Review Fundamental understanding and modeling of reactive sputtering processes S. Berg*, T. Nyberg The Angstrom Laboratory, Uppsala University, Box 534, 751 21 Uppsala, Sweden Accepted 21 October 2004 Available online 7 January 2005 Abstract Reactive sputtering is a commonly used process to fabricate compound thin film coatings on a wide variety of different substrates. The industrial applications request high rate deposition processes. To meet this demand, it is necessary to have very good process control of such processes. The deposition rate is extremely sensitive to the supply of the reactive gas. A too low supply of the reactive gas will cause high rate metallic sputtering, but may give rise to an understoichiometric composition of the deposited film. A too high supply of the reactive gas will allow for stoichiometric composition of the deposited film, but will cause poisoning of the target surface, which may reduce the deposition rate significantly. This behaviour points out that there may exist optimum processing conditions where both high rate and stoichiometric film composition may be obtained. The purpose of this article is to explain how different parameters affect the reactive sputtering process. A simple model for the reactive sputtering process is described. Based on this model, it is possible to predict the processing behaviour for many different ways of carrying out this process. It is also possible to use the results of the modeling work to scale processes from laboratory size to large industrial processes. The focus will be to obtain as simple a model that will still quite correctly describe most experimental findings. Despite some quite crude approximations, we believe that the model presented satisfies this criterion. D 2004 Elsevier B.V. All rights reserved. Keywords: Reactive sputtering; Thin film coatings; Reactive gas Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216 2. Definition of conditions for the reactive sputtering model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217 2.1. Flux of reactive gas in processing chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218 2.2. Conditions at the target surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218 2.3. Conditions at collecting area (substrate) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218 2.3.1. Flux of material approach. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218 2.3.2. Sputtered elemental target atoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219 2.3.3. Material conservation approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219 2.4. Deposition rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219 2.5. Modeling as a function of reactive gas partial pressure P . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219 2.6. Kinetics of the reactive gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220 2.7. Modeling as a function of the total reactive gas supply rate Qtot . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220 0040-6090/$ - see front matter D 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.tsf.2004.10.051 * Corresponding author. E-mail addresses: soren.berg@angstrom.uu.se (S. Berg)8 tomas.nyberg@angstrom.uu.se (T. Nyberg). Thin Solid Films 476 (2005) 215–230 www.elsevier.com/locate/tsf 3. Parameters influencing the hysteresis effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221 3.1. Influence of target material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 3.2. Influence of reactive gas. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 3.3. Influence of system pumping speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222 3.4. Target to substrate distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223 3.5. Target ion current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223 3.6. Target area: hysteresis-free operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224 4. Dissociation of sputtered compound molecules: effect on modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225 5. Reactive co-sputtering. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225 6. Reactive sputtering from an alloy target . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226 7. Processing with several reactive gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226 8. Transient behaviour: pulsed DC reactive sputtering. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227 9. Ion implantation: reactive gas. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228 10. Effect of secondary electrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 11. Non-uniform target current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 12. Modelling conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230 References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230 1. Introduction Reactive sputtering for thin film coatings is used in numerous industrial applications. By simply adding a gas that reacts with the sputtered material, it is possible to form a wide variety of useful compound thin film coatings. At first sight, this may seem very simple. However, the reaction mechanisms between the sputtered material and the reactive gas may cause some processing stability problems. Combining a high deposition rate and true compound stoichiometry of the deposited film turn out to appear as contradicting desires. In many cases, the main reason for this complication is that as well as forming a compound of the deposited film, compound formation will also take place at the surface of the sputtering target (target poisoning). Normally, the sputtering yield of the compound material is substantial lower than the sputtering yield of the elemental target material. This causes the deposition rate to decrease as the supply of the reactive gas increases. The relationship between the film composition and supply of reactive gas is very non-linear. This is also the case for the deposition rate vs. the supply of the reactive gas. Therefore, reactive sputtering processes controlled by the supply of the reactive gas exhibit quite complex processing behaviour [1,2]. Fig. 1 shows a typical experimental processing curve for the sputter erosion rate vs. the supply of the reactive gas for a reactive sputtering process (carried out with constant target current during processing). The characteristic feature of this curve is that it exhibits a hysteresis effect. The deposition rate does not decrease and increase at the same value with the supply of the reactive gas. The separation width between the decrease and increase denotes the width of the hysteresis region. A corresponding hysteresis effect is observed for the relationship between the partial pressure and the supply rate of reactive gas. This is shown in Fig. 2. This figure clearly illustrates a difference in increasing and decreasing the supply of the reactive gas. During the increase sequence of the reactive gas supply, the partial pressure of the reactive gas remains at a very low level until reaching the upper limiting value of the hysteresis width. During a decrease of the supply of the reactive gas, the partial pressure remains significantly higher in the hysteresis region than during the increase of the gas. This points out that more gas is consumed for compound formation during the increase sequence. This may be understood by comparing the sputter erosion rate curve for the corresponding sequences. A higher sputter erosion rate needs more reactive gas to form compound coatings. The hysteresis effect is one of the key problems in experimental reactive sputtering systems. In the following treatment, we will therefore demonstrate how the hysteresis is affected by different processing parameters. The goal is to better understand the influence of processing parameters on the overall behaviour of this process. Fig. 1. Typical experimental curve for a reactive sputtering process. The optical emission (OES) from sputtered metal atoms represents the sputter erosion rate. Qtot is expressed in standard cubic centimeters per minute (sccm). S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230216 In order to be able to predict the outcome of reactive sputtering processes, there is a need for a reliable model. In the following paragraphs, we will outline a simple model for reactive sputtering processes. The core idea of this model has been published earlier [3]. It has become the commonly accepted treatment and is frequently referred to as bBerg’s modelQ. However, the results should be considered as first order approximations. Despite the simplicity, however, the results have proved to fit surprisingly well to most experimentally found results. There exist process conditions, however, where a more sophisticated treatment must be considered. Even in these cases, the treatment can be based on a similar balance philosophy as suggested in the basic simple bBerg’s modelQ [3–8]. 2. Definition of conditions for the reactive sputtering model In order to understand the influence of different processing parameters for the reactive sputtering process, we will try to keep the number of parameters as low as possible. To satisfy this goal, we have as a first order approximation neglected some effects that may influence the overall processing behaviour. We will comment on this later in the article. A schematic of a simple sputtering equipment is shown in Fig. 3. The mathematical model describing this system is shown in Fig. 4a–b. We assume a target (area At) in front of a collecting surface (area Ac) in a vacuum chamber. There is a pump connected to the vacuum chamber having a constant pumping speed S for the gases involved. A reactive gas supply is connected to the vacuum chamber. We assume that the following conditions exist in the chamber: The total supply rate of the reactive gas is denoted Qtot. There is a uniform partial pressure P of the reactive gas in the chamber. Since there is reactive gas present, the reactions between elemental target atoms and the reactive gas will cause a fraction ht of the target to consist of compound molecules. This compound formation is of course uniformly distributed over the whole target surface. To clearly illustrate that a fraction ht of the target has a different composition than the remaining (1Àht) fraction, we treat the ht fraction as a separate continuous area. The remaining part (1Àht) of the target surface consists of elemental non-reacted target atoms. The same situation exists at the receiving area Ac where all sputtered material is assumed to be uniformly collected. The compound fraction at this surface area (deposited film) is denoted hc. Notice that, with this definition, hc is also a Fig. 2. The partial pressure, P, of the reactive gas corresponding to the curve in Fig. 1. Fig. 3. Schematic of a simple reactive sputtering system. Fig. 4. (a) Theoretical equivalent for the system shown in Fig. 3. The notations are referred to and explained in the text. (b) Illustration of flux of sputtered material to the substrate area Ac. The notations are referred to and explained in the text. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230 217 measure of the composition of the deposited film. We also assume that the ratio between the electron and ion currents does not change during processing. The ion current density J (Amps/unit area) is assumed to be uniformly distributed over the target surface At. The sputtering contribution by ionized reactive gas is neglected, which is valid when the partial pressure of the reactive gas is significantly lower than the partial pressure of the argon gas. We also assume that sputtering of a compound molecule from the target surface results in exposing the original non-reacted surface at that target position. This corresponds to forming only one monolayer of the compound by chemisorption. 2.1. Flux of reactive gas in processing chamber A uniform partial pressure P of the reactive gas will cause a uniform bombardment of neutral reactive molecules F (molecules/unit area and time) to all surfaces in the processing chamber. From gas kinetics [9], the relationship between F and P is F ¼ P ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2kTpm p ; ð1Þ where k is the Boltzmann constant, T is the temperature and m is the mass of the gas molecule. In the following, we will derive steady-state values for some representative parameters during different processing conditions. The definition for steady state is that there be no time variations in the processing parameters. The consequences will be discussed in the following sections. 2.2. Conditions at the target surface For simplicity, we assume the compound molecules at the target and substrate to consist of one reactive gas atom and one metal atom. Proper correction factors should be added for other stoichiometries. In the following description, we refer to Fig. 4a. During processing, the target will be sputter eroded. Elemental metal atoms are sputtered from the surface fraction (1Àht) and, for simplicity, the sputtered material from the surface fraction ht is assumed to be sputtered as molecules, irrespective of whether the material is ejected in atomic or molecular form. The only way sputtered compound molecules may be replaced is by reactions between neutral reactive gas molecules and the elemental non-reacted target material. We assume that no reactions between underlying elemental target atoms and the reactive gas can take place on the compound fraction ht of the target area At. A steady-state equation for the target may therefore be defined as J q Ycht ¼ a2F 1 À htð Þ; ð2Þ where Yc is the sputtering yield of compound molecules from the compound-covered fraction ht of the target, a is the probability (sticking coefficient) for a colliding neutral reactive gas molecule to react with an unreacted elemental target atom at the (1Àht) fraction of the target, q is the elementary electronic charge and F is multiplied by a factor of 2 because we assume two atoms per gas molecule. This equation states that at steady state the sputter erosion rate of compound molecules from the target must be identical to the compound formation rate by reactions between neutral reactive gas molecules and elemental target atoms. To further illustrate this, we have inserted a small compound area in front of the arrow corresponding to the incoming flux of reactive gas, Qt, to the target surface. This symbolizes the rate of compound formation at the target. We have also bopened upQ a small hole in the ht fraction area of the target. This is to illustrate the sputter removal rate of the compound material from the target. At steady state, the compound area to the right must breplaceQ the hole to the left. From Eqs. (1) and (2), it is possible to solve for ht as a function of partial pressure of the reactive gas. 2.3. Conditions at collecting area (substrate) We assume that all sputtered material from the target will be uniformly deposited at the collecting area Ac. At steady state, the composition of added film material must be identical to the existing film composition. There are several ways of describing this situation. 2.3.1. Flux of material approach By the following arguments, hc may be evaluated. The descriptions refer to Fig. 4b. The total number of outsputtered compound molecules per unit time from the target will be denoted Fc. Fc ¼ J q YchtAt; ð3aÞ where Yc represents the sputtering yield of the compound material. We assume that this compound material will be uniformly distributed onto the collecting surface Ac. Notice, however, that ( Fchc) will be deposited at the fraction hc of Ac that already has been defined to consist of compound material. Adding compound material to the already existing compound fraction of the collecting area will not change the local composition. Consequently, the addition of ( Fchc) to the hc fraction at the collecting area Ac will not influence the surface fraction hc. This contribution will therefore be neglected in this treatment. The remaining fraction Fc(1Àhc) of sputtered compound material will be deposited onto the fraction (1Àhc) of the collecting area Ac containing elemental non-reacted target atoms. This will convert some part of the (1Àhc)Ac area to become part of the hc fraction of compound material at the surface Ac. This fraction of the deposited compound material will thus contribute to an increase in the value of hc. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230218 2.3.2. Sputtered elemental target atoms A corresponding argument may be applied to the sputtered non-reacted elemental target atoms, for which the total number per unit time in Fig. 4b is denoted by Fm ¼ J q Ym 1 À htð ÞAt ð3bÞ The fraction of flux Fm(1Àhc) deposited onto the fraction (1Àhc) of the collecting area Ac will not influence the value of hc. Adding elemental non-reacted target atoms to an area already defined to consist of non-reacted atoms does not change hc. However, the flux Fmhc deposited onto the fraction hc of compound material at the collecting area Ac will contribute to a decrease in hc. Compound formation by reaction between reactive gas molecules and elemental non-reacted atoms at the collecting area will consume Qc reactive gas molecules. As for the target, we assume that no reactions between elemental target atoms and the reactive gas can take place on the compound fraction hc of the collecting area. Therefore, Qc ¼ aF 1 À hcð ÞAc: ð3cÞ For simplicity, we assume the same value for a as for the target. Since the reactive gas molecules consist of two atoms, Qc will contribute with two compound molecules adding to the value of hc. At steady state, the contributions supporting an increase in hc must be identical to the contributions that support a decrease in hc. This leads to the following balance equation for the collecting area Ac, 2Qc þ Fc 1 À hcÞ ¼ hcFm:ð ð3Þ From Eqs. (1)–(3), it is possible to solve for hc as a function of the partial pressure of the reactive gas. 2.3.3. Material conservation approach Some arguments have been raised against the above approach; namely, that the flux of materials approach described in Sections 2.3.1 and 2.3.2) above neglects some of the outsputtered material ( Fchc) and Fm(1Àhc). The following alternative treatment, however, serves to support the validity of the balance treatment. The ratio hc/(1Àhc) between the compound and elemental target atom fractions in the deposited film must be identical to the corresponding ratio ( Fc+2Qc)/Fm of the added film material. This relation defines a balance equation for the collecting area Ac, hc 1 À hc ¼ Fc þ 2Qc Fm ð3VÞ By solving Eqs. (3) and (3V) for hc, we find that both these equations result in identical expressions for hc. The bflux of materialQ and bmaterial conservationQ approaches give rise to identical mathematical conditions for the steady-state conditions at the collecting area Ac. 2.4. Deposition rate Using the above notation, it is easy to define an expression for the total sputter erosion rate R from the target (atoms+molecules per unit area and time), R ¼ J q Ym 1 À htÞ þ Ychtð Š:½ ð4Þ Proper constants may be added to convert R into desired units (e.g., angstrom/min, etc.). It must be understood, however, that more material is deposited at the collecting area Ac than sputter eroded from the target area At. This is because Qc of the reactive gas is added to the sputtered flux of material. The difference in mass between sputter eroded material and the mass of the deposited material will strongly depend on the content of reactive gas atoms in the compound molecule per sputtered elemental metal atom. An expression for the deposition rate D may be obtained in the following way. The total sputtering rate Rm of metal atoms (elemental+those included in sputtered molecules) is Rm ¼ J q Ym 1 À htð Þ þ Ycht½ ŠAt: ð5VÞ Knowing the compound fraction hc at Ac makes it quite easy to derive an expression for the deposition rate D. D ¼ c1Rm 1 À hcð Þ þ c2Rmhc; ð5Þ where c1 and c2 are constants accounting for unit conversions. The first term in Eq. (5) represents the contribution by elemental metal atoms, while the second term represents the contribution by compound material. We have chosen to express normalized D in units of mass/unit time. 2.5. Modeling as a function of reactive gas partial pressure P From Eqs. (1)–(5), it is possible to calculate R, D, ht and hc as a function of partial pressure P. Typical curves are shown in Figs. 5 and 6). Numerical values used for the calculations are given in Appendix Notice that all curves Fig. 5. Calculated target erosion rate R and substrate deposition rate D vs. partial pressure, P, of the reactive gas. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230 219 are bwell behavedQ and give single valued results at all partial pressures. Moreover, the deposition rate curve may have a maximum value before reaching complete target poisoning. These curves correspond to the reactive deposition of AlN. Unfortunately, it is not easy to experimentally control the partial pressure of the reactive gas during processing. Therefore, simple processing curves like Figs. 5 and 6 will not be obtained without quite sophisticated process control systems. The simplest and most common control function is the supply of reactive gas to the processing chamber. This calls for modeling equations based on the supply of the reactive gas rather than based on the partial pressure. 2.6. Kinetics of the reactive gas The relation between the reactive gas supply and the partial pressure may quite easily be solved from the model outlined above where the compound formation (reactions between elemental target atoms and reactive gas molecules) consumes a fraction of the incoming gas. A schematic diagram of the reactive gas pathways was shown in Fig. 4a. The consumption (number of molecules per unit time) at the target Qt is obtained from Eq. (2), Qt ¼ aF 1 À htð ÞAt; ð6aÞ while the consumption Qc at the collecting area Ac is obtained from Eq. (3c). Qc ¼ aF 1 À hcð ÞAc: ð6bÞ The remaining part Qp of the reactive gas will escape from the processing chamber through the pumping system. Assuming a pumping speed of S for the system pump gives Qp ¼ SP: ð6cÞ The total supply Qtot is the sum of all sources for reactive gas consumption, Qtot ¼ Qt þ Qc þ Qp: ð6Þ 2.7. Modeling as a function of the total reactive gas supply rate Qtot Adding Eq. (6) to the previous equations makes it possible to find the relation between Qtot and P, and consequently calculate R, D, ht, hc, Qt, Qc, Qp and P as a function of the supply rate of the reactive gas Qtot. Results are shown in Figs. 7–9. It is obvious that these curves show a more complex relation to Qtot than the curves in Figs. 5 and 6 did to the partial pressure P. The S-shaped behaviour defines a region where one value for Qtot may be satisfied by three different values of R, D, P, ht and hc. This region corresponds to the hysteresis width shown in Figs. 1 and 2 in the introduction. Four different points, P1–P4, are marked in Figs. 7–9. In the following, we discuss important phenomena that take place at these positions. Before elaborating on this, however, we clarify two different modes of operation. It is important to understand the difference between processes controlled by partial pressure of the reactive gas and by the supply of the reactive gas. This difference is illustrated in Fig. 10a–b. Selecting values for the partial pressure in Fig. 10a corresponding to positions 1, 2, . . ., 9 on the curve result in well-defined single-valued supply levels Qtot. Thus, by selecting continuously increasing Fig. 6. Calculated compound fractions ht and hc at the target and substrate, respectively, as a function of the partial pressure, P, of the reactive gas. Fig. 7. Calculated target erosion rate R and substrate deposition rate D vs. total supply rate of reactive gas, Qtot, for conditions identical to those used in generating Figs. 5 and 6. Fig. 8. Calculated partial pressure, P, vs. supply rate of the reactive gas, Qtot. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230220 values of the partial pressure of the reactive gas, the Sshaped curve shown in Fig. 10a will be generated. This is the signature of a partial pressure process control. As previously shown in Fig. 5 in Section 2.5, any selected partial pressure value also corresponds to a welldefined single value of the sputter erosion rate R. Thus, it is possible to obtain paired values (R, Qtot) corresponding to specifically selected partial pressures. These (R, Qtot) pair values will define a R vs. Qtot curve with the shape shown in Fig. 10b. It must be understood that experimentally this curve can only be generated by stepping values of the partial pressure not by stepping values of the reactive gas supply. In practice, this curve can only be obtained for processes having some feedback control system enabling, in a controlled way, variations in the reactive gas supply corresponding to the desired values of the partial pressure. A different behaviour will result if the supply rate of the reactive gas Qtot is used as the process control parameter. Continuously increasing Qtot can not generate the entire Sshaped curve in Fig. 10b. When Qtot reaches the value corresponding to position 2 on the curve, a dramatic effect occurs. A small additional increase in Qtot forces the rate to avalanche from position 2 down to position 7 on the curve. An even further increase of Qtot generates R values 8, 9, etc., following the curve. Starting at high gas flow values (corresponding to position 9 on the curve) and then decreasing Qtot generates values of the erosion rate R following the curve in reverse order to position 4. However, at this point, a further decrease in Qtot will result in R avalanching from position 4 to a point somewhat to the left of position 1. A further decrease in Qtot will cause the rate R to follow the curve. The fact that the sharp decrease and increase of R do not occur at the same value for Qtot illustrates the hysteresis effect with the hysteresis width defined by the separation between the two avalanche positions. This is the way that the R responds experimentally to having the supply Qtot as the input control parameter. It must also be pointed out that, in this mode of operation, it is not possible to reach any of the processing points on the segment between the positions 2 and 4 on the curve. Instead of following the S-shaped loop on this segment, the process will by-pass the loop and avalanche along the dotted lines. Thus, the positions labeled P1–P4 in Figs. 7–9 denote the processing region (shown shadowed) in which hysteresis occurs. 3. Parameters influencing the hysteresis effect There exist many bpopularQ explanations regarding the origin of the hysteresis effect in reactive sputtering processes. A too simplified explanation, however, may cause the experimentalists to draw wrong conclusions concerning how to optimize the process. We therefore want to clarify parametric interactions in somewhat more detail. It should be pointed out that, although hysteresis is generally observed in reactive sputter processing, there also exists conditions under which hysteresis does not occur. It is possible to use the mathematical model outlined above to analyze specific processes and predict if hysteresis will appear or not. In order to investigate the influence of processing parameters on the reactive sputtering process, we will introduce some illustrative curves. Plotting the curves for Qtot, Qt, Qc and Qp as a function of P may clarify the gettering situation. This is shown in Fig. 11a where it is clear that among the supply rate components, only Qc exhibits a negative derivative dQc/dP in a limited section of the curve. Qt and Qp do not have negative derivatives. It is therefore the negative derivative portion of the Qc curve that Fig. 9. Calculated compound fractions ht and hc for the target and substrate, respectively, vs. supply of the reactive gas, Qtot. P1–P4 are only marked for the ht-curve. Fig. 10. Schematic diagram illustrating the simulation procedure. (a) shows the general behavior for the reactive gas pressure, P, vs. reactive gas flow and (b) shows sputtering rate, R, vs. reactive gas flow. By selecting different values for the partial pressure in (a), the process operates at different points indicated by 1, 2, . . ., 9. These points correspond to points 1, 2, . . ., 9 in (b). S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230 221 causes a negative derivative in the Qtot vs. P curve. Having (or eliminating) hysteresis may be predicted by investigating the conditions for the derivative dQtot/dP. If dQtot/dPb0 in some region, the process will exhibit hysteresis. If dQtot/ dPN0 over the whole processing region, then the process will not exhibit any hysteresis. In Fig. 11b are shown the derivatives of Qtot, Qc, Qt and Qp as a function of P. The following condition is valid, dQtot dP ¼ dQt dP þ dQc dP þ dQp dP : ð7Þ Eq. (7) is quite useful when determining whether hysteresis will appear or not. (Notice that dQp/dP=S, system pumping speed.) The functions Qt, Qc and Qp may also be presented as a function of Qtot. Such a presentation is shown in Fig. 12. 3.1. Influence of target material The sputtering yields Ym for the elemental target material and Yc for the compound formed on the target surface are materials constants. For materials where YcbbYm, hysteresis is normally more pronounced than if Yc is close to Ym [10]. In Fig. 13, three R vs. Qtot curves calculated for different values of Yc, keeping Ym constant, are shown. As can be seen, there is a pronounced difference in hysteresis width for the curves. The model thus predicts that different target materials may respond differently in a reactive sputtering process. 3.2. Influence of reactive gas Some gases are more reactive to the target material than others. Normally, O2 is a more aggressive gas than N2. The influence of gas reactivity may be simulated by changing the value of the sticking coefficient a in the model formulas. In Fig. 14, the results of R vs. Qtot for different values of a are shown. Notice that a decrease in a causes a decrease in the width of the hysteresis region [11]. Also, a may be considered as a materials constant. However, this constant may be influenced by, e.g., target/substrate temperatures, surface morphology, etc. 3.3. Influence of system pumping speed In many reactive sputtering processing chambers there is a throttle valve between the pump and the chamber. The setting of this valve determines the pumping speed for the external pump. It is thus possible to experimentally control the pumping speed. Fig. 15 shows the effect of varying the pumping speed. The calculations predict that the hysteresis may be eliminated if the pumping speed is high enough. This is a well known effect [12] and it can also be seen directly by inspection of Eq. (7) and Fig. 11a–b. Since S=dQp/dP, the value of S may always be chosen high enough to make dQtot/dPN0. Unfortunately, the critical value of the pumping speed needed to eliminate the hysteresis is usually unrealistically high. Fig. 11. (a) shows calculated consumptions of the reactive gas vs. the partial pressure of this gas and (b) shows calculated derivatives of the consumption curves shown in (a). Fig. 12. Calculated consumption of the reactive gas vs. the supply, Qtot, of this gas. Fig. 13. Calculated sputter erosion rates, R, vs. the supply, Qtot, of reactive gas for different values of the sputtering yield Yc for the compound, Ym=1.5. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230222 3.4. Target to substrate distance The position of the target in front of the substrate may also influence the hysteresis width. This is illustrated by changing the size of the collecting area Ac in the calculations. An increasing collecting area Ac represents an increasing target to substrate distance. Fig. 16 shows the processing curves for two different sizes of Ac. Keeping the target to substrate distance small will cause a decrease in the hysteresis width. Notice that there will always be some spread of the sputtered material. Therefore, the condition that AcNAt will always exist in this configuration. Many sputtering systems have shutters relatively close in front of the sputtering cathodes. It is common practice by many users to pre-sputter the target behind a shutter before exposing the substrate to film deposition. The idea is to obtain stable processing conditions before exposing the substrate to deposition. The use of a target shutter in reactive sputter deposition, however, may significantly influence the processing conditions. Pre-sputtering the target in reactive mode behind the shutter will be equivalent to sputtering with a smaller collecting area than the collecting area when the shutter is removed. The reactive sputtering process will then QjumpQ from one processing curve corresponding to the shutter area to another corresponding to the deposited area when the shutter is removed (illustrated by the arrow in Fig. 16). After the shutter removal, a new stabilization period will take place before reaching the new steady-state condition. During this period, the composition of the deposited film may vary significantly. This initial phase of deposition determines the film/substrate interface. This may affect the film adhesion and other film/substrate sensitive properties. However, having a protective shutter close to the substrate instead of close to the target will make it possible to pre-sputter in reactive mode without any significant change in mode of operation when the shutter is removed (minor change in Ac). 3.5. Target ion current At first, it might seem possible to eliminate the hysteresis by applying high power to the target and thus reducing the target poisoning effect by a high sputter erosion rate. Unfortunately, the process does not respond in this way. In Fig. 17, results of calculations for different ion current levels are shown. Increasing the ion current only causes a magnification of the curves. In fact, the calculations predict Fig. 15. Calculated sputter erosion rates, R, vs. reactive gas supply, Qtot, for three different pumping speeds, S. Fig. 16. Calculated processing curves corresponding to sputtering with a target shutter (Ac=700 cm2 ) and without a shutter (Ac=2500 cm2 ). Fig. 17. Calculated sputter erosion rates, R, vs. reactive gas supply, Qtot, for different values of the argon ion current, I. Dashed lines represent constant compositions ht and hc. Fig. 14. Calculated sputter erosion rates, R, vs. reactive gas supply, Qtot, for three different values of the sticking coefficient a. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230 223 that the target poisoning and the film composition effects will be identical along straight lines starting from the origin. It is thus not possible to eliminate the hysteresis by increasing the target ion current. The hysteresis effect may also cause confusion in decisions concerning the processing point. It is often desired to operate the process close to the avalanche point. A suitable operating point for the I=2.5 A curve in Fig. 17 should be point X. This point may be reached by first applying 2.5 A to the target and then slowly increasing Qtot to c13 sccm. Note, however, that, if 13 sccm of reactive gas is supplied to the processing chamber before the target current is applied, the process will end up at processing point XV. Identical sets of values for Qtot and I yield different processing conditions. This serves to illustrate that under certain conditions it is important to report in which order different parameters are applied to the process. Sometimes, the target current or power supplied to the target may be used as a control processing parameter. If a constant supply of reactive gas is fed to the chamber, this will give the curves in Fig. 18a–b for the sputter erosion rate R and partial pressure P as a function of target ion current. Notice that, for low values of the ion current, the target is poisoned, while the metallic target mode corresponds to high values of ion current. This same information can also be obtained from Fig. 17. 3.6. Target area: hysteresis-free operation It is possible to select the size of the target. The model predicts that the shape of the processing curves depend strongly on the size of the beffectiveQ target area. This is shown in Fig. 19a. As can be seen, there will be no hysteresis for a small size target area under these processing conditions [13]. It must be pointed out that this is not an effect caused by the corresponding increase in ion current density (having the same total target current for all target areas). To further illustrate this, calculations for the small area target for different target ion currents are shown in Fig. 19b. As can be seen, the hysteresis is eliminated irrespective of ion current density values. This is a consequence of the results described in Fig. 17. Increasing or decreasing the ion current only causes a magnification or demagnification of the curves. It does not change the shape of the curves. A small target exhibits almost ideal processing conditions. Fig. 19c shows the compound fractions ht and hc corresponding to the same conditions as in Fig. 19a for At=3 cm2 . Note that the target remains primarily in the high rate metallic mode htc0 all the way up to the position where hcc1 (stoichiometric film formation). Fig. 18. Calculated processing curves using ion current, I, as the independent parameter. (a) Target erosion rates, R. (b) Partial pressure of reactive gas, P. Fig. 19. (a) Calculated sputter erosion rates, R, vs. reactive gas supply, Qtot, for different values of the target area, At, assuming a constant total target ion current, I. (b) Calculated sputter erosion rates, R, vs. reactive gas supply, Qtot, for different values of target ion current, I, for the 3-cm2 target in (a). (c) Calculated values for ht and hc vs. reactive gas supply, Qtot, for the 3-cm2 target in (a). S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230224 It should be noticed, however, that the critical target size for hysteresis-free operation will depend on several other processing parameters. The critical size has to be estimated for the specific reactive sputtering process. Notice also that the power dissipation over a relatively small area may give rise to a considerable heating of the target. This has to be considered when designing a hysteresis-free reactive sputtering process based on this phenomenon. 4. Dissociation of sputtered compound molecules: effect on modelling In the above treatment, we assumed that the sputtered compound molecules from the target reached the collecting area without breaking up into free atoms [14]. This is generally not believed to occur for compound molecules during sputtering. During the sputter event, the compound molecule is believed to dissociate into its elemental atoms. It is easy, however, to modify the above model to take molecule dissociation into account. This may be illustrated by the following arguments. (a) In the case of compound dissociation, there will be no net gettering of the reactive gas at the target surface. The reason for this is that the gas molecules that are consumed to form the compound at the target surface will return to the gas when the compound molecule is sputtered and dissociated into its original components. We simply assume that these bgas atomsQ will return back to the gas phase. (b) There will be an additional flux of elemental target atoms to the collecting area Ac. Compound molecules will be sputter eroded from the target at a rate ( J/q)Ychc. The metal fraction of these outsputtered compound molecules will, after compound dissociation, reach the collecting area as elemental target metal atoms. In Fig. 20, results from calculations based on the dissociation of the compound molecules are compared to the results based on sputtered compound molecules. It is observed that there is only a minor difference between the two curves. Detailed comparisons of ht, hc, Qc, Qt and P as a function of Qtot give the same minor deviations as for the R vs. Qtot curve shown in this figure. This illustrates that the status of the sputtered material does not primarily determine the general shapes of these curves. Other parameters have a far more significant influence on the overall process. In the following treatments, we therefore choose to continue to allow sputtered compound molecules to reach the substrate intact. 5. Reactive co-sputtering Complex compound thin films may be deposited by reactive co-sputtering from several elemental targets [15]. The processing behaviour for such a set-up can be predicted by calculations based on the simple model presented above. It is necessary, however, to somewhat modify the conditions at the collecting area compared to that outlined for a single elemental target. We suggest the following treatment. A schematic of the model for the processing situation for a reactive co-sputtering process having two separate targets is shown in Fig. 21. The notation corresponds to the definitions given in Fig. 4a–b. Eq. (2) may be used to define the balances for each separate target. J1 q Yc1ht1 ¼ a12F 1 À ht1Þð ð8aÞ for target 1 and J2 q Yc2ht2 ¼ a22F 1 À ht2ð Þ ð8bÞ for target 2. The conditions at the collecting area Ac are somewhat more complicated than for the simple single target process. We choose to use the material conservation approach (Eq. (3V)) to solve for the overall composition y of the collecting area Ac. The total number T1 of sputtered elemental atoms from target 1 is given by T1 ¼ J1 q A1 Ym1 1 À ht1ð Þ þ Yc1ht1½ Š; ð9aÞ assuming one metal and one gas atom in the compound molecule. A corresponding equation may be defined for the second target, T2 ¼ J2 q A2 Ym2 1 À ht2ð Þ þ Yc2ht2½ Š: ð9bÞ From T1 and T2, the fraction y of metal 1 atoms in the deposited film may be calculated as y ¼ T1 T1 þ T2 : ð9ÞFig. 20. Calculated sputter erosion rates, R, vs. reactive gas supply, Qtot, including the effect of molecular dissociation and neglecting this effect. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230 225 These equations may then be used to calculate the compound fractions hc1 and hc2 at the surface fractions yAc and (1Ày)Ac of the collecting area. Since y was defined to be the fraction of atoms sputtered from target 1, we simply assume that all material sputtered from target 1 is collected at the y fraction of Ac. Therefore, collecting of materials from the two separate targets may be treated separately. A relationship between the partial pressure and supply of the reactive gas can be found that is analogous to the treatment for one gas and one target. Notice, however, that all sub-areas At1(1Àht1), At2(1Àht2), yAc(1Àhc1) and Ac(1Ày)(1Àhc2) consume reactive gas. Calculations based on the assumptions made here are shown in Fig. 22. We have chosen processing conditions to demonstrate that the width and position of the hysteresis obtained by simultaneous sputtering of both targets (cosputtering) may deviate significantly from operating any of the targets individually. Notice also that we only obtain a single hysteresis and not a complex mixture of two hystereses. Due to the different reactivity of the two targets, and also the different ratios Yc/Ym of the targets, the film composition y may vary quite dramatically as a function of the reactive gas supply Qtot. In Fig. 23, the composition y as a function of reactive gas supply is shown for the co-sputtering curve shown in Fig. 22. The loop illustrates that quite complex composition behaviour may be obtained for reactive cosputtering from two separate targets. Increasing the number of targets further increases the complexity. We also want to point out that due to different reactivities between targets and the reactive gas, it may be impossible to operate both targets in the metallic mode (high rate sputtering) and obtain a fully reacted coating. However, this can only be achieved for certain specific compositions. This effect will exist for both reactive alloy and co- sputtering. 6. Reactive sputtering from an alloy target Due to the materials conservation, modeling of reactive sputtering from an alloy target is almost as straight forward as sputtering from a single element target. Under steadystate conditions, the fraction y of one of the materials in the deposited film will be constant (identical to the target alloy composition) and independent of the value of the reactive gas supply. It should be understood, however, that, due to differences in reactivity with the reactive gas, the different metals in the alloy target may end up with quite different compound fractions hc1 and hc2 in the deposited film. It should also be pointed out that, during processing, the surface composition of the alloy target normally will deviate from the target bulk composition. The fact that the overall composition of the sputter-eroded material must correspond to the target bulk composition makes it quite easy to determine the target surface concentration. The relationship between the bulk and surface composition may be found from the condition that the rate of sputtered atom emission from material 1 divided by the rate of emission from material 2 must be equal to the target bulk composition. A detailed description of modelling of reactive alloy sputtering may be found in Ref. [16]. 7. Processing with several reactive gases Sometimes, it may be desirable to form an alloy such as an oxy-nitride by reactive sputtering from a metallic Fig. 21. Schematic diagram of a reactive co-sputtering system having two targets. Fig. 22. Calculated partial pressures, P, vs. supply, Qtot, of the reactive gas for the cases in which the targets are operating individually or both are operating simultaneously. Fig. 23. Calculated fraction, y, of material 1 in mixed deposited film vs. reactive gas supply, Qtot, for reactive co-sputtering (solid line) and reactive alloy sputtering (dashed line). S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230226 target with more than one reactive gas. The basic model may easily be modified to define a set of equations describing such a process. We assume in the following example a process having a single element metallic target sputtered in argon and two reactive gases (e.g., oxygen and nitrogen). A schematic diagram of the target conditions is shown in Fig. 24. Two reactive gases (gas 1 and gas 2) act on the target surface forming compound 1 at surface fraction ht1 and compound 2 at surface fraction ht2. There will be two balance equations defining the target steady-state condition. Analogous to Eq. (2), the balance for gas 1 and compound 1 will be: J q Yc1ht1 ¼ a12F1 1 À ht1 À ht2Þð ð10Þ A corresponding equation may be defined for the second gas. The treatment of the situation at the collecting surface follows the technique outlined in the previous paragraphs. The results of the calculations show that the process response is quite complex. The two gases compete for compound formation at the target and substrate surfaces. Both gases contribute to target poisoning. The gas gettering will strongly depend on the non-poisoned fractions (1Àht1Àht2) of the target and (1Àhc1Àhc2) of the collecting surfaces. Increasing the supply of gas 1, Q1, therefore directly influences the gettering of gas 2 even without altering the supply of gas 2, Q2. This intrinsic feedback behaviour of the process makes process control difficult [17,18]. There is a unique phenomenon that can occur during reactive sputtering with two reactive gases. It is possible to enter the target poisoning mode by increasing the supply of the first gas while keeping the second gas supply constant. During certain conditions, however, it is not possible to return to the metallic (non-poisoned) mode even by decreasing the supply of the first gas to zero. The process may be btrappedQ in the target poisoned mode. The metallic mode can only be reached again if the supply of the second gas is decreased markedly below its original constant value. This phenomenon further complicates reliable control of the two gas reactive sputtering process. However, the effect is predicted by the results of the modelling calculations as shown in Fig. 25. The trapping effect starts to appear for approximately 4 sccm of gas 2. Once gas 1 is increased beyond the point for avalanching to the compound mode, it is not possible to return to the metallic mode even if the gas supply of gas 1 is set to zero. It is necessary to also decrease the supply of gas 2 to be able to return to the metallic mode [19,20]. 8. Transient behaviour: pulsed DC reactive sputtering During the last decade, pulsed DC reactive sputtering has become the dominating technique to power the target during reactive sputter deposition processing. This technique eliminates, or significantly reduces, the arcing problem caused by a thin insulating layer building up on the target surface. In the simulations described above, we have introduced ht to describe the fraction of the target consisting of compound material. If this is an insulating compound layer, it will be positively charged by the bombarding positively charged inert argon ions. A certain net charge will cause a voltage drop over the insulating layer. At a critical electrical field, an electrical breakdown of the insulating layer will take place giving rise to arcing. By changing the polarity (applying a positive voltage) from the power supply over a few percent of the duty cycle, electrons are attracted to the target surface. Proper positive voltage and duty time allow for the complete discharge of the insulating layer. In this way, arcing is prevented despite having some areas of insulating material on the target surface. The frequency of pulsed DC systems is normally in the range of 10–300 kHz. Assuming a frequency of 100 kHz and a target erosion rate of 10 Am/min is equivalent to eroding Fig. 24. Schematic of the target conditions for reactive sputtering with two reactive gases. Fig. 25. Calculated sputter erosion rates, R, vs. reactive gas supplies, Q1 and Q2. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230 227 0.0017 2/pulse from the target surface. Such a small erosion rate per pulse is not enough to markedly change the poisoning status (ht) at the target. This example serves to illustrate that ht will not respond to the frequency of the power supply at these high frequencies. A more detailed analysis [21] shows that similar conditions will be valid for the response of uc to the applied frequency. It should also be pointed out that there will be a transit time before sputter eroded material from the target reaches the collecting area. Due to scattering with gas atoms, there will be a dispersion in transit times for these sputtered particles. At a target to substrate distance of 10 cm and a total pressure of 5 mTorr there will be a substantial spread in transit times for sputtered Al atoms. This is illustrated in Fig. 26. The dashed short pulse (duration 0.1 ms) causes sputter erosion from an Al target. The solid curve shows the probability distribution, f(t), of the transit time for sputtered atoms to reach the substrate. This result indicates that consecutive pulses from the target (frequency N1 kHz) will supply sputtered material to the collecting area with a large time overlap. The higher the frequency, the more overlap between consecutive pulses. The net result will be that the pulsed-character of arriving sputtered material will disappear. Instead, there will be a continuous flux of sputtered material as in the case of pure DC sputtering. This is illustrated in Fig. 27. From the above arguments, we can conclude that the behaviour of the pulsed DC reactive sputtering very much will be the same as for continuous DC sputtering. The positive pulse only serves to discharge insulating parts on the target surface. The gas kinetics and flow of sputtered material will not respond fast enough to the DC pulses applied by the power supply. The short positive pulse, however, may cause some energetic bombardment at the substrate, which may influence the morphology of the deposited film. One advantage of pulsed DC operation is, however, that it allows for higher plasma densities during the bpulse onQ interval as compared to the plasma density obtained for a corresponding continuous DC-plasma having the same average power. 9. Ion implantation: reactive gas We have described some major effects that take place during reactive sputtering processes. As mentioned in the introduction we chose to include as few parameters as possible in the model, while still obtaining results which are well correlating with experimental experiences. We are aware of other effects that also take place during processing. Some of the reactive gas molecules will be ionized and gain energy from the electric field in front of the target and get implanted in the target surface region. Upon penetrating into the surface region of the target, these ions can undergo chemical reactions with the target atoms. This will cause some additional target poisoning and contribute to an increase in ht. In our simplified treatment, however, we assumed that the ion current is solely carried by the argon gas. This is valid when the partial pressure of the reactive gas is small compared to the partial pressure of the argon gas in the processing chamber. A detailed analysis of the effect of ion implantation of reactive gas molecules have been carried out by D. Depla and R. De Gryse. Their theoretical calculations predict that this effect alone may give rise to a hysteresis in the erosion rate vs. partial pressure curves [22]. So far, however, there exists no experimental evidence for such behaviour. The relative importance of reactive gas ion implantation and chemisorption at the target surface remains to be further investigated. Our preliminary studies indicate that the outer 20–50 2 layer at the target surface will be influenced by bknock-onQ effect of energetic argon onto chemically adsorbed and/or implanted reactive gas atoms. However, irrespectively of reactive gas incorporation mechanism, the gas atoms will be evenly distributed within the whole altered target surface layer. Fig. 27. Results corresponding to Fig. 26 but for a pulsed frequency of 5 kHz (dashed line). Fig. 26. Probability distribution, f(t), (solid line) of the transit time for sputtered Al atoms to reach the substrate after one short (0.1 ms) sputtering pulse (dashed line). S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230228 10. Effect of secondary electrons When the target surface composition changes, the amount of secondary electrons emitted by ion bombardment of the target surface changes. This will, in turn, change the ratio between ion and electron current. It is not possible to distinguish the electron and ion currents in the external electrical circuit. Therefore, a constant total current during processing does not necessary imply that there is a constant ion current bombarding the target. If the secondary electron emission coefficient can be estimated for the different target surface conditions, this effect can be included in the modelling work. Some authors have recently done this [23,24]. The results indicate that this effect does not change the general shapes of the calculated curves. To some extent, however, it offers a possibility to include the plasma conditions into the calculations. It is well known that the target voltage will change due to target poisoning if a reactive sputtering process operates in the constant target current mode. This can be predicted by simple calculations. We will briefly summarize the treatment by Pflug et al. [23] who assumed that the total target current IT is composed of both ions and electrons. IT ¼ I 1 þ cð Þ; ð11Þ where I denotes the actual ion current and Ic denotes the electron current generated by the secondary emission coefficient c. The secondary electron emission factor c depends on both the target voltage, U, and the degree of target poisoning. Pflug et al. [23] suggested the expression c ¼ cm 1 À htð Þ þ cchtð Þ U U0 ð12Þ in which cm and cc are the secondary electron emission coefficients at the metal and compound areas, respectively, at the target surface and U0 is a reference voltage required for keeping the equation dimensionless. It is possible to combine these equations with the basic gas kinetic equations outlined in Section 2 above and calculate the variations in target voltage as a function of the reactive gas supply for constant target current. Calculated results following the treatment of Pflug et al. are shown in Fig. 28. It should be noticed that cm may be either smaller or larger than cc. Consequently, the voltage may either increase or decrease for a large supply of reactive gas. It depends on the target material and the choice of reactive gas. We may notice, however, that the processing curve is still S-shaped. It must be pointed out, however, that the results shown in Fig. 28 are based on very simplified assumptions about the electrical characteristics of the plasma. A detailed analytical analysis can not easily be evaluated. 11. Non-uniform target current In magnetron sputtering, the ion current is extremely non-uniform at the target. The current density in the pronounced erosion zone may be significantly higher than outside this zone. We therefore normally estimate the area of the erosion zone and assume that the current acts only on this zone. The shapes of the modeling curves, however, will not change if we assume the target to consist of different segments having different ion current densities. The degree of target poisoning, however, will be different for the different segments. 12. Modelling conclusions The basic model presented in the first sections of this article serve to predict the general shapes of processing curves for a wide variety of reactive sputtering processes. We have also demonstrated that this model can be slightly modified quite easily to account for more complex reactive sputtering processes. It must be pointed out that the simplifications made in the model will, of course, restrict the validity of the calculated results. To our experience, however, the theoretical curves remarkable well mirror most experimental findings. We are well aware of that much work remains to be done before there will exist a reliable detailed model which also accounts for the properties of the plasma, non-uniformities in gas kinetics, more complex chemical reactions, variations in secondary electron emission coefficients, ion implantation of the reactive gas, etc. Several attempts already exist to use finite element method (FEM) calculations to solve for some of these effects [25]. We will continue to search for new simple ways of introducing additional effects into the description of reactive sputtering processes. The basic idea is to find an extended reliable process model, which is able to predict the effects of more individual parameters. Acknowledgements The work presented in this article has primarily been financial supported by the Swedish Foundation for Strategic Research in the former project bThe Angstrom Consortium Fig. 28. Calculated target voltage, U, as a function of the supply of the reactive gas, Qtot, for reactive sputter processing at constant total target current. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230 229 for Thin Film ProcessingQ and the ongoing project bThe Low-Temperature Thin Film Synthesis ProgramQ. Additional financial contributions have been obtained from the Knut and Alice Wallenberg Foundation, the Swedish Research Council for Engineering Sciences and the Swedish Agency for Innovation Systems. Appendix A The following parameters have been used for calculating the curves in the figures, unless otherwise is indicated. Figs. 5–19: S=80 l/s, Yc=0.3, Ym=1.5, a=1, At=150 cm2 , Ac=2500 cm2 , I=0.5 A. Figs. 21 and 22: S=80 l/s, Yc1=0.3, Ym1=1.5, Yc2=0.8, Ym2=3.0, a1=1, a2=0.5, At1=150 cm2 , At2=150 cm2 , Ac=2500 cm2 , I=0.5 A. Fig. 24: S=10 l/s, Yc1=0.03, Ym1=0.4, Yc2=0.03, Ym2=0.4, a1=1, a2=1, At=18 cm2 , Ac=1650 cm2 , I=2 A. Fig. 27: P=100 W, cm=0.4, cc=1, S=80 l/s, Yc=0.3, Ym=1.5, a=1, At=150 cm2 , Ac=2500 cm2 . References [1] I. Safi, Surf. Coat. Technol. 127 (2000) 203. [2] S. Berg, T. Nyberg, H.-O. Blom, C. Nender, in: D.A. Glocker, S.I. Shah (Eds.), Handbook of Thin Film Process Technology, Institute of Physics Publishing, Bristol, UK, 1998, p. A5.3:1. [3] S. Berg, H.-O. Blom, T. Larsson, C. Nender, J. Vac. Sci. Technol., A, Vac. Surf. Films 5 (1987) 202. [4] S. Berg, T. Larsson, C. Nender, H.-O. Blom, J. Appl. Phys. 63 (1988) 887. [5] H. Bartzsch, P. Frach, Surf. Coat. Technol. 142/144 (2001) 192. [6] V.A. Koss, J.L. Vossen, J. Vac. Sci. Technol., A, Vac. Surf. Films 8 (1990) 3791. [7] H. Ofner, R. Zarwasch, E. Rille, H.K. Pulker, J. Vac. Sci. Technol., A, Vac. Surf. Films 9 (1991) 2795. [8] H. Sekiguchi, T. Murakami, A. Kanzawa, T. Imai, T. Honda, J. Vac. Sci. Technol., A, Vac. Surf. Films 14 (1996) 2231. [9] J.F. O’Hanlon, A User’s Guide to Vacuum Technology, John Wiley & Sons, New York, NY, 1980. [10] S. Zhu, F. Wang, W. Wu, L. Xin, C. Hu, S. Yang, S. Geng, M. Li, Y. Xiong, K. Chen, International Journal of Materials and Product Technology, Inderscience Enterprises, Guilin, China, 2001, p. 101. [11] J. Schulte, G. Sobe, Thin Solid Films 324 (1998) 19. [12] S. Kadlec, J. Musil, J. Vyskocil, Vacuum 37 (1987) 729. [13] S. Berg, T. Nyberg, Patent No. WO 03/006703A1 (2001). [14] A.J. Stirling, W.D. Westwood, Thin Solid Films 7 (1971) 1. [15] N. Martin, C. Rousselot, Surf. Coat. Technol. 114 (1999) 235. [16] M. Moradi, C. Nender, S. Berg, H.-O. Blom, A. Belkind, Z. Orban, J. Vac. Sci. Technol., A, Vac. Surf. Films 9 (1991) 619. [17] W.D. Sproul, D.J. Christie, D.C. Carter, S. Berg, T. Nyberg, Proceedings of 46th Annual SVC Technical Conference, Society of Vacuum Coaters, San Francisco, CA, 2003, p. 98. [18] N. Martin, C. Rousselot, J. Vac. Sci. Technol., A, Vac. Surf. Films 17 (1999) 2869. [19] P. Carlsson, C. Nender, H. Barankova, S. Berg, J. Vac. Sci. Technol., A, Vac. Surf. Films 11 (1993) 1534. [20] H. Barankova, S. Berg, P. Carlsson, C. Nender, Thin Solid Films 260 (1995) 181. [21] L.B. Jonsson, T. Nyberg, I. Katardjiev, S. Berg, Thin Solid Films 365 (2000) 43. [22] D. Depla, J. Haemers, R. De Gryse, Plasma Sources Sci. Technol. 11 (2002) 91. [23] A. Pflug, B. Szyszka, V. Sittinger, J. Niemann, Proceedings of 46th Annual SVC Technical Conference, Society of Vacuum Coaters, San Francisco, CA, 2003, p. 241. [24] Y. Matsuda, K. Otomo, H. Fujiyama, Thin Solid Films 390 (2001) 59. [25] A. Pflug, B. Szyszka, J. Niemann, Proceedings of 4th ICCG, Braunschweig, Germany, 2002, p. 101. S. Berg, T. Nyberg / Thin Solid Films 476 (2005) 215–230230