating capability to change the slurry
`film thickness over time based on the predicted wafer surface wear.
`Using this method they were able to determine the amount of ma-
`terial removal that occurs to the surface roughness after a given
`amount of polishing time. They were also able to derive an empiri-
`cal model for the instantaneous polish rate with respect to time.
`Muldowney28 used the commercial numerical solver Fluent 6.1
`to model the slurry velocity field, thermal field, and chemical reac-
`tions between the pad and wafer during CMP. The CFD domain in
`this study is shown in Fig. 10. For this simulation, the wafer was
`modeled as being flat and smooth, while the pad was modeled as
`having a rough topography in the form of a series of concentric
`circular grooves that are separated by circular concentric “asperi-
`ties.” The gap between the “asperities” and the wafer surface was
`modeled as being porous in order to account for the porous flow of
`slurry through the asperities. The pad and the wafer surfaces were
`modeled as no-slip/no-penetration boundaries, while the slurry was
`assumed to be a Newtonian fluid. From this analysis, the author was
`able to show the resultant velocity profile of the slurry at the wafer–
`pad interface. Figure 11 shows the predicted velocity profile, which
`appears to have a stagnation/backflow region that is comparable to
`the midgap 共z* = 0.6兲 velocity field predicted by Thakurta et al.14
`共Fig. 9兲.
`Rogers et al.29 used a combination of numerical modeling and
`experimental testing to analyze the flow of slurry during CMP. For
`their numerical study they used Fluent, a commercial fluid flow
`solver, to analyze the flow of both the slurry and the surrounding air
`outside the wafer–pad interface. Their 2D flow domain consisted of
`the linearly moving pad surface, the surrounding air/slurry volume,
`and the fixed wafer surface, which was represented as a rigid punch.
`The motion of the air and slurry were analyzed using Fluent’s vol-
`ume of fluid 共VOF兲 solver, which modeled the air and slurry as two
`immiscible fluids. Both the wafer and pad walls were modeled as
`no-slip boundaries, while the side walls were modeled as cyclic
`
`Downloaded on 2016-02-25 to IP
`
`104.129.194.122
`
`
`
`ecsdl.org/site/terms_use address. Redistribution subject to ECS terms of use (see
`
`) unless CC License in place (see abstract).
`
`Raytheon2030-0005
`
`

`
`
`
`K20K20
`
`Journal of The Electrochemical Society, 153 共6兲 K15-K22 共2006兲
`
`Figure 11. Resultant slurry velocity field from Muldowney et al.28 Reprinted
`with permission from Materials Research Society.
`
`boundaries. From their numerical study they found that the slurry
`hydrodynamic pressure was greater than ambient everywhere inside
`the wafer–pad gap and decreased linearly along the length of the
`wafer. These results are in contrast to the analytical results of Levert
`et al.30 共discussed in the Experimental section of this paper兲, who
`found a subambient pressure region in the wafer–pad gap. They
`attributed this discrepancy to the fact that Levert et al. operated in
`the asperity contact regime, while they worked in the hydrodynamic
`lubrication regime.
`
`numerical
`Discussion of numerical modeling studies.— These
`studies have shown that it is possible to analyze surface wear, pres-
`sure distribution, and chemical reactions in a CMP domain using a
`numerical solver. Each of these studies took into account various
`aspects of the CMP process, such as chemical reactions, the erosion
`of wafer roughness, the rotation of both the wafer and the pad, and
`the interaction of atmospheric air with the flow of the slurry. How-
`ever, the simulations described in each study necessitated the use of
`various assumptions in order to simplify the problem and minimize
`computational expense. Examples include the assumption of per-
`fectly flat, perfectly parallel wafer and pad surfaces in the study by
`Yao et al.27 and the 2D assumption in the study by Rogers et al.29
`Because CMP is a complex process involving several physical
`phenomena, it is desirable to have a CMP numerical model which
`accounts for as much of the CMP physical phenomena as possible. It
`is expected that CMP numerical simulations will become increas-
`ingly sophisticated and realistic as computing resources continue to
`improve.
`
`Experimental Studies in CMP Hydrodynamics
`Hydrodynamic experiments in CMP have primarily served to
`examine parameters such as the slurry pressure field, the slurry film
`thickness distribution, and the wafer coefficient of friction. The im-
`plications of the slurry pressure field and film thickness distribution
`are described in the previous section. The slurry coefficient of fric-
`tion provides insight into the amount of abrasive wear that the wafer
`experiences.
`Levert et al.31 conducted a series of experiments to determine the
`slurry pressure distribution, wafer–pad coefficient of friction, and
`the wafer surface wear during CMP. They performed two sets of
`tests: 共i兲 hydrodynamic CMP tests, which were conducted with a
`light load so that the wafer would hydroplane on top of the pad
`surface, and 共ii兲 commercial CMP tests, which were conducted with
`a heavier load such that the pad and wafer asperities touched. These
`experiments were conducted using a bench-top polisher that was
`outfitted with an overhead wafer carrier. The wafer carrier itself was
`attached to an array of capacitance probes which helped to measure
`the wafer surface wear. From these experiments it was found that the
`hydrodynamic regime causes the wafer surface to wear away at a
`rate which is three orders of magnitude lower than commercial CMP
`rates. The authors thus concluded that CMP must have contacting
`wafer–pad asperities in order to polish the wafer surface adequately.
`
`Figure 12. Slurry pressure test fixture from Zhou et al.34 Reprinted with
`permission from Elsevier, © 2002.
`
`Bullen et al.32 designed and performed a series of experiments in
`order to analyze the slurry pressure distribution during CMP. Their
`experimental facility consisted of a tabletop polisher which was
`pressed upon by a drill press which served as the wafer carrier. The
`wafer itself was outfitted with a series of pressure taps at various
`radii that were connected to a pressure transducer, which rotated
`with the wafer. Before each experiment, the pad surface was condi-
`tioned using a diamond-grit polisher and the wafer surface was pol-
`ished to a convex shape in order to prevent the possibility of a
`vacuum at the wafer–pad interface. A series of tests were conducted
`with both a rotating and nonrotating wafer surface. From the results
`of this study it was found that the pressure distribution in the non-
`rotating wafer had two peaks and two valleys but did not have a
`significantly large subambient pressure region, as predicted in Shan
`et al.4,5 It was in fact predicted that the applied load was supported
`by the center of the wafer, where the pressure was predicted to be
`the highest. It is possible that the discrepancy between the results of
`this study and Shan et al.’s study can be attributed to the convex
`shape of the wafer in this study, which causes a positive hydrody-
`namic pressure from lubrication theory. The authors in this study
`also investigated the dynamic pressure distribution for a rotating
`wafer. They found that although the dynamic pressure distribution
`still had a semblance of peaks and valleys like in the static case, the
`ampli

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket