Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods: Process, Properties and Applications
Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods: Process, Properties and Applications
Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods: Process, Properties and Applications
Ebook678 pages3 hours

Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods: Process, Properties and Applications

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Oxynitride thin film technology is rapidly impacting a broad spectrum of applications, ranging from decorative functions (through optoelectronics) to corrosion resistance. Developing a better understanding of the relationships between deposition processes, structure and composition of the deposited films is critical to the continued evolution of these applications. This e-book provides valuable information about the process modeling, fabrication and characterization of metallic oxynitride-based thin films produced by reactive sputtering and some related deposition processes. Its contents are spread in twelve main and concise chapters through which the book thoroughly reviews the bases of oxynitride thin film technology and deposition processes, sputtering processes and the resulting behaviors of these oxynitride thin films. More importantly, the solutions for the growth of oxynitride technology are given in detail with an emphasis on some particular compounds. This is a valuable resource for academic learners studying materials science and industrial coaters, who are concerned not only about fundamental aspects of oxynitride synthesis, but also by their innate material characteristics.

LanguageEnglish
Release dateJun 21, 2013
ISBN9781608051564
Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods: Process, Properties and Applications

Related to Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods

Related ebooks

Materials Science For You

View More

Related articles

Reviews for Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Metallic Oxynitride Thin Films by Reactive Sputtering and Related Deposition Methods - Bentham Science Publishers

    Part 1

    Reactive Process – Experimental, Modelling and Simulation

    Modelling of Reactive Sputter Deposition of Oxynitrides

    Sören Berg¹, Tomas Nyberg¹, *, Diederik Depla²

    ¹Solid State Electronics Division, The Angstrom Laboratory, Uppsala University, Box 534, 75121,Uppsala, Sweden and ²Department of Solid State Sciences, Ghent University, Krijgslaan, 281(S1), 9000Ghent, Belgium

    Abstract

    Reactive sputter deposition is frequently carried out in a mixture of argon and oxygen or nitrogen to obtain oxides and nitrides. The behavior of such single-reactive gas processes has been explained theoretically and verified by numerous industrial thin film deposition applications. However, mixing two reactive gases with the argon sputtering gas in order to carry out reactive sputter deposition of oxy-nitride films is a far more complicated process. A first order simple process model for such a mixed process is presented. Modelling indicates that altering the supply of one gas will not only cause a change of the partial pressure of this gas but also may significantly change the partial pressure of the other reactive gas. Moreover, different reactivities of the reactive gases result in stoichiometries that are very different from the relative reactive gas supplies. This linked behavior between the reactive gases may cause severe process control problems. In a second part of the chapter, a more advanced model is presented, which includes reactive ion implantation and knock-on implantation. First this model is tested vs. experimental results published in literature. Based on the good agreement in the noticed trends, the time dependence of the poisoning behavior of the target is discussed. Finally, the influence of the deposition profile on the hysteresis behavior, and the composition of the oxynitride is discussed, showing the importance of a complete model.

    Keywords: : Sputtering, reactive sputtering, oxynitrides, process modelling, chemical reactivity, target poisoning, process control, hysteresis.


    * Address correspondence to Tomas Nyberg: Angström Laboratory, Uppsala University, Box 534, 751 21 Uppsala, Sweden; Tel: 46-(0)-18-471-3164; Fax: 46-(0)-18-55 50 95; E-mail: tomas.nyberg@angstrom.uu.se

    INTRODUCTION

    Reactive sputtering is a well known technique to deposit oxides and nitrides from different metal targets. Processes involving a single elemental metal target and a gas mixture of argon and one reactive gas (e.g. oxygen or nitrogen) have been

    extensively studied and well described in the literature. Normally this type of reactive sputtering process exhibits a hysteresis behaviour that may cause complications concerning processing stability. It should be understood that despite the relative complicated processing behaviour the one metal one reactive gas sputtering process is the simplest reactive sputtering process to describe. A simple model has been developed that describes the main features of this process [1]. Results from this model may predict the target sputter erosion rate and the partial pressure of the reactive gas as a function of the supply of the reactive gas. Typical results are shown in Figs. 1 and 2.

    Figure 1)

    Schematic of processing curve for partial pressure of reactive gas vs. gas supply using constant sputtering current.

    Figure 2)

    Schematic of processing curve for target sputter erosion vs. supply of reactive gas using constant sputtering current.

    The S-shaped processing curves can only be obtained if the partial pressure of the reactive gas is used as control parameter in a feedback control system. The solid lines define the processing curves if the supply of the reactive gas will be used as control parameter. The dotted segment will be reached if the partial pressure of the reactive gas is used as control parameter. The width of the S-shaped curves defines the hysteresis width of the process. It should be understood that if the supply of the reactive gas is used as the input control parameter it will not be possible to reach the processing points on the S-shaped curves inside the hysteresis width. This gives a fundamental limit for what compositions that can be obtained for the deposited films.

    Replacing the single element metal target with 2 different single element targets or an alloy target consisting of 2 metals significantly changes the processing conditions. Since the two metals generally have different reactivity to the reactive gas it will be necessary to adjust the supply of the reactive gas to a level where the less reactive metal will fully react with this gas. This normally forces the process into a low rate sputtering mode. This low rate may sometimes be as low as one magnitude lower than what is possible to obtain if a single element target was used.

    It is also possible to use one single element target and add 2 reactive gases to the argon processing gas. By this technique it is possible to reactive sputter deposit e.g. oxy-nitrides. Both the nitrogen and the oxygen have to react with the sputtered metal atoms. Also for this system the difference in reactivity of these gases will give new restrictions in the processing control. In addition a complication in controlling the partial pressures of the 2 reactive gases will appear. Varying the partial pressure of one of these gases will affect the partial pressure of the other gas. Therefore quite advanced feed-back control systems are needed to obtain full control of such a reactive sputtering process [2].

    BASIC MODELLING OF SPUTTERING WITH TWO REACTIVE GASES

    In this chapter we will describe somewhat in detail the mechanisms responsible for the processing behaviour for a reactive sputter deposition process carried out from sputtering one single element target in a mixture of argon and 2 reactive gases (oxygen/nitrogen). The purpose of the presentation is to clarify what kind of additional complications will arise that will not be present when sputtering one single element target in a mixture of argon and one reactive gas. We believe that the simplest way to illustrate the differences may be to outline a simplified model of the process. From this model it will be possible to predict processing behaviour caused by the different involved processing parameters. As a first approximation we will follow the description earlier presented in reference [1].

    We will assume a system consisting of one single element target having an area At. Power may be applied to this target generating an argon ion current density J evenly distributed onto the target surface. All sputtered material will be collected and evenly distributed at the collecting surface having area Ac. Two reactive gases (e.g. oxygen and nitrogen) will also be present in the chamber having partial pressures pO and pN respectively. It will be assumed that these partial pressures are so small so that they do not give rise to any contribution to the ion current bombarding the target. Thus only argon ions are responsible for the sputter erosion from the target. There will be a probability that the gases will react with the pure metal atoms at the At and Ac surfaces. These probabilities will be denoted αO and αN for oxygen and nitrogen respectively. For simplicity we will assume the same probabilities at both surfaces. It is far too complicated to include all possible chemical reactions that may occur during processing. However, we will consider one possible effect. It is well known that oxygen normally is more reactive than nitrogen. In fact it may be possible that the oxygen also may react with the nitride and convert it into oxide. This can only happen at surfaces where nitrides have been formed. The probability for this to happen will be denoted αON.

    We assume the following to happen in the processing chamber. Due to the presence of oxygen in the chamber some fractions ΘtO and ΘcO of the surfaces At and Ac respectively will be covered to oxide. The fractions covered by nitride will be denoted ΘtN and ΘcN.

    Based on the assumptions above it is possible to define a number of equations that together determine the expected processing behaviour. The outline in Fig. 3 will serve to illustrate a simplified model of this type of reactive sputtering process.

    Figure 3)

    Schematic of incoming gas fluxes and ion current to the different area fractions at the target.

    We will assume steady state conditions.

    The relation between the partial pressure p and the flux of molecules F (number/ unit area and time) that this pressure generates at all surfaces will be:

    The flux of sputter eroded metal atoms out from the target will be denoted RM.

    Where J denotes the current density of argon ions and e denoted the elemental electronic charge and Ym denotes the partial sputtering yield of metal by the argon ions. Oxide and nitride are formed at the target surface. We will assume that sputtering of these compounds will result in erosion of the corresponding molecules. Consequently the flux of sputter eroded oxide molecules denoted RO will be:

    Where YO denotes the sputtering yield of oxide molecules. For simplicity we neglect that the oxide molecule may decompose into metal and oxygen atoms. Thus a sputtered oxide molecule will be transferred to the collecting area and be deposited as an oxide molecule. The flux of sputter eroded nitride molecules denoted RN will be:

    where YN denotes the sputtering yield of nitride molecules. We assume that also the sputtered nitride molecules will be deposited as nitride molecules.

    Since we have assumed steady state conditions it is possible to define a number of balance equations at the target and the collecting surfaces. At steady state there will be a balance for the nitride coverage at the target where the formation of nitride must be equal the removal of nitride. For simplicity we will assume that the metal nitride will have the form M1N1 and that the oxide molecule will have the form M1O1. This defines the following balance equation for the nitride coverage:

    Where the term 2αONΘtNFO defines the number of nitride molecules at the target surface being converted to oxide molecules by the flux of oxygen molecules to the fraction ΘtN of the target surface already covered by nitride. The factor 2 in the gas terms originates from the fact that one gas molecule contains 2 atoms. Compounds with other stoichiometries should be compensated for by introducing proper constants in the equation.

    The corresponding equation for the oxygen coverage at the target surface will be:

    It is possible to define two balance equations at the collecting area Ac, one for the oxide formation and one for the nitride formation. An illustration for the simple model for the collecting area is shown in Fig. 4.

    Figure 4)

    Schematic of fluxes of sputtered particles from target to substrate.

    The balance for the nitride formation will be:

    And the corresponding equation for the formation of oxide will be:

    In addition to these balance equations it is possible to calculate the total consumption QO and QN of oxygen and nitrogen respectively for every value of the partial pressures of the reactive gases.

    where the terms SpO and SpN denotes the throughput of the gases to the vacuum pump having a pumping speed S. It is also possible to obtain an expression for the target sputter erosion rate R. We will define the sum of the sputtered atoms and molecules as a measure of the target sputter erosion rate R.

    In this simplified presentation R is a direct measure of the sputter eroded target metal atoms. From the defined balance equations above it is possible to calculate both the target sputter erosion rate R and the partial pressures pN and pO of the reactive gases as a function of QO for different constant supplies of nitrogen QN.

    Fig. 5 shows a typical behaviour for the partial pressures pO and pN respectively as a function of QO for a constant supply of nitrogen QN. In this calculation we have assumed that αON = 0.3.

    Figure 5)

    Calculated partial pressures of oxygen and nitrogen vs. supply of oxygen for a constant supply of nitrogen.

    It should be noticed that the partial pressure of nitrogen pN is significantly affected by the supply of oxygen. The reason for this is of course that the added oxygen will contribute to an increase of poisoning of the target resulting in a lower sputter erosion rate. This will of course decrease the nitrogen consumption and consequently show up as an increase of pN.

    Fig. 6 shows the sputter erosion rate R as a function of the oxygen supply QO for different supplies QN of nitrogen. Also here we assume that αON = 0.3 but a somewhat lower YN than in Fig. 5.

    Figure 6)

    Calculated target sputter erosion rates vs. oxygen supply for three different constant supplies of nitrogen.

    The results point out that it is possible to be trapped in the poisoned mode if the supply of nitrogen is sufficiently large. In this situation it is not possible to return to the starting point (zero oxygen supply) without also decreasing the supply of nitrogen. This is a serious limitation that may create significant problems for the process control system.

    The relation between the sputter yields of the nitride, oxide and metal will significantly influence the process behaviour. Results from calculations with the same parameters as in Fig. 5 are shown in Fig. 7.

    Figure 7)

    Calculated target sputter erosion rates for three different constant supplies of nitrogen.

    It should be noticed that for this parameter combination the hysteresis will decrease as the supply of nitrogen is increased and as shown here eventually totally disappear. This prediction has been verified experimentally for reactive sputtering of Zr in a mixture of nitrogen and oxygen [3]. For metals exhibiting this behaviour it may be difficult to form oxynitrides by reactive sputtering.

    It should be understood that all the above outlined calculations are based on quite crude simplifications. Despite this it has turned out that calculated predictions surprisingly well give the same general processing behaviour as observed experimentally. However, this is only true for steady state processing conditions. Investigating dynamic behaviour needs a more detailed treatment of some additional phenomena involved in the target sputter erosion mechanism.

    ADVANCED STRATEGIES FOR REACTIVE SPUTTERING MODELLING

    The straightforward model described in the previous section enables the fast calculation of different process parameters (such as deposition rate, reactive gas partial pressure) as a function of the reactive gas flow rates during sputter deposition of oxynitrides. In this way one can get a good, but qualitatively, idea of the process behaviour. This flexibility can however only be achieved by a simplified description of the processes occurring at the target and the substrate. For a steady state calculation, as shown above, these approximations have a limited impact on the shape of processing curves and the experimental curves can be mimicked by the model. One must however realize that the obtained parameters generally can only have a qualitative meaning. More important, to reach a more quantitative model a better description of the process kinetics is needed. For a one reactive gas/one target combination the target condition is described by the simple model as,

    with no,s the target atomic surface density. The equation describes the balance between compound formation by chemisorption (first term) and the compound sputtering (second term). Based on equation (13), setting F equal to zero, one can calculate a time constant for the sputter cleaning of a fully poisoned target.

    This gives a value of 0.06 s for the sputter cleaning of a fully poisoned Ta target (Yc = 0.18, Id 1.8x10¹⁷ ions.cm-2.s-1 (0.03 A/cm²) and no,s 2x10¹⁵ cm-2). It is common knowledge that the sputter cleaning of a poisoned target [1] takes typically a second or more, indicating that the compound layer is several monolayers thick. An example of the sputter cleaning of a poisoned tantalum target is shown in Fig. 8.

    Figure 8)

    Sputter cleaning of a poisoned Ta target. Prior to the sputter cleaning process the target was sputtered in pure oxygen (0.3 Pa, discharge current 0.3 A) until a stable discharge voltage was reached. After the poisoning process the oxygen gas was replaced by pure argon (0.3 A) and the discharge voltage was recorded during the sputter cleaning process (discharge current 0.3 A). The calculated time constants (τ) for the sputter cleaning of an oxide layer formed by chemisorption (1 monolayer) and an oxide layer formed by direct ion implantation (thickness 2 nm) are indicated.

    One can of course fix the time constant by increasing the number of monolayers in equation (14) by changing the value of the target atomic surface density, or including more monolayers. However, there is a fundamental problem with this approach. Generally speaking, chemisorption is a fast process for the oxidation of the first monolayer, but the formation of a thicker compound layer is too slow to compete with the sputtering process. This simple reasoning shows that there is a need for extra oxidation processes. One, suggested by D. Depla et al. is reactive ion implantation [4]. The formed reactive ions in the plasma are accelerated towards the negative target, and can be implanted in the target. For the typically discharge voltage used during magnetron sputtering, the implantation depth is in the order of 2 nm. The thicker layer gives rise to a time constant, for the same example, of 0.35 seconds, which corresponds with the sputter cleaning experiments (see Fig. 5). A second oxidation process for the target was suggested by Berg et al.: knock-on implantation of chemisorbed species into target. TRIDYN simulations show that chemisorbed oxygen atoms can be knock-on implanted into the target by the bombardment with inert gas ions [5]. The current status of modelling includes therefore three target processes: i) chemisorption, ii) knock-on implantation and direct reactive ion implantation. A model describing the influence of these processes on the reactive behaviour has been published [4] using one reactive gas, and the text below takes the model a step further, i.e., the implementation of these processes for two different reactive gases as in the case of oxynitride deposition.

    Although the focus of reactive sputter modelling has been for a long time on the target processes, the importance of the substrate processes may not be neglected. The shape of the hysteresis behaviour is influenced by the deposition profile, and hence to reach a better fit between model and experiment, the deposition profile should be included. Simulation of deposition profiles can be performed in an analytical way, but using Monte Carlo (MC) based particle trajectory codes, the same goal can be reached. Analytical methods are typically valid for a given set-up, and these methods do not include the details of the deposition set-up. This kind of problems can be tackled using the MC method. An example of such calculation will be discussed in the following paragraphs. Including the deposition profile does not influence the fundamental description of the substrate processes. Most authors describe the compound formation on the substrate by a chemisorption process. The large discrepancy between published sticking coefficient for chemisorption of oxygen on metals, and measured values of the sticking coefficient during deposition shows that the compound formation is better described as the incorporation of the reactive gas in the growing layer [6, 7]. The impact of this result is not clear yet. Nevertheless, the last results just show that the modelling of reactive magnetron sputtering is still a vivid research topic.

    The Target Processes

    The different target processes are schematically shown in Fig. 9.

    Figure 9)

    The different target processes during reactive magnetron processes. The target is characterized by three fractions, i.e., θm, θr and θc.To distinguish with the first bulk layer these surface fractions are indicated in the figure by θsm, θsr and θsc. The latter fraction is related to chemisorption and is only present at the surface. Reactive ions and knock on implanted chemisorbed species are implanted into the target according an implantation profile. The transfer due to the target erosion is depicted schematically be the right to left arrows.

    For one target/one gas combination the three target processes have been described before [4], and the derived equations, in a somewhat different form, are given below.

    The first three equations (a), (b), and (c) describe the target surface condition. The target surface condition is described by three fractions: i) the metal fraction θm, ii) the compound formed by implanted reactive ions θr, and iii) the compound formed by chemisorption of reactive gas molecules θc. Equation (a) describes the time dependence of the metal surface fraction θm. The surface metal fraction is defined by the transfer of metal from the bulk to the surface. This transfer process is defined by the erosion rate vs, and the concentration of metal in the subsurface layer (nm for layer I = 2, in Fig. 9 indicated as θbm). The sputter process removes metal atoms from the surface towards the vacuum. The flux of sputter metal atoms is calculated from the ion current density Id, the metal sputter yield Ym, and the metal surface fraction θm. Another process which reduces the surface metal fraction is the chemisorption of reactive gas molecules. The latter process is defined by the flux of reactive gas molecules F, the sticking coefficient αt, and the surface metal fraction. This process is balanced by the knock-on ion implantation of the chemisorbed species. By the knock-on process the compound formed by chemisorption is converted back into metal. The last term of equation (a) describes this process. Besides the ion current density and the compound fraction formed by chemisorption θc, the knock-on yield β defines this process. The second surface fraction, i.e., the compound fraction formed by ion implantation θr, is described by equation (b). The equation contains only two terms: the transfer of compound from the bulk towards the surface (see Fig. 9, θbr), and the sputter removal of the compound with Yr the compound sputter yield. The last surface fraction, i.e., the compound fraction formed by chemisorption, is described by equation (c). This equation is given by the balance between chemisorption (see description of equation (a)), the knock-on implantation (see description of equation (a)), and the sputter removal of the compound with Yc the compound sputter yield.

    Equations (d) and (e) describe the bulk processes. To describe these processes the target is subdivided in i layers. Material (compound and metal) is transferred from one cell to another by the erosion process, which is described by the first term in both equation (d) and equation (e). This transfer process is defined by the erosion rate vs. The erosion rate is defined by the target surface condition,

    i.e., the target sputter yield is weighted according to the fraction of the different surface species.

    The chemical reaction between the implanted reactive atoms is given by the second term in both equation (d) and equation (e). The reaction rate constant k defines the rate of the chemical reaction. The last term in equation (e) describes the direct ion implantation and the knock-on implantation. It is assumed in a first approximation that the implantation profile p(x) is identical for both mechanisms and that the implantation profile does not change as a function of the target oxidation. The number of reactive gas atoms implanted in the target depends on the reactive gas fraction of the ion current. It is assumed that this fraction is equal to the reactive gas fraction in the plasma. The knock-on implantation process has been described above.

    In reference [4] this model was used to explain several aspects of the reactive gas process for one target/one gas combination. When a second reactive gas is introduced in the vacuum chamber, one can wonder about the target processes. Of course more equations will be needed to describe the presence of the second reactive gas, but in essence little will change.

    Similar equations as in equation (15) can be written down. Equations (17) show the changes needed to describe the interaction of the target with the second reactive gas. As shown in the previous section, the less stable compound (generally the nitride) can be converted into the more stable compound (generally the oxide). The description of this process brings two additional terms in these equations. The first is the chemical reaction between the formed compound nrN with the implanted species nfO, described as kcnrNnfO with kc the reaction rate constant for this process. The second term is similar as described in the previous section, i.e., the chemisorption of the first reactive gas (oxygen) on the formed compound, i.e., 2FOαONθcN with αON the conversion coefficient from oxide to nitride.

    Before studying the time dependence of the poisoning process (see section 3.2), the model was tested by comparing it with published experimental measurements [8]. It is not the goal to completely fit the experimental results, but just to see if the simulated results follow the same trend as the experiment. In the paper of Martin et al. [8] a titanium target is sputtered in two reactive gasses, i.e., oxygen and nitrogen, to deposit an oxynitride coating. Beside the trapping effect discussed in the previous section, they also show some 2D plots indicating three regions as a function of the two reactive gas flows. These three regions correspond with the metal mode, the instability region, and the poisoned mode. A synergy between oxygen and nitrogen has been observed for these 2D diagrams since the boundary delimiting the metallic mode to the complete poisoning of the target is linear and only depends on the oxygen and nitrogen flow rates. This is an interesting experimental result to compare with the simulations. To simulate such a result experimental parameters should be given. This would mean a detailed analysis and dedicated experiments to retrieve these parameters. Instead of this approach, we have reduced to problem to its essence, i.e., the reactive sputtering of a target material in a mixture of two reactive gasses with one gas more reactive than the other. Hence, the reaction rate constant for the first gas is set at 5x10-23 cm³.s-1 while for the second gas its value was five times lower, i.e., 1x10-23 cm³.s-1. A same reasoning was made for the sticking coefficients. For gas 1 the sticking coefficient on the target and substrate was set at 0.15 while 0.05 was used for the second gas. As it is known that the sputter yield of nitrides is higher than oxides, compound 1 has a lower sputter yield (0.05) than compound 2 (0.1). The implantation of oxygen and nitrogen in a metal results in a very similar implantation profile and therefore the parameters defining the Gaussian implantation profile were equal for both gases (2 nm for the implantation depth, and 1 nm for the ion straggle). The same value for the knock on yield was used based on the same reasoning (β = 0.3). The other experimental parameters are given in the caption of Fig. 10.

    Figure 10)

    Hysteresis behaviour of the oxygen and nitrogen pressure as a function of the oxygen flow. Simulation parameters: I = 0.4 A, target size =10 cm², S (pumping speed) = 50 L/s, Substrate area = 1000 cm².

    Fig. 10 shows the hysteresis behaviour of the reactive gases as a function of the oxygen flow rate for different fixed values of the nitrogen flow rate. The addition of nitrogen to the plasma results in a shift of the critical point towards lower oxygen gas flows, and a narrowing of the hysteresis. A better way to describe this behaviour using a 2D plot indicating the critical point as function of the two reactive gas flows. This is shown in Fig. 11.

    Figure 11)

    Critical point as a function of the oxygen and nitrogen flow for the hysteresis shown in Fig. 10.

    An almost linear behaviour is found similar to the experiments by Martin et al. [8]. Also the narrowing of the hysteresis by increasing the N2 flow rate was shown by Martin et al. [8]. In summary, the experimental tendency can be easily mimicked by the experiment. In the following section this linear behaviour can be understood when studying the time dependence of the poisoning process.

    The Time Dependence of the Poisoning Process

    In Fig. 12 the simulated time evolution of the poisoning process is shown for the same experimental parameters as in Fig. 11. First the target is sputtered in a mixture of argon and nitrogen. In Fig. 12 this is indicated by the first grey zone. The target condition is hardly affected by the presence of nitrogen. Indeed, θm the metal fraction remains almost equal to one, i.e., a pure metal target. The reason for this behaviour can be explained by the getter action of the deposited metal. Indeed, the partial pressure remains quite low, and the nitride substrate fraction (right figure) increases up to approximately 0.4. After 30 seconds the oxygen flow rate is increased. Target poisoning sets immediately in, and after 100 seconds the target is completely poisoned.

    Figure 12)

    Time evolution of the partial pressure of nitrogen and oxygen (figure left). Time evolution of the target metal fraction (figure middle). Time evolution of the fraction nitride and oxide on the substrate (figure right).

    From this behaviour one can understand the almost linear behaviour of the critical points as a function of the nitrogen flow (see Fig. 11). Although the target is not affected by the nitrogen gas, a substantial fraction of the deposited metal is converted into nitride. This limits the getter ability for oxygen. Hence, as less oxygen can be consumed by the deposited metal, poisoning starts at a lower oxygen flow. After complete poisoning of the target (θm = 0, 100 seconds), one notices a further change of the substrate condition as indicated

    Enjoying the preview?
    Page 1 of 1