- 1 Introduction
- 2 New precursor chemistry of Group I elements
- 3 New precursor chemistry of Group II elements
- 4 New precursor chemistry of Group III and lanthanide elements
- 5 New precursor chemistry of Group IV elements
- 6 New precursor chemistry of Group V elements
- 7 New precursor chemistry of Group VI and VII elements
- 8 New precursor chemistry of Group VIII, IX, X and XI elements
- 9 New precursor chemistry of Group XII elements
- 10 New precursor chemistry of Group XIII elements
- 11 New precursor chemistry of Group XIV elements
- 12 New precursor chemistry of Group XV elements
- 13 Concluding remarks
Recent developments in molecular precursors for atomic layer deposition
-
Published:16 Nov 2018
-
Special Collection: 2018 ebook collection
A. L. Johnson and J. D. Parish, in Organometallic Chemistry: Volume 42, ed. N. J. Patmore and P. I. P. Elliott, The Royal Society of Chemistry, 2018, vol. 42, pp. 1-53.
Download citation file:
One field of organometallic and materials chemistry that has seen great advancements over the last 20 years is that of atomic layer deposition (ALD), and in particular the development of precursors for the deposition of thin films of highly functional materials. This review focuses on newly developed ALD precursors for metals and metalloid elements (Groups I to XV). New precursors are necessary for a wide range of both established and emerging high-tech applications. A brief overview of recent advances in precursor chemistry is given.
1 Introduction
In recent years the industrial, consumer driven demand for smaller and more sophisticated microelectronic devices has directed research into nanoscale thin film deposition processes such as chemical vapour deposition (CVD) and atomic layer deposition (ALD). Common CVD methods employ volatile compounds, which are vaporised and continuously drawn into a deposition chamber. These precursor compounds chemically react either in the vapour phase or directly on a heated substrate and build up a film of target material, while unwanted volatile components are removed by reduced pressure or an inert gas purge. In the case of CVD, single source precursors (SSPs), i.e. molecules which already contain the necessary atoms in a proper ratio, corresponding to the desired complex phase required, come from a single precursor molecule. Alternatively, multiple source precursors can be used to produce the desired materials, in which two or more precursors act as vectors for individual atoms, the ratio of which in the final desired complex is tuned by varying the amount of precursor introduced to the reaction chamber, incorporating the desired multiple components into the growing film. Because of the fundamental reaction chemistries involved in CVD processes, impurities such as carbon may be introduced into growing thin films from organic ligand combustion products.
Another way to promote the desired chemistry is to employ co-reagents, i.e. separate reactive gases, which are simultaneously introduced into the reaction chamber. Oxygen, water and ozone are commonly used for oxides, and ammonia for nitrides. CVD rapidly covers the substrate in a thin film of the target material, tuneable to nanometres in thickness. However, as device requirements push toward smaller and more spatially demanding structures, methods such as CVD and various physical vapour deposition (PVD) techniques are being superseded by Atomic Layer Deposition (ALD). ALD offers advantages over alternative deposition methods due to its simplicity, reproducibility, control over film composition and thickness, and high conformality of deposited films at the atomic level. These unique characteristics originate from the cyclic, self-saturating nature of ALD processes.1–3
Unlike CVD, ALD is based on sequential surface reactions of molecular precursors separated by an inert gas purge. In this manner, the reactants are kept separated until the adsorbed species react at the surface in a self-limiting process, without the influence of gas phase reactions. Inherent to the ALD process is the ability to accurately control the thickness of the deposited films at an atomic level by simply varying the number of deposition cycles. Another important feature of ALD resides in the self-limiting nature of the surface reactions taking place during deposition, which precludes multilayer growth (i.e. allows control of the film thickness) and can simplify the study of the mechanism of surface grafting and/or film growth. A growth cycle for ALD is shown schematically in Fig. 1. For a more comprehensive summary of ALD and its many applications, the reader is referred to existing reviews on the topic. Several excellent reviews provide an overview of ALD in all of its facets, while this article aims to summarise the most recent advances in new precursor development. However, in order to understand the need for new precursors for ALD, it is important to briefly outline some very general principles that determine precursor properties and aid precursor selection.
The key to ALD is the design and synthesis of new precursors specifically for the aforementioned purposes. For many years precursors for ALD have been sourced from successful precursors originally developed for CVD, and specifically the more reactive precursors used for metalorganic-CVD (MOCVD). In order to achieve the unique characteristics of ALD, and to be suitable as a practical vapour deposition process, precursors must display specific properties.
Ideally, compounds must in the first instance be sufficiently volatile (minimum values of 0.1 Torr equilibrium vapour pressure have been suggested).4 Precursors should also vaporise rapidly and at a reproducible rate, conditions that are most typically met by liquid precursors. In order to achieve a self-terminating surface reaction, precursors must be thermally robust, with no intramolecular reactions or thermal decompositions on the substrate or in the gas phase. However, the precursor must be sufficiently chemically labile to react with the substrate surface and subsequently with the second precursor pulse. The implications of this are that at the point where the precursor has reacted with all the available reactive sites, there is a plateauing of film growth for the cycle, and the following step can begin. This self-limiting growth manifests itself in a typical saturation/growth per cycle plot, as can be seen from the plot in Fig. 2. Assuming that the precursor is thermally stable at the deposition temperature, no further film growth reactions can occur once the maximum surface saturated dose has been delivered. Whilst this holds for an ideal situation, in reality the saturation of 100% of the surface sites is unlikely, primarily because of kinetic and steric factors which limit surface coverage. Concordantly, theoretical growth rate per cycle (GPC) for classical ALD will always be suboptimal. As a result of the properties outlined previously, most ALD growth processes exhibit an “ALD window”; a temperature window where the growth rate is independent of substrate temperature (Fig. 2B). This can vary in size and is highly dependent on precursor.5
Precursors must also be highly reactive towards surface adsorbed species, and towards subsequent co-reagents. High reactivity generally results in relatively fast kinetics, lower ALD temperatures, and faster surface saturation, resulting in shorter cycle times.6
A further precursor requirement is that reaction by-products must be volatile and thus easily purged in order to prepare for the subsequent half-cycle. Moreover, by-products should not be, in and of themselves, corrosive or react with the growing thin film, which would result in non-uniform films, film etching and corrosion of the ALD reactor. Precursors which react with both the surface and with co-reagents exothermically tend to produce high purity films, as ligands are completely removed. This large thermodynamic driving force (large negative ΔG values) also typically allows lower deposition temperatures, which in turn, can result in the formation of high quality films.
For many novel precursor applications which require scale up, thermal ALD processes are often too slow, primarily because the precursors in use find their true application in MOCVD and were never originally designed with ALD chemistries in mind. In these instances, the need to develop new ALD-only precursors has to-date been somewhat circumnavigated by the use of alternative ALD chemistries and the application of more reactive co-reagents (e.g. H2O2 in-place of H2O), or highly reactive gaseous species such as plasmas, radicals, ions and electrons which can be formed by using energy enhanced ALD processes such as plasma enhanced-ALD. By using these energy enhanced processes, less reactive precursors, foraged from MOCVD applications have been able to be applied to ALD.7
The success of ALD is built on chemistry, and unfortunately the time is rapidly approaching where the engineering of ALD processes using “established” precursors is no longer be able to keep pace with demand for both well-known industrially relevant materials and new curiosity-driven materials research. With ALD maturing into the deposition method of choice for many applications, its advancement is restricted by the availability of precursors which can keep pace with the demands of industry, such as low temperature ALD windows and the compatibility of ALD with other upstream or downstream device fabrication processes. For this, and several other reasons, new precursors distributed across the periodic table are needed.
It is important to note that whilst many references are made to ALD properties such as “growth per cycle (GPC)”, direct comparisons between precursors and processes – particularly for different materials – are not absolute, with many other factors exerting influence on this film thickness, including material densities, relative size of atoms, crystallinity and interatomic planes (i.e. wide-spaced laminar materials). Further to this, the authors do not attempt to dwell on the GPC of multinary systems, where the overall growth per cycle is not necessarily equivalent to the sum of a number of independent processes. It must also be put forward that whilst there is a good understanding of the “textbook” ALD process for simple systems, the lines between “atomic layer deposition” and “chemical vapour deposition” become less distinct with processes that move beyond monolayer growth, processes in which physisorption of precursor plays a more important role than surface-based chemical reactions, and processes that do not have the ability to react in the traditional binary fashion. A degree of contention exists over the definition of “true” ALD, and exactly which characteristics of ALD growth are most important in defining it as a process.
This review is intended to give the reader to a brief overview of “new” metal and metalloid precursors focussing on selected examples of some the most recent advances in new precursor development. This review is not intended to be exhaustive, and focusses on what the authors believe to be significant works and contributions to the field within recent years. The authors apologise in advance that not all work can be covered and afforded the same depth of coverage. As such, for chemistries we have not covered, particularly in the field of precursors as yet untested for ALD, we would direct interested readers to reviews in the area.8–11
2 New precursor chemistry of Group I elements
It is perhaps unsurprising that the many advantages of ALD have been turned towards the continued development of thin film battery materials. For some time now, considerable attention has been drawn to the benefits of various ALD coatings for electrodes, which for the most part rely on standard metal oxide materials such as Al2O3 and TiO2, in addition to the development of ALD routes to the electrode materials themselves. Further to this, more recent research has turned towards the fabrication of solid-state electrolytes (SSE) deposited in a similar fashion, and intrinsic to these have been the development of precursors for group 1 metals.
Initial studies into the suitability of a range of lithium compounds as ALD precursors were made by Putkonen et al. in 2009.12 The study focussed on five lithium systems with differing ligand functionalities: Li(THD) (THD=2,2,6,6-tetramethyl-3,5-heptanedionato), LiCp, LinBu, LiNCy2 and LiOtBu, with the lithium β-diketonate, Li(THD), and the lithium alkoxide, LiOtBu, proving most effective at the deposition of Li-containing films. Whilst the volatility of tetrameric Li(THD) is somewhat limited13 with sublimation under operating pressures of ca. 3 mbar occurring in the region of 175–200 °C, films of LiCO3 were successfully deposited with sequential pulsing of precursor and ozone. Depositions were carried out between 185 °C and 300 °C onto soda lime glass, with a consistent growth rate of ca. 0.30 Å per cycle between 185 °C and 225 °C, after which a significantly diminished growth rate is observed. LiNCy2 was found to decompose at temperatures required to ensure volatility, whilst LiCp/H2O and LinBu/H2O deposited films were found to be irreproducible and non-self-limiting respectively. Subsequent work on the efficacy of LiOtBu/H2O as a lithium oxide precursor by Aaltonen et al. saw growth rates of 2.8 Å per cycle when used along with AlMe3/H2O to deposit films of Li2O–Al2O3 at 225 °C.14 It is worth noting that the H2O process towards lithium containing films greatly reduces the amount of carbon in the film and does not lead so readily to the formation of LiCO3, as is observed with the Li(THD)-O3 process. Furthermore, reaction of LiOtBu with H2S and H2S/Al(NMe2)3 has been found to effectively deposit Li2S and LixAlyS thin films respectively, for use as cathodic and SSE materials in LiS batteries.15,16
Alongside the widespread use of Li(THD) and LiOtBu, lithium hexamethyldisilazide (LiHMDS, see Fig. 3) has also been shown to be effective in the deposition of lithium containing films. The reaction of LiHMDS with H2O has become a standard research process, which when coupled with CO2 affords films of lithium carbonate, and when coupled with NH3 affords LiN.17 Further to this, LiHMDS was shown by Hämäläinen et al. to act as a single source lithium and silicon precursor when deposited with ozone pulses.18 Lithium trimethylsilanolate (LiTMSO, Fig. 3) has been investigated by Ruud et al.19 for the deposition of a range of lithium containing thin films, including its use as a single-source precursor for both lithium and silicon within LixSiyOz, in a study that discounted the efficacy of a range of other potential lithium precursors including lithium benzoate, lithium trifluoroacetate and lithium acetate.
Interesting precursor pairings by Nisula et al.20 and Kozen et al.21 have utilised LiHMDS and LiOtBu respectively, coupled with diethyl phosphoramidate, (H2N(O)P(OEt)2), to deposit high quality lithium phosphorus oxynitride (LiPON) films for use as a solid state electrolyte material, whilst other phosphorus precursors such as triethyl phosphate have been shown to yield films of Li3PO4.22 Other more complex structures such as LiFePO4,23 LixAlySizO24 and LixAlyO14 rely on the integration of LiHMDS/H2O or LiOtBu/H2O into multicomponent processes. A recent interest in LiF deposition has led to a range of processes being developed utilising Li(THD)25,26 and LiOtBu27 with TiF4.
The ubiquity of lithium within battery applications has resulted in limited research into the atomic layer deposition of other alkali metal containing films. Despite this, a number of precursor compounds based on the efficacy of their lithium analogues have been explored, and the relatively higher abundance of sodium has resulted in a greater expansion of research into the incorporation of sodium into next generation batteries. Sodium containing thin films were first successfully deposited by Østreng et al.28 in a study that investigated a range of simple sodium and potassium systems including Na/K(HMDS), Na/K(TMSO), Na/K(OtBu). Of the six compounds investigated, both HMDS derivatives and the potassium trimethylsilanolate proved unsuitable for deposition. Sodium and potassium aluminate films were obtained from Na/K(OtBu) using both H2O and O3 as oxygen sources with variable super cycles of AlMe3:H2O pulses. Less success was achieved in an attempt to realise sodium silicate films through the use of NaTMSO, with film conformity proving unachievable. Further research by Stønsteby et al. has utilised the alkali tert-butoxides with H2O in multicycle processes to deposit sodium and potassium perovskite materials via ALD.29 Examples of Group (I) metal precursors used in generating thin film products is shown in Table 1.
Group (I) precursor . | Co-reactant . | Target material . |
---|---|---|
Li(THD) | O3, TiF4 | Li2CO3,12 LiF25,26 |
LiOtBu | H2O, H2S, H2N(O)P(OEt)2, (EtO)3PO, O2 (plasma), TiF4 | (Li,La)xTiyOz,12,30 Li2O–Al2O3,14,31 LixAlyS16 Li2S,15 LiPON,21,32 Li2PO2N,32 Li3PO4,22 LiCoO2,33 LiF27 |
LiHMDS | H2N(O)P(OEt)2, (EtO)3PO, H2O/CO2, NH3, O3 | LiPON,20 Li3PO4,22 LiCO3,17 LiN,17 Li2O,23 LixSiyOz18 |
p-HOC6H4OH | [Li2{p-OC6H4O}]34 | |
p-HO2C–C6H4–CO2H | [Li2{p-O2C–C6H4–CO2}]35 | |
LiTMSO | H2O/CO2, H2O/O3, TMA/O3 | (LiCO3, LixSiyOz, LixAlyOz)19 |
NaOtBu | H2O, O3 | NaxAlyOz,28 (NaNbO3, NaTiO3)29 |
KOtBu | H2O, O3 | KxAlyOz,28 (KNbO3, KTiO3)29 |
NaTHD | HO2C–C6H4–CO2H | [Na2{p-O2C–C6H4–CO2}]35 |
KTHD | HO2C–C6H4–CO2H | [K2{p-O2C–C6H4–CO2}]35 |
RbOtBu | Ti(OiPr)4/H2O | Rb:TiOx36 |
Nb(OEt)5/H2O | RbNbO336 |
Group (I) precursor . | Co-reactant . | Target material . |
---|---|---|
Li(THD) | O3, TiF4 | Li2CO3,12 LiF25,26 |
LiOtBu | H2O, H2S, H2N(O)P(OEt)2, (EtO)3PO, O2 (plasma), TiF4 | (Li,La)xTiyOz,12,30 Li2O–Al2O3,14,31 LixAlyS16 Li2S,15 LiPON,21,32 Li2PO2N,32 Li3PO4,22 LiCoO2,33 LiF27 |
LiHMDS | H2N(O)P(OEt)2, (EtO)3PO, H2O/CO2, NH3, O3 | LiPON,20 Li3PO4,22 LiCO3,17 LiN,17 Li2O,23 LixSiyOz18 |
p-HOC6H4OH | [Li2{p-OC6H4O}]34 | |
p-HO2C–C6H4–CO2H | [Li2{p-O2C–C6H4–CO2}]35 | |
LiTMSO | H2O/CO2, H2O/O3, TMA/O3 | (LiCO3, LixSiyOz, LixAlyOz)19 |
NaOtBu | H2O, O3 | NaxAlyOz,28 (NaNbO3, NaTiO3)29 |
KOtBu | H2O, O3 | KxAlyOz,28 (KNbO3, KTiO3)29 |
NaTHD | HO2C–C6H4–CO2H | [Na2{p-O2C–C6H4–CO2}]35 |
KTHD | HO2C–C6H4–CO2H | [K2{p-O2C–C6H4–CO2}]35 |
RbOtBu | Ti(OiPr)4/H2O | Rb:TiOx36 |
Nb(OEt)5/H2O | RbNbO336 |
Quite remarkably the application range for atomic layer deposition (ALD) has now been extended to include the deposition of rubidium-containing films. Rubidium t-butoxide has been utilised as a precursor for the deposition of rubidium containing materials. The deposition of rubidium containing films is reported as proof of concept through the incorporation of rubidium into two very different systems. Doping amounts of Rb have been added to TiO2 to yield a Rb doped oxide, Rb:TiOx with up to 20 at% Rb. Studies revealed that higher amounts of Rb lead to the uncontrolled formation of rubidium carbonate, as determined by XPS. In the rubidium niobate system, the level of rubidium incorporation can be controlled up to the stoichiometric 1 : 1 Rb:Nb ratio allowing the formation of epitaxial films of perovskite RbNbO3.36
In a significant development porous lithium aryloxide films, [Li2{p-C6H4O2}], have been produced by Karppinen and co-workers, using LiHMDS as the lithium source, along with p-hydroxyquinone (HQ) in an atomic/molecular layer deposition process. Remarkably these Li2Q metal–organic frameworks are crystalline, and undergo the reversible coordination of H2O without the loss of structural integrity.34 In related studies the use of LiTHD, along with NaTHD and KTHD have been used to generate a novel molecular layer deposition process for the production of a family of group 1 metal-terephthate frameworks, i.e. [Li2{p-O2CC6H4CO2}], [Na2{p-O2CC6H4CO2}] and [K2{p-O2CC6H4CO2}] respectively.35
3 New precursor chemistry of Group II elements
Beryllium oxide (BeO) has garnered much interest as a highly thermally conductive high-κ dielectric material, with a dense structure that permits its use as a gas-diffusion barrier.37,38 Due to the primary perceived application of BeO within the microelectronics industry, routes towards the atomic layer deposition of BeO have been readily exploited. As with many other standard ALD processes for which metal-alkyl systems display unparalleled reactivity and volatility, the standard beryllium precursor for this purpose is [BeMe2], and has found use in a vast number of literature studies.39–47 Growth per cycle values of 0.14 nm at 150 °C have been achieved with H2O as an oxidant, whilst at higher temperatures, where the H2O process demonstrates limited growth, the use of O3 facilitates growth rates of up to 0.11 nm per cycle.39
The atomic layer deposition of MgO was first achieved through the use of diethylmagnesium and H2O.48 This process was followed by the widespread use of the sandwich complex bis-cyclopentadienyl magnesium [Mg(Cp)2], or its alkylated derivative, [Mg(η5-C5H4Et)2], with H2O,49–53 along with magnesium β-diketonate, [Mg(THD)2] and H2O2 or O3.54,55 Due to the coordinatively unsaturated magnesium environment within [Mg(THD)2], a tendency towards oligomerisation or the formation of adducts is observed and a wide range of melting points and other reported properties are attributed to this.55–57 Work has been done in an attempt to elucidate these properties and their effect on the efficacy of [Mg(THD)2] as a precursor by Hatenpää et al.,55 where alcoholic adducts of [Mg(THD)2] were found to be monomeric but unsuitable for ALD. Films of MgF2 have subsequently been deposited through the use of [Mg(THD)2] and TiF4, which demonstrated the use of a significantly less hazardous fluoride precursor than more conventional HF, which also exhibits a tendency to etch silica substrates.58,59 The most significant recent advance in magnesium precursor chemistry is the development of monomeric and volatile bis(N,N′-di-sec-butylacetamidinato)magnesium, Fig. 4, [Mg{(iPrN)2CMe}2], by Lou et al. which was used to deposit films of MgxCa1−xO onto GaN substrates.60 The volatilities and potential utility of these amidinate systems was highlighted by Sadiq et al. in 2001, though no deposition was attempted.61 An interesting computational investigation by Kazadojev et al. into the theoretical efficacies of a variety of chelating ligands for MgO and CaO deposition has also been reported.62
Unsurprisingly, the familiar THD complex of calcium has long been utilised in the deposition of calcium containing thin films, with films of CaS, SrS and BaS being deposited in the late 1980s by Tammenmaa et al. via reaction of the group 2 chelates with H2S.63,64 In addition to the low reactivity towards H2O,65 and in line with the other alkaline earths, coordinative unsaturation within these complexes is problematic,66 with [Ca{THD}2] often existing in trimeric form, prompting investigation into monomeric adducts such as the tetraethylenepentamine complex [Ca{THD}2(tetraen)], Fig. 5, which exhibits higher volatility, but slightly lower growth rates than [Ca{THD}2] when used to deposit CaS.67 The latter is likely due to the increased steric hindrance of the spectator amine ligand. A natural advancement in the precursor chemistry of calcium ALD, in an attempt to increase reactivity towards H2O was the utilisation of calcium cyclopentadienyl complexes, with bis(1,2,4-triisopropylcyclopentadienyl)calcium being used in conjunction with H2O by Kukli et al. in 2006 to deposit CaO.68 Precursor couplings of ozone and calcium hexafluoroacetlyaceonate, {hfac}, have subsequently been used to deposit films of CaF,69 whilst more recently, Kim et al. demonstrated the use of novel dimeric bis(N,N′-diisopropylformamidinato)calcium and bis(N,N′-diisopropylacetamidinato)calcium, (Fig. 5) to deposit CaS and CaO.60,70 One of the more interesting developments within calcium ALD is the use of bis(tris(pyrozolyl)borate)calcium, [Ca{Tpb}2], by Saly and co-workers in 2010, to afford CaB2O4 in a reaction with H2O.71 This is a little reported example of a single-source ALD precursor with demonstrable ability to confer two metals to a film in precise stoichiometries. A similar reaction was used to deposit analogous films of SrB2O4 and BaB2O4 with related precursor systems by the same authors.72,73
The use of strontium β-diketonate, [Sr{THD}2] is well established63 within atomic layer deposition, facing many of the same challenges as the magnesium, calcium and barium variants, being oligomeric in the solid state and exhibiting low volatility, low reactivity towards H2O, and undesirable decomposition to incorporate carbon impurities.74–76 In situ [Sr{THD}2] synthesis has in the past sought to eliminate the short shelf-life of the precursor with limited success,77 and more reliable precursor pairings subsequently developed have eliminated many of the disadvantages of [Sr{THD}2]. These processes include the reactions of cyclopentadienyl derivatives [Sr(Cp)2], [Sr(η5-C5Me5)2] and [Sr(η5-C5H3iPr3)2] with both H2O and H2S, and [Sr(η5-C5H4tBu)2] and [Sr(η5-C5Me4Pr)2] with dimethoxyethane and H2O, with all processes successfully depositing the desired strontium chalcogenide films. A variety of adducts of these cyclopentadienyl complexes have also been investigated.75,78–81 Some of this work forms part of an interesting experimental-computational precursor study by Holme and Prinz75 in which a variety of well-documented precursor ligand systems are assessed on the basis of a number of factors including metal-ligand bond strength, electron withdrawing substituents and intra-ligand bond strengths with a view to preventing ligand decomposition. Focus was primarily on cyclopentadienyl complexes in addition to a range of alkylated and fluorinated acetylacetonates, or “acac” compounds, the parent family of the widely used β-diketonate systems. The study indicates the presence of weaker metal-ligand bonding in cyclopentadienyl complexes, in part explaining their higher reactivity towards H2O and oxygen, whilst the presence of weak intra-ligand bonding within acac species supports the propensity of these complexes to undergo thermal decomposition over an extended temperature range.75
In a bid to improve upon the precursor properties of [Sr{THD}2], a novel precursor was synthesised and trialled by Kim et al. in 200782 that featured the addition of a chelating pendant arm that served to disrupt oligomerisation and increase the volatility of the system. The new precursor, [Sr{MTHD}2] (Fig. 6), possesses a melting point 60 °C lower than [Sr{THD}2] and was thermally stable to 400 °C, advantageous over the disputed thermal stability of its more established parent compound.76,82 In a similar vein, the heteroleptic derivative of [Sr{THD}2], [Sr{THD}{demamp}] (demampH={1-[2-(dimethylamino)ethyl]methylamino}-2-methylpropan-2-ol), was described in 2015 by Lee et al.83 The new complex displays an intermediate level of volatility and reactivity towards oxygen, sitting roughly between [Sr(η5-C3H2iPr3)2] and [Sr{THD}2].83 More recently, dimers of strontium and barium imidazolates (Imid) have been shown by Norman et al.84 to deposit the respective group 2 oxides, discussed in greater depth within the barium precursor overview.
Many of the drawbacks associated with the β-diketonate {THD} complexes discussed previously are accentuated within barium deposition, with oligomerisation of the larger alkaline earth metal more pronounced, existing in tetrameric form amongst others, with most non-adducted species susceptible to degradation over time.66,85 However, ALD of BaS has been achieved with [Ba{THD}2] synthesised in situ, as with its strontium counterpart, in addition to more conventionally synthesised precursor.63,86 Significantly more focus has been directed towards the deposition of barium containing films utilising cyclopentadienyl derivatives as ligand systems, such as [Ba(η5-C5H2tBu3)2] · THFx, [Ba(η5-C5Me5)2] · THFx, [Ba(η5-C5H2iPr3)2], [Ba(η5-C5Me4Pr)2] and [Ba(η5-C5Me4C2H4NMe2)2] with O2 plasma or H2O and [Ba(η5-C5Me5)2] with H2S78,80,87–89 (Fig. 7). Systems such as these offer a range of parameters and deposition windows to the prospective user.
More recent investigation into novel barium containing precursor systems has seen highlighted a number of chelating systems based around β-ketiminato backbones in an attempt to increase both volatility and reactivity whilst maintaining the structural integrity of ligands at elevated temperatures.74,90 These complexes have not been trialled for atomic layer deposition to date, but exhibit good thermal stability and volatility, in addition to controllable coordinative saturation. A recent publication by Acharya et al.91 described the successful trial of a novel pyrrole-based barium precursor [Ba(PyMe4)2] with H2O as an oxidant, facilitating low temperature deposition of BaO. No structural detail was described, and with such coordinatively unsaturated systems it is unclear as to the extent of oligomerisation or possible adducts present within the system. Other notable advances in precursor development attest to the application of aforementioned dimers of strontium and barium, di-strontium and di-barium tetra(2-tert-butyl-4,5-di-tert-amylimidazolate), (Imid), Fig. 7, in the deposition of metal oxide films with O3 as an oxidant.84 Examples of Group (II) metal precursors used in generating thin film products are shown in Table 2.
Group (II) precursor . | Co-reactants . | Target material . |
---|---|---|
[BeMe2] | H2O, O3 | BeO39–45,92,93 |
[Mg(Cp)2] | H2O | MgO48 |
[Mg(Cp)2] | H2O, O2 (plasma) | MgO49–51,94 |
[Mg(η5-C5H4Et)2] | H2O | MgO52,53,95 |
[Mg{THD}2] | H2O2, O3, TiF4 | MgO54,55,58,59 |
[Mg{(iPrN)2CMe)}2] | H2O | MgO60 |
[Ca{THD}2] | H2S, H2O, CO2/O3, O3 | CaS,63 CaO,64 CaCO3,65 CaO65 |
[Ca{THD}2(tetraen)] | H2S | CaS67 |
[Ca(Cp)2] | H2O | CaO68 |
[Ca{(iPrN)2CH}2] | H2S | CaS70 |
[Ca{(iPrN)2CMe}2] | H2S, H2O | CaS,70 CaO60 |
[Ca{hfac}2] | O3 | CaF69 |
[Ca{Tpb}2] | H2O | CaB2O471 |
[Sr{THD}2] | H2S, H2O | SrS,63,76 SrO76 |
[Sr(Cp)2] | H2S, H2O | SrS,80 SrO75 |
[Sr(iPr3C5H2)2] | H2O, H2S, O3 | SrS,80 SrO78,96 |
[Sr(PrMe4Cp)2 · dme] | H2O, O3 | SrO75,97 |
[Sr(tBuC5H4)2] | H2O | SrO79,98 |
[Sr{MTHD}2 | O3 | SrO82 |
[Sr{THD}(demamp)] | O3 | SrO83 |
[Sr{Imid}2] | O3 | SrO84 |
[Sr(Tpb)2] | H2O | SrB2O473 |
Ba(THD)2 | H2S | BaS63 |
Ba(C5Me5)2 | H2S, H2O | BaS,80 BaO87 |
Ba(iPr3C5H2)2 | H2O, O2 (plasma) | BaO89 (BaTiO3)99,100 |
Ba(tBu3C5H2)2 | H2O | BaO78,87,88 |
Ba(PrC5Me4)2 | H2O | BaO101 |
Ba(Me2NC2H4C5Me4)2 | H2O | BaO87 |
Ba(imid)2 | O3 | BaO84 |
Ba(PyMe4)2 | H2O | BaO91 |
Ba(TpbEt2)2 | H2O | BaB2O472 |
Group (II) precursor . | Co-reactants . | Target material . |
---|---|---|
[BeMe2] | H2O, O3 | BeO39–45,92,93 |
[Mg(Cp)2] | H2O | MgO48 |
[Mg(Cp)2] | H2O, O2 (plasma) | MgO49–51,94 |
[Mg(η5-C5H4Et)2] | H2O | MgO52,53,95 |
[Mg{THD}2] | H2O2, O3, TiF4 | MgO54,55,58,59 |
[Mg{(iPrN)2CMe)}2] | H2O | MgO60 |
[Ca{THD}2] | H2S, H2O, CO2/O3, O3 | CaS,63 CaO,64 CaCO3,65 CaO65 |
[Ca{THD}2(tetraen)] | H2S | CaS67 |
[Ca(Cp)2] | H2O | CaO68 |
[Ca{(iPrN)2CH}2] | H2S | CaS70 |
[Ca{(iPrN)2CMe}2] | H2S, H2O | CaS,70 CaO60 |
[Ca{hfac}2] | O3 | CaF69 |
[Ca{Tpb}2] | H2O | CaB2O471 |
[Sr{THD}2] | H2S, H2O | SrS,63,76 SrO76 |
[Sr(Cp)2] | H2S, H2O | SrS,80 SrO75 |
[Sr(iPr3C5H2)2] | H2O, H2S, O3 | SrS,80 SrO78,96 |
[Sr(PrMe4Cp)2 · dme] | H2O, O3 | SrO75,97 |
[Sr(tBuC5H4)2] | H2O | SrO79,98 |
[Sr{MTHD}2 | O3 | SrO82 |
[Sr{THD}(demamp)] | O3 | SrO83 |
[Sr{Imid}2] | O3 | SrO84 |
[Sr(Tpb)2] | H2O | SrB2O473 |
Ba(THD)2 | H2S | BaS63 |
Ba(C5Me5)2 | H2S, H2O | BaS,80 BaO87 |
Ba(iPr3C5H2)2 | H2O, O2 (plasma) | BaO89 (BaTiO3)99,100 |
Ba(tBu3C5H2)2 | H2O | BaO78,87,88 |
Ba(PrC5Me4)2 | H2O | BaO101 |
Ba(Me2NC2H4C5Me4)2 | H2O | BaO87 |
Ba(imid)2 | O3 | BaO84 |
Ba(PyMe4)2 | H2O | BaO91 |
Ba(TpbEt2)2 | H2O | BaB2O472 |
4 New precursor chemistry of Group III and lanthanide elements
As with many other oxides of metals, group 3 and their rare earth (lanthanide) counterparts have long attracted attention as versatile materials with a range of potential applications. Because of the range and size of their band gap (4–5 eV) and high relative permittivity or dielectric constant (κ=12–30) together with high band offsets to silicon they have been considered and studied as potential dielectric materials in microelectronic devices. These properties along with others unique to this range of metals, including high thermal conductivities, electroluminescence and mechanical strength, have made ALD processes in which these materials can be deposited long sought after. ALD processes for the deposition of these rare earth metal oxides (RE2O3, RE=Sc, Y La, Ce, Pr, Nd, Pm, Eu, Gd, Tb, Dy, Ho, Er, Tm, Yb and Lu) have conventionally been dominated by systems in which the bulky ligands such as β-diketonate, substituted cyclopentadienyl or bis(trimethylsilyl)amide ligands have been employed in an attempt to satisfy the steric demands of the RE3+ ions. Whilst the volatility and stability of these compounds is well known and many are commercially available, from a precursor perspective these materials are not ideal. Thin films grown from materials such as [RE{THD}3]x require high growth temperatures, as well as activation with strong oxidising agents such as O3 and O2-plasma due to the low reactivity of these systems with H2O. In contrast the tris-cyclopentadienyl complexes, [RE{CpR}3] R=Me, Et, iPr, are considerably more reactive towards H2O, resulting in higher growth rates per cycle, and often display monomeric structures (depending on both the size of the RE3+ ion and the substituent groups on the Cp-ligand), which results in greater volatilities. However, the thermal stability of RE-Cp systems is, in general, lower than the RE β-diketonate complexes, resulting in concomitant MOCVD alongside the ALD, which leads to non-uniform growth rates and an uncontrolled deposition process. While bis(trimethylsilyl)amide complexes, [RE{N(SiMe3)2}3], are generally easily prepared and the bulky {SiMe3} substituents aid volatility, the reactivity of the N–Si bonds often results in the formation of metal silicate thin films and are often avoided when pure RE-oxide materials are desired. While these precursors continue to dominate as “go-to” precursors because of their availability, reproducibility, and known chemistry, there has been some progress in the development of new precursors for rare earth metals. Examples of innovative ALD precursors used for Group III and lanthanide thin films are presented in Table 3.
Precursor . | Co-reactant . | Target material . |
---|---|---|
RE(OCMe2iPr)3 | H2O | Y2O3, LaOx/cubic and hexagonal-La2O3 and Gd2O3.103,105 |
(RE=Y, La, Gd, Pr) | ||
RE(OCiPr3)3 (RE=Y, La, Gd) | ||
RE(OCEt2tBu)3 (RE=Y, La, Gd) | ||
RE(OCMe2CH2OMe)3 (RE=Gd, Pr) | H2O | Gd2O3 and PrOx104 |
[La{(RN)2CMe}3] (R=iPr or tBu) | H2O | La2O3106,107 |
[Er{(tBuN)2CMe}3] | O3 | Er2O3108 |
[M{(iPrN)2CMe}3] (M=Sc, Y) | H2O | Sc2O3,109 Y2O3110 |
[(η5-C5H4Me)2Y{(iPrN)2CMe}] | H2O, H2O/O2 plasma | Y2O3111 |
[(η5-C5H4Me)2Er{(iPrN)2CMe}] | H2O/O2 plasma | Y2O3,112,113 Er2O3114 and Dy2O3113,114 |
H2O/O3 | Pr2O3,113 Gd2O3113 | |
[(η5-C5H4iPr)2RE{(iPrN)2CMe}] (RE=Y, Pr, Gd Dy) | H2O | Hf-doped Y2O3115 |
[Y{(iPrN)2CNMe2}3] | H2O | Y2O3116 |
[Er{(iPrN)2CNMe2}3] | 3,5-(HO2C)2C5H3N | [Er2{(O2C)2C5H3N}3]∞117 |
Precursor . | Co-reactant . | Target material . |
---|---|---|
RE(OCMe2iPr)3 | H2O | Y2O3, LaOx/cubic and hexagonal-La2O3 and Gd2O3.103,105 |
(RE=Y, La, Gd, Pr) | ||
RE(OCiPr3)3 (RE=Y, La, Gd) | ||
RE(OCEt2tBu)3 (RE=Y, La, Gd) | ||
RE(OCMe2CH2OMe)3 (RE=Gd, Pr) | H2O | Gd2O3 and PrOx104 |
[La{(RN)2CMe}3] (R=iPr or tBu) | H2O | La2O3106,107 |
[Er{(tBuN)2CMe}3] | O3 | Er2O3108 |
[M{(iPrN)2CMe}3] (M=Sc, Y) | H2O | Sc2O3,109 Y2O3110 |
[(η5-C5H4Me)2Y{(iPrN)2CMe}] | H2O, H2O/O2 plasma | Y2O3111 |
[(η5-C5H4Me)2Er{(iPrN)2CMe}] | H2O/O2 plasma | Y2O3,112,113 Er2O3114 and Dy2O3113,114 |
H2O/O3 | Pr2O3,113 Gd2O3113 | |
[(η5-C5H4iPr)2RE{(iPrN)2CMe}] (RE=Y, Pr, Gd Dy) | H2O | Hf-doped Y2O3115 |
[Y{(iPrN)2CNMe2}3] | H2O | Y2O3116 |
[Er{(iPrN)2CNMe2}3] | 3,5-(HO2C)2C5H3N | [Er2{(O2C)2C5H3N}3]∞117 |
Homoleptic alkoxide complexes of rare earth complexes are one group of precursors which have been used in the ALD of RE oxide films: The sterically bulky alkoxide complexes [RE{OCMe2iPr}3] (RE=Y, Gd, La Pr), [RE{OCiPr3}3] (RE=Y, Gd, La) and [RE{OCEt2tBu}3] (RE=Y, Gd, La) shown in Fig. 8 have been utilised as ALD precursors, with water as the co-reagent over a range of temperatures (250–400 °C) with varying success. However, not many details regarding the self-limiting nature of the process were reported.102,103 Other examples of Group (III) metal precursors used in generating thin film products are summarised in Table 3. While these systems show some degree of volatility, high TG residues indicated a degree of thermal instability, a feature which was replicated in the ALD experiments displaying inconsistent growth rates. Homoleptic alkoxide complexes (Fig. 8) containing the bidentate ligand 1-methoxy-2-methyl-2-propanolato (mmp) have also been the subject of investigation; PrOx and Gd2O3 films have been both been deposited by liquid injection ALD using a toluene solution of [RE(mmp)3]x/tetraglyme and water as precursors. The growth rate per cycle was found to increase with increasing precursor dose for both PrOx and Gd2O3, indicating the lack of a saturating, self-limiting ALD-type growth.104
The homoleptic N,N′-dialkyl-2-alkyl amidinates of the general formula [RE{(R′N)2CR}3] are the most recent class of ligands to be utilised in the ALD of rare earth oxide thin films. In 2003 Lim et al. first reported the deposition of La2O3 thin films in an water-based ALD process using the tris-amidinates [La{(iPrN)2CMe}3] and [La{(iPrN)2CtBu}3] as precursors.106,107 Growth rates of 0.9 Å per cycle at a deposition temperature of 300 °C were reported, however limited details on the growth behaviour, film composition, or properties were given. Whilst Winter and co-workers108 have reported the synthesis, characterisation and molecular structures of a range of rare earth amidinate complexes, [RE{(tBuN)2CMe}3] (RE=Y, La, Ce, Nd, Eu, Er, Lu), all monomeric in nature, they were significantly less volatile than the related iPr derivatised complexes reported by Gordon.109,110
In addition, the compounds were reported to be significantly less reactive towards H2O and required the use of stronger oxidizing agents such as O3. Of the systems described only the Erbium complex was investigated for ALD, yielding nearly stoichiometric Er2O3 films, with low impurity levels reported. Growth rates for the Er2O3 films were reported to increase from 0.37 Å per cycle at 225 °C to 0.55 Å per cycle at 300 °C.
[Sc{(iPrN)2CMe}3]/H2O and [Y{(iPrN)2CMe}3]/H2O precursor combinations have also been employed for the ALD growth of Sc2O3 and Y2O3 films respectively.109,110 For the Y2O3 process a self-limiting ALD-type growth was observed, with a growth of 0.8 Å per cycle at optimized deposition temperatures, with low impurity concentrations and relatively good dielectric properties (κ=11–12). However, the films were reported to be oxygen rich (O/Y ratio=1.7–2.0). In the case of Sc2O3 only very few details were given on the growth behaviour, film composition or properties however a growth rate of 0.7 Å per cycle was reported.
In an attempt to blend the properties of the rare earth tris-cyclopentadienyl complexes, [RE{Cp}3], with the reactivity observed in the tris-amidinate complexes, heteroleptic species of the general form [(η5-C5H4R′)2RE{(iPrN)2CMe}] (R′=Me or iPr) have been synthesised and investigated (Fig. 8). In 2014 Park et al. described the liquid precursor [(η5-C5H4iPr)2Y{(iPrN)2CMe}]. The heteroleptic precursor, which exhibits a vapour pressure of 1 Torr at 168 °C, displays self-limiting growth in an ALD window from 350 to 450 °C and a growth rate of 0.6 Å per cycle.111 Subsequent studies focusing on the development of mixed Hf/Y oxide materials showed a decreased growth rate of 0.4 Å per cycle at a deposition temperature of 180 °C.115 The related heteroleptic complexes [(η5-C5H4iPr)2RE{(iPrN)2CMe}] (RE=Pr, Gd Dy) as well as the yttrium derivative have also been investigated with both H2O and O3 as the oxygen source. The thermal stability of the precursors was reported to decrease in the order Y>Dy>Gd>Pr (i.e. with increasing ionic radius). In general, growth rates were found to be higher with H2O as the co-reagent compared to O3, except in the case of PrOx deposition.113 In related studies Er2O3, Dy2O3 and Y2O3 have been produced from [(η5-C5H4Me)2Er{(iPrN)2CMe}] and [(η5-C5H4iPr)2RE{(iPrN)2CMe}] (RE=Y or Dy), with H2O (thermal ALD) and O2 plasma (PE-ALD) counter oxidants. Both systems are reported to show ALD growth characteristics.114
Work on the development of rare earth guanidinate complexes has also been initiated with Devi and co-workers, developing both the yttrium and erbium complexes [RE{(iPrN)2CNMe2}3] (RE=Y or Er) (Fig. 8). In the case of [Y{(iPrN)2CNMe2}3] with H2O as the co-reagent, ALD in the temperature range of 175 °C to 250 °C displayed saturative behaviour consistent with an ALD process and a growth rate of 1.1 Å per cycle.116 The corresponding erbium derivative has been exploited in a molecular layer deposition (MLD) process with 3,5-pyridine dicarboxylic acid to generate a luminescent metal organic framework (MOF), [Er2{(O2C)2C5H3N}3]∞, with an exceptional growth rate of 6.4 Å per cycle at 245–280 °C.117
5 New precursor chemistry of Group IV elements
Group 4 metal oxide materials, either as simple binary oxides, i.e. TiO2, ZrO2 and HfO2, or as mixed metal oxides have been the focus of considerable interest for some time. In the case of TiO2, its application in photo-catalysis, as a transparent conducting oxide and as a material for dielectric and microelectronic applications have made it one of the most actively researched materials for deposition by ALD. As a result the topic of TiO2 and the fabrication of thin films by ALD has recently been the subject of an extensive review by Karppinen and co-workers.118 Independently, Devi8 and Blanquart et al.119 also have produced excellent, and extensive, reviews of the precursor chemistry of Group IV elements. Similarly ZrO2 and HfO2 based materials are also of considerable interest, and are currently exploited in high volume manufacturing as high permittivity dielectric materials.
The desire for thin films of TiO2, ZrO2 and HfO2 has driven the development of precursors for some time. However, with the established utility of a number of highly efficient ALD processes based around precursors such as group 4 metal halides, e.g. TiCl4, ZrI4 and HfCl4, metal alkoxides e.g. Ti(OMe)4, Zr(OiPr)4 and Hf(OCMeEt2)4, metal amides e.g. Ti(NMe2)4, Zr(NMeEt)4 and Hf(NMeEt)4, and metal cyclopentadienyl complexes e.g. [CpTi(NMe2)3], [(MeCp)2ZrMe2], [Cp2ZrCl2] and [(MeCp)2Hf(OMe)2], there has been relatively little development in new precursors for the ALD of TiO2, ZrO2 and HfO2.
While Group IV metal β-diketonates have played an extensive role in the development of MOCVD applications there is only one homoleptic β-diketonate complex, namely Zr(THD)4, with ozone, that has been reported as a prospective precursor for the ALD of any Group IV metal oxide. However it is important to note the heteroleptic species [Ti(OiPr)2{THD}2] and [Ti(OtAmyl)2{THD}2] (Fig. 9) have both been shown to achieve self-limiting growth above 360 °C with O3 as the oxygen source.120 Otherwise, this general class of precursors is associated with the low GPC characteristic of β-diketonate complexes (Table 4).
Precursor . | Co-reactant . | Target material . |
---|---|---|
[Ti(OR)2{THD}2] R=iPr or tAmyl | O3 | TiO2120 |
[Ti(OiPr)2(NMe2)2] | H2O and O3 | TiO2121 |
[Hf(OtBu)(NMeEt)3] | O3 | HfO2122 |
[Hf{mmp}2(OtBu)2] | H2O | HfO2123 |
[Ti{dmae}2(OiPr)2] | H2O | TiO2124 |
[Ti{dmap}(NMe2)3] | O2-plasma | TiO2125 |
[Zr{(MeN)2CMe}4] | H2O | ZrO2126 |
[Ti(OiPr)3{(iPrN)2CMe}] | H2O/D2O | TiO2121,127 |
[Ti(NMe2)3{(iPrN)2CNMe2}] | H2O | TiO2128 |
[Ti(NEtMe)3{(iPrN)2CNEtMe}] | H2O or O3 | TiO2129 |
[Zr(NEtMe)3{(iPrN)2CNEtMe}] | H2O or O3 | ZrO2129 |
[Zr(NEtMe)2{(iPrN)2CNEtMe}2] [Hf(NMe2)2{(iPrN)CNMe2}2] | H2O or O3 | ZrO2130 |
H2O or O3 | HfO2131 | |
[(η5-C5H5)Ti(η7-C7H7)] | O3 | TiO2132 |
[(η5-C5H4-Me)Zr(η7-C7H7)] | O3 | ZrO2132 |
Precursor . | Co-reactant . | Target material . |
---|---|---|
[Ti(OR)2{THD}2] R=iPr or tAmyl | O3 | TiO2120 |
[Ti(OiPr)2(NMe2)2] | H2O and O3 | TiO2121 |
[Hf(OtBu)(NMeEt)3] | O3 | HfO2122 |
[Hf{mmp}2(OtBu)2] | H2O | HfO2123 |
[Ti{dmae}2(OiPr)2] | H2O | TiO2124 |
[Ti{dmap}(NMe2)3] | O2-plasma | TiO2125 |
[Zr{(MeN)2CMe}4] | H2O | ZrO2126 |
[Ti(OiPr)3{(iPrN)2CMe}] | H2O/D2O | TiO2121,127 |
[Ti(NMe2)3{(iPrN)2CNMe2}] | H2O | TiO2128 |
[Ti(NEtMe)3{(iPrN)2CNEtMe}] | H2O or O3 | TiO2129 |
[Zr(NEtMe)3{(iPrN)2CNEtMe}] | H2O or O3 | ZrO2129 |
[Zr(NEtMe)2{(iPrN)2CNEtMe}2] [Hf(NMe2)2{(iPrN)CNMe2}2] | H2O or O3 | ZrO2130 |
H2O or O3 | HfO2131 | |
[(η5-C5H5)Ti(η7-C7H7)] | O3 | TiO2132 |
[(η5-C5H4-Me)Zr(η7-C7H7)] | O3 | ZrO2132 |
Whilst a range of homo- and heteroleptic metal amides and metal alkoxide complexes of Group IV are well established, they have lent themselves as suitable departure points for the limited number of new Ti, Zr and Hf ALD precursors that have been developed recently. The mixed alkoxide-alkylamide compounds, [Ti(OiPr)2(NMe2)2]121 and [Hf(OtBu)(NEtMe)3]122 have both been used successfully for the ALD of TiO2 and HfO2, respectively. Whilst homoleptic alkoxide and alkylamide precursors usually have thermal stabilities well below 300 °C, these heteroleptic compounds exhibited self-limiting growth above 300 °C, with both processes displaying high growth rates and affording oxide films of high purity. TiO2 grown from [Ti(OiPr)2(NMe2)2] and either ozone or water, did not exhibit a constant GPC, however self-limiting growth was confirmed at 325 °C with a remarkably high GPC, for TiO2, of 0.8 Å. HfO2 was similarly grown using a [Hf(OtBu)(NEtMe)3]/O3 ALD process at 300 °C, and a GPC of 1.6 Å.122
In a similar vein, donor-functionalised alkoxide precursors have also been developed. Although several alkoxide−donor-functionalized alkoxide precursors have been reported in the literature, only the studies on the complexes [Hf(OtBu)2(mmp)2]123 (Hmmp=HOCMe2CH2OMe, 1-methoxy-2-methylpropan-2-ol) or [Ti(OiPr)2(dmae)2]124 (dmae=dimethylaminoethoxide) using water for the generation of HfO2 and TiO2 thin films reported self-limiting growth. Disappointingly high carbon and hydrogen contamination was found in the films grown from [Hf(OtBu)2(mmp)2]. The self-limited growth of TiO2 from [Ti(OiPr)2(dmae)2] was demonstrated at the low temperature of 100 °C, but no compositional data relating to the thin films were reported. TiO2 thin films have also been produced using the novel precursors [Ti(NMe2)3{dmap}] (Hdmap=1-dimethylamino-2-propanol). Plasma enhanced ALD using [Ti(NMe2)3{dmap}]/O2 plasma has allowed the deposition of high quality TiO2 in an ALD window between 60 and 120 °C at a high growth rate of 0.9 Å per cycle.125
Amidinate ligands have been shown to significantly improve the thermal stability of M–N containing compounds whilst maintaining the reactivity of related metal amide systems, resulting in amidinate complexes having good reactivity with ozone and water. Their high reactivity results in the efficient removal of the ligands, and low impurity levels in the deposited films. Whilst a number of amidinate compounds have been reported, very few have reported the application of these systems in ALD processes. To date only tetrakis(N,N′-dimethylacetamidinate) zirconium, [Zr{(MeN)2CMe}4]126 and the alkoxide-amidinate precursor [Ti(OiPr)3{(iPrN)2CMe}]121 have been reported for the ALD of Group 4 metal containing materials, i.e. ZrO2 and TiO2 respectively. In the case of [Zr{(MeN)2CMe}4]/H2O self-limiting growth of ZrO2 could be achieved at 300 °C, but with a low GPC (0.24 Å). Unfortunately, the study did not involve any compositional analysis.126 In contrast [Ti(OiPr)3{(iPrN)2CMe}], which is a liquid at room temperature with a low evaporation temperature (65 °C under 5–10 mbar), presented good reactivity toward both ozone and water. Using water as the oxygen source, the deposition exhibited a region of constant GPC between 300 and 350 °C. Self-limiting growth was confirmed at 325 °C with saturation of the GPC at 0.5 Å. Deposition within the temperature range 275–375 °C, resulted in what was described as “very pure” thin films with carbon and nitrogen contamination below 1 at.%. Notwithstanding the deposition temperature, the titanium dioxide films deposited adopted the anatase phase. Whilst the ozone process presented nearly identical data with respect to purity, crystallinity and growth behaviour, the ALD window was extended from 275 to 350 °C with ozone as the oxygen source. However, saturated GPC was slightly lower (0.4 Å).121 In situ mechanistic studies have also been performed on [Ti(OiPr)3{(iPrN)2CMe}]/D2O using quartz crystal microbalance and quadrupole mass spectroscopy at 275 °C.127
Whilst amidinate complexes of Group IV metals have, in general, been little explored, the related guanidinate systems have seen more attention. The mono guanidinate complexes [Ti(NMe2)3{(iPrN)CNMe2}],128 [Ti(NMeEt)3{(iPrN)CNMeEt}]129 and [Zr(NMeEt)3{(iPrN)CNMeEt}]129 have been investigated as ALD precursors. The investigation into the ALD of TiO2 from [Ti(NMe2)3{(iPrN)CNMe2}] highlighted a precursor with a thermal stability and a process with a high GPC, and self-limiting growth confirmed at 330 °C.128 The related monoguanidinate compounds [Ti(NMeEt)3{(iPrN)CNMeEt}] and [Zr(NMeEt)3{(iPrN)CNMeEt}] were also evaluated and exhibited similar properties. Evaluation of these complexes as ALD precursors was performed with both H2O and ozone as the oxygen source. Using water, both precursors exhibited self-limiting behaviour at 275 °C. Both water and ozone processes presented an ALD window of between 225–325 °C, with a GPC of 0.4 Å for the growth of TiO2 across this range, independent of the oxygen source, compared to a growth of 0.8 Å for ZrO2 using water. For the ALD process with ozone as the oxygen source, the GPC increased steadily with the deposition temperature (1.0 Å per cycle).129
In related studies both the zirconium and hafnium bis-guanidinate complexes [Zr(NMeEt)2{(iPrN)CNMeEt}2]130 and [Hf(NMe2)2{(iPrN)CNMe2}2]131 precursors have been evaluated for the ALD of ZrO2 and HfO2 thin films. The zirconium precursor, when water was used as the oxygen source, displayed a steady GPC (0.85 Å) over a temperature range of 225 to 400 °C, whereas with ozone the GPC increased steadily with the deposition temperature from 0.8 Å at 250 °C to 1.3 Å at 400 °C. The characteristic self-limiting ALD growth was confirmed by the study of the GPC as a function of the precursor pulse length at 300 °C, with high GPCs of 0.90 and 1.15 Å per cycle with water and ozone, respectively.130 For the hafnium complex self-limiting growth was achieved at 200 °C with a high GPC of 1.2 Å.131
As noted above, Group IV metallocene complexes have long been established as ALD precursors for the deposition of thin films by ALD.119 Recently this family has been extended beyond cyclopentadienyl complexes to introduce cycloheptatrienyl complexes, specifically [(Cp)Ti(CHT)] and [(MeCp)Zr(CHT)] (CHT=C7H7), for the ALD of TiO2 and ZrO2. These precursors do not react with water and require ozone as the oxygen source. The zirconium compound appeared to perform well with self-limiting growth occurring up to a temperature of 350 °C, with a GPC rate of 0.8 Å. Contrastingly the titanium compound was significantly less thermally stable (Tmax=300 °C) and had only a modest GPC (0.35 Å). Interestingly, results show that the films deposited from [(Cp)Ti(CHT)] had a strong tendency to form the high permittivity rutile phase upon annealing at temperatures below 600 °C.132
It should be noted that although there are a plethora of reports which feature established materials, such as TiN, as well as new materials, e.g. TiS2, which feature ALD as the primary fabrication process, these materials in and of themselves have not resulted in the design, synthesis and development of new precursors, relying on established metal precursors, such as TiCl4, with co-reagents other than H2O and O3.
6 New precursor chemistry of Group V elements
As with group 4 precursors, much focus in the development of group 5 precursors has focused on their development for the generation of oxide materials, and as with oxides of the Group 4 elements oxides of vanadium, niobium and tantalum, have been of primary interest because of their application in electronic devices. Nb2O5 and Ta2O5 are both high-κ materials, with large band gaps that have been investigated for numerous applications such as antireflective coatings, dielectric layers and diffusion barriers.119 In stark contrast to Nb2O5 and Ta2O5, the properties of vanadium oxide thin films depend significantly on their stoichiometry, structure, and morphology.133 Vanadium exists in many stable oxidation states, and as such results in a number of oxide materials, i.e. VO, V2O3, VO2 and V2O2, and in several polymorphic forms, therefore depositing thin films of vanadium oxide with a specific oxidation state and structure has been a challenge, with particular attention focused on +4 and +5 oxidation states of vanadium oxide. Vanadium dioxide (VO2) displays a semiconductor-to-metal phase transition that occurs close to 67 °C which is accompanied by an abrupt change in its resistivity and near-infrared transmission. This property makes VO2 of interest for a number of applications including smart windows, resistive memories, and switches in microelectronics.134 Vanadium pentoxide (V2O5) meanwhile, has potential applications as an electrode in lithium ion batteries.135
As with Group 4, ALD processes which use the Group 5 precursors are dominated by a handful of highly established ALD processes based around metal halides i.e. VCl4, NbCl5, NbF5, TaCl5, TaF5 and TaI5, metal alkoxides and oxo-alkoxides e.g. [OV(OiPr)3], [V(OtBu)4], [Nb(OEt)5] and [Ta(OEt)5], and metal amide and metal amide-imido complexes e.g. [V(NEtMe)4], [Ta(NMe2)5], [Ta(NEt2)5], [Nb(NtBu)(NEtMe)3] and [Ta(NtBu)(NEt2)3], all of which can, and do, act as suitable departure points for precursor modification and development.119
A summary of Group V complexes used as precursors for thin films is shown in Table 5. Amongst the precursors reported for the ALD growth of vanadium based materials, only the compounds [OV(OiPr)3], [V(NEtMe)4] and [V{(iPrN)2CMe}3] (Fig. 10) have been reported for vanadium oxide deposition. [OV(OiPr)3] is the common precursor for the ALD of vanadium oxide, and the deposition process is well characterized, however OV(OiPr)3 has limited thermal stability (below 170 °C) and only a low GPC.136,137
Precursor . | Co-reactant . | Target material . |
---|---|---|
[OV(OiPr)3] | H2O | V2O5136,137 |
[V(NEtMe)4] | H2O | VOx138 |
O3 | VO2139 | |
[V{(iPrN)2CMe}3] | H2O2 (anneal in H2) | V2O3140 |
H2O/H2 | VO2140 | |
O3 or H2O/O2 | V2O5140 | |
H2S | VS4141 | |
[In{(iPrN)2CMe}3]/H2S | VxIn(2−x)S3142 | |
[{tBuN}Nb(NEt2)3] | H2O | Nb2O3143 |
O3 | ||
[{tBuN}Nb(NEtMe)3] | H2O | Nb2O3143 |
O3 | ||
[{tBuN}Ta(NEt2)3] | H2O | Ta2O3144,145 |
O3 | ||
[{tBuN}Ta(NEtMe)3] | H2O | Ta2O3145 |
[(tBuN)Ta{3,5-di-tert-butylpyrazolate}3] | O3 | Ta2O3146 |
[(tBuN)Ta{(iPrN)2CMe}2(NMe2)] | H2O | Ta2O3147 |
Precursor . | Co-reactant . | Target material . |
---|---|---|
[OV(OiPr)3] | H2O | V2O5136,137 |
[V(NEtMe)4] | H2O | VOx138 |
O3 | VO2139 | |
[V{(iPrN)2CMe}3] | H2O2 (anneal in H2) | V2O3140 |
H2O/H2 | VO2140 | |
O3 or H2O/O2 | V2O5140 | |
H2S | VS4141 | |
[In{(iPrN)2CMe}3]/H2S | VxIn(2−x)S3142 | |
[{tBuN}Nb(NEt2)3] | H2O | Nb2O3143 |
O3 | ||
[{tBuN}Nb(NEtMe)3] | H2O | Nb2O3143 |
O3 | ||
[{tBuN}Ta(NEt2)3] | H2O | Ta2O3144,145 |
O3 | ||
[{tBuN}Ta(NEtMe)3] | H2O | Ta2O3145 |
[(tBuN)Ta{3,5-di-tert-butylpyrazolate}3] | O3 | Ta2O3146 |
[(tBuN)Ta{(iPrN)2CMe}2(NMe2)] | H2O | Ta2O3147 |
ALD of the amide system [V(NEtMe)4], with water as the oxygen source displays a constant GPC of 0.5 Å in the temperature range of 125 to 200 °C. Self-limiting film growth was confirmed at 150 °C with both ozone and water. The water process had the lower GPC of 0.45 Å. While the films were reported to be smooth and possess low degrees of contamination (C impurities) at 200 °C, it was found that the vanadium oxide thin films contained a mixture of +4 and +5 oxidation states of vanadium.138 Greater control over oxidation state of the final materials was found to be achievable when ozone was used a co-reagent in the ALD process alongside [V(NEtMe)4]. Self-limited growth was established at 200 °C, with a GPC of 0.3 Å. Analysis of the thin films by XPS revealed the films to be VO2.139
Hock and co-workers have recently reported the development of the vanadium 3+ precursor [V{(iPrN)2CMe}3]. As part of an investigation into the application of this precursor with various oxidising agents, the reaction of [V{(iPrN)2CMe}3] with H2O, H2O2 and O3 was monitored in situ using quartz crystal microbalance (QCM) and quadrupole mass spectroscopy (QMS). With ozone, self-limiting growth of V2O5 was reported from 150 to 225 °C with QCM studies showing an average mass gain of 45.3±3.4 ng cm−2 per cycle. In contrast, reaction with O2 showed no mass gain during in situ monitoring, however when O2 was dosed after H2O ALD growth at 200 °C was observed (average mass gain of 17.7±0.9 ng cm−2 per cycle). Whilst the as-deposited films were amorphous by Raman and XRD spectroscopy, after annealing in a N2 atmosphere at 450 °C films were identified as V2O5. ALD (200 °C) with H2O2 followed by a post deposition anneal in H2 showed the presence of V2O3 by Raman spectroscopy. In an attempt to encourage selective V4+ deposition, H2 (4%) was added the ALD sequence after a H2O2 pulse. Self-limiting growth of VO2 was reported at 200 °C with QCM studies showing an average mass gain of 17.5±1.3 ng cm−2 per cycle.140 In related studies the same group have reported the ALD of VS4, (Parónite), in which [V{(iPrN)2CMe}3] reacts with H2S to oxidise V3+ to V4+ while reducing S2− (from H2S) to {S22−}. Self-limiting growth of VS4 was established by QCM between 150 ° C and 200 °C, with a GPC of 0.33 Å.141 [V{(iPrN)2CMe}3] has also been used as a co-reagent in the ALD of thin films of VxIn(2−x)S3.142
As noted earlier, the imido-amide complexes [{tBuN}Nb(NEt2)3]143 [{tBuN}Nb(NEtMe)3],143 [{tBuN}Ta(NEt2)3],144,145 and [{tBuN}Ta(NEtMe)3]145 have all been reported as precursors for the growth of Nb2O5 and Ta2O5 respectively, with water as the oxidant at 275 °C (Nb) and 325 or 350 °C (Ta), with GPC values of 0.4 Å 0.5 Å, 0.65 Å and 0.6 Å reported. With the exception of [{tBuN}Ta(NEtMe)3] these precursors have also been utilised alongside ozone as the oxygen source at identical temperatures.143,145
As a derivative of the imido-amide complexes, the highly thermally stable pyrazolate complex [(tBuN)Ta{3,5-di-tert-butylpyrazolate}3], has been utilised as an ALD precursor for the formation of Ta2O3. With a relatively low vapour pressure, a source temperature of 160 °C under a reduced reactor pressure of 2–3 mbar was necessary to sublime the precursor. Self-limiting growth was confirmed at 325 °C when ozone was used as the oxygen source. Analysis of the thin films indicated low amounts of impurities in the films, however the GPC was low (0.3 Å) compared to more volatile complexes in this class.146
With the proliferation of amidinate ligands across the periodic table in a wide array of ALD precursors it is perhaps not surprising that Group 5 metals have also received attention. The heteroleptic amide-imido-amidinate complex compound, [(tBuN)Ta{(iPrN)2CMe}2(NMe2)]147 has been investigated for the ALD of Ta2O5, using water as the oxygen source. Self-limiting growth was confirmed at 325 °C with a rather low GPC of 0.28 Å.
7 New precursor chemistry of Group VI and VII elements
Chromium, molybdenum and tungsten precursors for application in ALD are extensively focused around the development of precursors for the deposition of metal oxide thin films, such as CuCrO2, Cr2O3, MoO3 and WO3, with the exception of tungsten, where the ALD of metallic tungsten thin films (W) as well as tungsten nitride (WxN) dominates. Precursors for these materials have been reviewed elsewhere,148 however, as with other elemental groups, established precursors for ALD are selected from a handful of reactive MOCVD precursors, specifically species such as CrO2Cl2, Cr(acac)3, MoCl5 and WF6. For the metal nitrides, MoxN and WxN, the bis-imido-amide complexes [(tBuN)2M(NMe2)2] (M= Mo or W) have been comprehensively explored. Similarly for Group VII elements where only manganese has been investigated for the production of MoOx, MnS, MnTe and MnAs thin films by ALD, established precursors include MnCl2, [Mn{THD}3], [(η5-C5H4-Me)2Mn] and [(η5-C5H4-Et)2Mn].148
Despite the relative inactivity with the arena of precursor development over the past two decades several new precursors have been reported recently. Kalutarage et al. reported the development of a family of transition metal (Cr, Mn, Fe, Co, Ni, Cu) complexes bearing α-imino alkoxide ligands (Fig. 11). These complexes can reportedly be sublimed, between 90 and 160 °C at 0.05 Torr, on ∼0.5 g scales in ∼3 h with <5% nonvolatile residues. In the case of the manganese derivative [Mn2{tBuNCHC(tBu)(Me)O}4] the sublimation temperature is higher than that of the other complexes in this range, because of its dimeric structure. ALD at 225 and 180 °C for the Mn and Cr complexes, [M{tBuNCHC(tBu)(Me)O}2] (Fig. 11), with the BH3(NHMe2), showed self-limiting growth of metallic Cr, and in the case of the manganese precursor, MnOx films respectively, as identified by XPS analysis: oxide generation is thought to be the result of ex situ oxidation. For the Cr precursor the GPC was reported to be 0.08 Å and for the Mn precursor 0.095 Å.149 Subsequent investigation of the manganese precursor [Mn2{tBuNCHC(tBu)(Me)O}4] demonstrated its utility as a manganese vector in the deposition of Cu/Mn alloy thin films alongside the copper precursor [Cu{OCHMeCH2NMe2}2] and the reducing agent BH3(NHMe2). Deposition rates of about 0.09 Å were reported on a variety of substrates, with Cu : Mn ratios of about 70 : 30 obtained by controlling the number of Cu and Mn cycles. X-ray photoelectron spectroscopy supported the presence of metallic Cu and Mn within the Cu/Mn alloys.150
Chabal and co-workers reported a novel molybdenum precursor, [(η5-C5H4-Si(CH3)3)Mo(CO)2{η3-2-CH2CMeCH2}], and its use with ozone as the co-reactant in an effort to grow MoO3 films at low temperatures by ALD. The ALD process at 250 and 300 °C is characterized by a nucleation delay attributed to the reaction barrier for the initial reaction of the metal precursor and the formation of the surface products. The results suggest an initial delay in the ALD process, which can be negated if higher substrate temperatures (350 °C) during the initial Mo precursor pulses are used, after which a lower-temperature (250–300 °C) ALD process is possible with no incubation period, yielding what are reported as good quality MoO3 films.151
Atomic layer deposition of crystalline molybdenum oxide thin films α-MoO3, β-MoO3, and an unidentified sub-oxide MoOx phase (2.75≤x≤2.89) has been demonstrated using a new, commercially available precursor [MoO2(THD)2] with O3 at 250 °C. The ALD process (approx. 1 Å per cycle) reportedly displays a well-controlled film growth with good uniformity and conformality with low levels of impurities. Studies showed that the relative amounts of the different phases could be controlled by changing the deposition conditions. Post-deposition annealing treatments studied by high-temperature X-ray diffraction enabled further control of film composition. In particular, single-phase α-MoO3 and MoO2 films were achieved by annealing in O2 and ‘forming gas’, respectively.152 Marks and co-workers have also reported the synthesis and design of new oxo-complexes for the fabrication MoO3 and WO3. Remarkably, these two isostructural precursors [MoO2{(tBuN)2CMe}2] and [WO2{(tBuN)2CMe}2] (Fig. 11) display very different reaction profiles. Using the molybdenum precursor and O3 amorphous, ultrathin molybdenum oxynitride (approximate composition MoO2.48N0.18, as determined by XPS) films are grown at 200 °C. In contrast using [WO2{(tBuN)2CMe}2] and H2O (200 °C) yields amorphous WO3 films. For both processes GPC data is not reported. Despite attempts to produce effective ALD growth with [MoO2{(tBuN)2CMe}2]/H2O and [WO2{(tBuN)2CMe}2]/O3 combinations all attempts failed underscoring a very different reaction chemistry for the two isostructural complexes, the origin of which is not clearly understood.153 Related studies by the same group have gone on to develop the cyclohexyl and isopropyl derivatives, [MoO2{(CyN)2CMe}2] and [MoO2{(iPrN)2CMe}2]. Quartz-crystal microbalance and X-ray photoelectron spectroscopic studies confirm that the isopropyl derivative is an improved ALD precursor versus the R=t-butyl derivative for the deposition of nitrogen free MoO3. A linear growth rate was observed at 150 °C using [MoO2{(iPrN)2CMe}2] and O3, with a reported growth rate of ∼0.065 Å per cycle, as determined by ex situ by spectroscopic ellipsometry measurements, which was slightly smaller than the growth rate determined from QCM (∼0.085 Å per cycle).154 A summary of ALD precursors used to generate thin films for Group VI and VII metals is shown in Table 6.
Precursor . | Co-reactant . | Target material . |
---|---|---|
[Cr2{tBuNCHC(tBu)(Me)O}2] | BH3(NHMe2) | Cr149 |
[Mn2{tBuNCHC(tBu)(Me)O}4] | BH3(NHMe2) | MnOx149 |
[Cu{OCHMeCH2NMe2}2]/BH3(NHMe2) | Cu7Mn3150 | |
[(η5-C5H4–Si(CH3)3)Mo(CO)2{η3-2-CH2CMeCH2}] | O3 | MnO3151 |
[MoO2(THD)2] | O3 | α-MoO3, β-MoO3, MoO(2.75–2.89)152 |
[MoO2{(tBuN)2CMe}2] | O3 | MoO2.48N0.18153 |
[WO2{(tBuN)2CMe}2] | H2O | WO3153 |
[MoO2{(iPrN)2CMe}2] | O3 | MoO3154 |
Precursor . | Co-reactant . | Target material . |
---|---|---|
[Cr2{tBuNCHC(tBu)(Me)O}2] | BH3(NHMe2) | Cr149 |
[Mn2{tBuNCHC(tBu)(Me)O}4] | BH3(NHMe2) | MnOx149 |
[Cu{OCHMeCH2NMe2}2]/BH3(NHMe2) | Cu7Mn3150 | |
[(η5-C5H4–Si(CH3)3)Mo(CO)2{η3-2-CH2CMeCH2}] | O3 | MnO3151 |
[MoO2(THD)2] | O3 | α-MoO3, β-MoO3, MoO(2.75–2.89)152 |
[MoO2{(tBuN)2CMe}2] | O3 | MoO2.48N0.18153 |
[WO2{(tBuN)2CMe}2] | H2O | WO3153 |
[MoO2{(iPrN)2CMe}2] | O3 | MoO3154 |
8 New precursor chemistry of Group VIII, IX, X and XI elements
Thin films of the first row transition metal elements, and their derivatives (i.e. oxides and sulphides) have many important and current applications; metals such Fe Co and Ni have potential application in spintronic devices155 whereas Cu is the conductor of choice for interconnects in electronic devices. In addition to the interest in metallic thin films, oxides and sulphide materials such as Fe2O3, NiO, Co3O4, Cu2O, FeS, NiS, Co9S8 and CuS, which find application in a range of semiconducting energy harvesting conversion and storage applications. In 2013 Kinsley et al. published a review of precursors, and their chemistry, for the ALD of metallic first row transition metals.10 Precursors and processes for metal oxide148 and metal sulphide156 materials deposited by ALD have also been reviewed. For this reason, this section will cover examples of new precursors for Fe, Co, Ni and Cu which have, in general, been published subsequently to these reviews.
As noted in the previous section Kalutarage et al. in 2013 reported the development of a family of α-imino alkoxide complexes of the metals Cr, Mn, Fe, Co Ni and Cu of the general form [M{tBuNCHC(tBu)(Me)O}2] (Fig. 12).149 As with the Mn complex discussed earlier the Fe complex is dimeric and displays low volatilities, whereas the Co, Ni and Co analogues are monomeric and display appreciable volatilities (sublimation is achieved between 70–80 °C at 0.05 Torr). Self-limiting growth rates of 0.07 (Fe), 0.07 (Co) and 0.09 (Ni) Å per cycle were reported for an ALD process at 180 °C using BH3(NHMe2) as the co-reagent/reductant. XPS analysis was used to confirm the deposition of metallic Fe, Co and Ni metal thin films. In the case of the Cu derivative, ALD was not explored as the precursors showed volatilities below that of comparable Cu precursors, specifically [Cu{dmap}2] and presented no obvious advantages as a prospective ALD precursor.
In 2001 Winter and co-workers reported the design and synthesis of a family of first row transition metal complexes of chromium(ii), manganese(ii), iron(ii), cobalt(ii), and nickel(ii), containing 1,4-di-tert-butyl-diaza-1,3-butadienyl (tBu2DAD) ligands (Fig. 12).157 Since then, the cobalt and nickel complexes have been utilised in a number of ALD processes. In a 3-step ALD process for the deposition of cobalt metal Winter et al. describe the utility of the cobalt complex [Co{tBu2DAD}] with formic acid and 1,4-bis(trimethylsilyl)-1,4-dihydropyrazine.158 Reaction of the Co(ii) formate intermediate with 1,4-bis(trimethylsilyl)-1,4-dihydropyrazine at 180 °C resulted in the formation of pure (>97.1% by XPS) Co films. A saturative growth rate of 0.95 Å per cycle, which is significantly higher than comparable reported processes for Co metal deposition, is reported. The as-deposited Co films are reported to be of high purity, in addition to possessing a sheet resistivity close to that reported for bulk Co metal.159,160 In an extension to this work the Co and Ni complexes, [Co{tBu2DAD}2] and [Ni{tBu2DAD}2], have both been used in a 2 component ALD process with either tBuNH2 or Et2NH as the co-reactant to form Co and Ni metal films respectively.161,162 For the Co system, growth on a range of substrates proved to be relatively uniform across the temperature range 160 to 220 °C, with a self-limiting ALD window between 170–200 °C and GPCs of 0.95 Å (Pt substrate) 0.98 Å (Cu substrate) and 0.98 Å (Ru substrate).161 In the case of the Ni precursor, saturation plots demonstrated a self-limited growth window between 180–195 °C with a GPC of 0.6 Å.162 In both cases the metallic thin films were determined by XPS to be of high purity; >98% (Co) and >97% (Ni).
The same Co precursor, [Co{tBu2DAD}2], has also been reported in the ALD of Co3O4 thin films with O2 as the co-reagent, across a temperature range of 125–300 °C. Interestingly, below 265 °C mixed phase CoOx predominated (a mixture of CoO and Co3O4), whereas above 265 °C thin films appeared to be phase pure Co3O4. Growth rate data was not included as part of the report.163
A family of N-heterocyclic carbene (NHC) stabilised Co complexes of the type [Co(CO)(NO)(NHC)(PR3)] have been reported alongside an evaluation of their thermal properties and an evaluation of their application to ALD of Co based thin films. ALD experiments on the complexes [Co(CO)(NO)(iPrIm)(PMe3)], [Co(CO)(NO)(iPrIm)2], [Co(CO)(NO)(MetBuIm)2] and [Co(CO)2(NO)(iPrIm)] (Fig. 12) were all run at 250 °C using a NH3/H2 mixed gas as co-reagent, however no details of ALD growth rates, or saturation plots were disclosed.164
Perhaps not unsurprisingly given their proliferation across the other elements of the periodic table amidinate complexes of Co have also been reported. The amidinate complex bis(N-tert-butyl-N′-ethylpropionamidinato) cobalt(ii) [Co{(tBuN)2CMe}2] has been used at 265 °C by atomic layer deposition from H2 to produce Co metal thin films.165 ALD half reactions between [Co{(tBuN)2CMe}2] and H2, on Cu substrates, have also been the focus of XPs investigations. Surface chemistries between these two reagents were evaluated by XPS. Adsorption of [Co{(tBuN)2CMe}2] proved self-limiting and the precursor was shown to be reduced readily on Cu with and without H2 co-reactant to form Co0.166
ALD of the pyrite like metal disulfides FeS2, CoS2 and NiS2 have also been reported. Deposition using the complexes [M{(tBuN)2CMe}2] (M=Fe, Co or Ni) with H2S plasma (3% in Ar) at 200 °C displayed GPCs of 1.06 (Fe), 1.29 (Co) and 1.24 (Ni) Å, with an apparently large ALD window (80–200 °C).167 Contrastingly, application of the Co system [Co{(tBuN)2CMe}2] with H2S in a thermal ALD process displayed self-limited growth behaviour at 120 °C. Saturated growth rates were determined to be 0.27 Å per cycle producing the Co rich sulphide Co9S8, as determined by XPS.168 Similar studies using [Ni{(tBuN)2CMe}2] and H2S again in a thermal ALD process, to produce NiSx Films, have also been reported.169,170 In one instance the reaction was also monitored by QCM studies with a GPC of 9.3 ng cm−2 reported.169
The new cobalt complex [Co{DMOCHCOCF3}2] has been prepared by the reaction of cobalt(ii) acetate tetrahydrate with two equivalents of deprotonated ligand 1-(dimethyl-1,3-oxazol-2-yl)-3,3,3-trifluorprop-1-en-2-ol (DMOCHCOHCF3). This new precursor has been reported to show a high degree of thermal stability. Optimised ALD processes and growth studies were performed on Si (100) and carbon/TiO2 nanofibers substrates, at 145 °C with ozone (O3) as the co reagent. Studies revealed a well-defined temperature range with constant growth per cycles of 0.2 Å per cycle between 150 °C and 200 °C of phase pure Co3O4.171
In a radical departure from complicated and often very air sensitive ALD precursors Väyrynen et al. have reported the simple diamine adduct [CoCl2{κ2-Me2NCH2CH2NMe2] alongside a full ALD study of the precursor and its application to the fabrication of CoO thin films. Using water as a precursor the process was investigated over a 225–300 °C range, below the limit of the precursor's thermal stability. Saturation of the film growth with respect to both precursors was determined. At 275 °C high purity CoO was deposited at a growth rate of 0.2 Å per cycle regardless of Co precursor pulse length.172 The implications of this new complex could be profound and is the first of what will be a much larger number of simple adduct systems to be assessed in the future.
As noted earlier in this section Cu has an assured place as the conductor of choice for interconnects in electronic devices.173 As such, it has long been a material which has garnered attention as have precursors for the deposition of high purity Cu metal and the selective deposition of semiconducting oxide materials such as Cu2O and CuO. Since the development of Cu(i) amidinates as precursors to copper metals by Barry and Gordon174 only a small number of new precursors have been developed including the amino-alkoxide complex [Cu{dmap}2].175,176 The NHC-copper(i) amide system [{iPr2Im}Cu{N(SiMe3)}] has been shown to deposit Cu metal, with growth rate per cycle of 0.2 Å, at 225 °C in a plasma enhanced ALD process using H2 plasma as the co-reagent.177
Atomic layer deposition has long been an attractive method by which thin films for advanced technological applications such as microelectronics and nanotechnology can be deposited. One material group in ALD that has matured in past 15 years and proven to be of importance to a wide range of technological applications are the noble metals, specifically the elements Ru, Os, Rh, Ir, Pt, Pd, Ag and Au. While these metals are known to be good conductors of both electricity and heat, they are perhaps even better known because of their application to catalysis and because of their resistance to corrosion and oxidation.178
A summary of ALD precursors used to generate thin films for Group VIII, IX, X and XI metals is shown in Table 7. Whilst the noble metals have attracted attention, their oxides are not often considered. However, for materials such as RuO2 and IrO2, which are conductive and biocompatible and thus are possible candidate materials for biological applications and implantable devices, in addition to being structurally compatible electrode materials with high-κ dielectrics such as TiO2 and SrTiO3. In conjunction with the review by Miikkulainen et al. which provides an overview of precursors and general ALD trends across the periodic table,148 a more recent review by Hämäläinen et al. provides comprehensive coverage of the topic across the noble metals.178 The review by Hämäläinen et al. is comprehensive, exploring both noble metal containing precursors and their thermal ALD to produce noble metals and noble metal oxides. The review covers reaction mechanisms in various types of processes as well as specific issues regarding nucleation, in addition to tabulating and comparing the deposition temperatures, film growth rates, purities as well as evaporation temperatures for noble metal precursors. A review highlighting the specific chemistries of ruthenium β-diketonates and their application to ALD has also been published.182 As a result we shall only describe here the development of new precursors which have appeared in the literature since the publication of these reviews.
Precursor . | Co-reactant . | Target material . |
---|---|---|
[Fe{tBuNCHC(tBu)(Me)O}2] | BH3(NHMe2) | Fe149 |
Co149 | ||
Ni149 | ||
[Co{tBuNCHC(tBu)(Me)O}2] | ||
[Ni{tBuNCHC(tBu)(Me)O}2] | ||
[Co{tBu2DAD}] | HCO2H/1,4-bis(trimethylsilyl)-1,4-dihydropyrazine | Co158 |
[Co{tBu2DAD}] | tBuNH2 or Et2NH | Co161 |
[Ni{tBu2DAD}2] | tBuNH2 | Ni162 |
[Co{tBu2DAD}2] | O2 | CoOx and Co3O4163 |
[Co(CO)(NO)(iPrIm)(PMe3)] [Co(CO)(NO)(iPrIm)2] [Co(CO)(NO)(MetBuIm)2] [Co(CO)2(NO)(iPrIm)] | NH3/H2 | Co164 |
[Co{(tBuN)2CMe}2] | H2 | Co165,166 |
[M{(tBuN)2CMe}2] (M=Fe, Co, Ni) | H2S-plasma | FeS2, CoS2 & NiS2167 |
[Co{(tBuN)2CMe}2] | H2S | Co9S8168 |
[Ni{(tBuN)2CMe}2] | H2S | NiSx169,170 |
[Co{DMOCHCOCF3}2] | O3 | Co3O4171 |
[{κ2-Me2NCH2CH2NMe2}CoCl2] | H2O | CoO172 |
[{iPr2Im}Cu{N(SiMe3)}] | H2-plasma | Cu177 |
[(η4-C6H8Et)Ru(η6-C6H5Et)] | O2 | Ru179 |
[(Et3P)Au{N(SiMe3)2}] | BH3(NHMe2) | Au180 |
[Me2Au{S2CNEt2}] | O3 | Au181 |
Precursor . | Co-reactant . | Target material . |
---|---|---|
[Fe{tBuNCHC(tBu)(Me)O}2] | BH3(NHMe2) | Fe149 |
Co149 | ||
Ni149 | ||
[Co{tBuNCHC(tBu)(Me)O}2] | ||
[Ni{tBuNCHC(tBu)(Me)O}2] | ||
[Co{tBu2DAD}] | HCO2H/1,4-bis(trimethylsilyl)-1,4-dihydropyrazine | Co158 |
[Co{tBu2DAD}] | tBuNH2 or Et2NH | Co161 |
[Ni{tBu2DAD}2] | tBuNH2 | Ni162 |
[Co{tBu2DAD}2] | O2 | CoOx and Co3O4163 |
[Co(CO)(NO)(iPrIm)(PMe3)] [Co(CO)(NO)(iPrIm)2] [Co(CO)(NO)(MetBuIm)2] [Co(CO)2(NO)(iPrIm)] | NH3/H2 | Co164 |
[Co{(tBuN)2CMe}2] | H2 | Co165,166 |
[M{(tBuN)2CMe}2] (M=Fe, Co, Ni) | H2S-plasma | FeS2, CoS2 & NiS2167 |
[Co{(tBuN)2CMe}2] | H2S | Co9S8168 |
[Ni{(tBuN)2CMe}2] | H2S | NiSx169,170 |
[Co{DMOCHCOCF3}2] | O3 | Co3O4171 |
[{κ2-Me2NCH2CH2NMe2}CoCl2] | H2O | CoO172 |
[{iPr2Im}Cu{N(SiMe3)}] | H2-plasma | Cu177 |
[(η4-C6H8Et)Ru(η6-C6H5Et)] | O2 | Ru179 |
[(Et3P)Au{N(SiMe3)2}] | BH3(NHMe2) | Au180 |
[Me2Au{S2CNEt2}] | O3 | Au181 |
Popovici et al. have described the ALD of Ru thin films starting from the new precursor (ethylbenzyl)(1-ethyl-1,4-cyclohexadienyl)Ru(0) (C16H22Ru) in combination with molecular oxygen (O2) as co-reagent on both SiO2 and TiN starting surfaces in the temperature range 225 and 325 °C. The study shows that the Ru growth behaviour is strongly dependant on the starting surface and explores how this can influence the film properties and growth kinetics.179
Other developments in the area of precious metal precursor are centred on the development of a small number of new Au-precursors. In 2016 Mäkela and colleagues described their evaluation of a family of seven new gold precursors for application in the formation of gold containing thin films. From this shortlist of perspective precursors the liquid precursor [(Et3P)Au{N(SiMe3)2}] was selected for investigation with potential reducing agents and hydrogen sulphide. The study describes the best in show results with respect to growth rate and film properties, obtained from [(Et3P)Au{N(SiMe3)2}] and BH3(NHMe2). Unfortunately self-limiting growth profiles, characteristic of true ALD could not be confirmed; however the process did produce polycrystalline Au films of relatively high purity.180 In studies by the same group of researchers another family of Au(iii)dithiocarbamate complexes were screened for potential ALD application. Of the systems investigated one complex, namely [Me2Au{S2CNEt2}] showed optimal thermal behaviour, being volatile while displaying thermal stability. ALD growth experiments with O3 at temperatures between 120–250 °C were investigated. Self-limited growth was confirmed at 180 °C with a rate of 0.9 Å per cycle growth of Au. The report describes the as deposited thin films of Au as uniform polycrystalline and highly conductive, with only trace amount of impurities (as determined by XPS).181
9 New precursor chemistry of Group XII elements
Perhaps even more so than any other group in the periodic table, the ALD of thin films containing Group XII elements, and in particular zinc containing materials is dominated by the alkyl metal reagents, such as diethyl zinc (DEZ) and to a lesser extent, dimethyl zinc (DMZ).148,183 Whilst other precursors such as [Zn(OAc)2], ZnCl2, [ZnMe(OiPr)] and [Zn{THD}2] have been explored in very specific ALD processes, in the past 20 years there is only one example of a new and novel ALD precursor for zinc based thin films.
In 2016 Devi and co-workers reported the design, synthesis, characterisation and viability as precursors for atomic layer deposition (ALD) of ZnO, of a family of zinc ketoiminates. Of the seven complexes explored one compound was outstanding with respect to its thermal properties, [Zn{OC(Me)CHC(Me)NCH2CH2OEt}2] ([Zn(eeki)2]) (Fig. 13). Deposition of ZnO thin films by ALD was studied using [Zn(eeki)2] as the metal and water as the oxygen sources, respectively, over a temperature range of 150–300 °C. An ALD window was observed from 175–300 °C with an average growth rate of ∼1.3 Å per cycle. Growth rates were shown to be minimal at low temperatures highlighting that the activation energy (i.e. reaction with either the surface or H2O) of the process was not sufficient to form a film as it was reactivity limited. The authors suggest that the broad ALD window provides a means of lengthening the window for stable high temperature ALD processes such as MLD processes as it widens the precursor cross-over window for organic molecules as well as dopants. The crystallographic orientations of the deposited thin film layers were found to be temperature dependent and contrary to the literature observation of c-axis dominant ZnO thin films grown using diethyl zinc. UV–Vis measurements revealed that the transparency is high in the visible range with a band gap of 3.29 eV.184
For cadmium materials the precursor library is even more meagre, with dimethyl cadmium (DMCd) being the only reported ALD precursor. To the best of our knowledge there have been no developments in the arena of cadmium precursors in the past two decades. In the case of mercury, to the best of our knowledge no mercury based precursor has been reported to date.
10 New precursor chemistry of Group XIII elements
The atomic layer deposition of group 13 materials is perhaps the most established and widely studied within the field, owing largely to the high reactivity of the simple boron-group alkyl compounds. The deposition of Al2O3 is the de facto “gold standard” of ALD processes, with a vast number of applications across multiple fields, whilst the subsequent use of (III–V) materials such as aluminium, gallium and indium nitrides, phosphides and arsenides in electronic applications has seen a wealth of research in the area over the last few decades. A description of the widespread use of simple alkyl, chloride, alkoxide and amide precursors is neither novel or within the scope of this overview, however examples of simple processes and precursors for many boron-group materials can be found within a 2013 review of inorganic ALD films by Miikkulainen et al.148 Some examples of ALD precursors used to generate Group XIII materials is shown in Fig. 14 and summarised in Table 8.
Precursor . | Co-reactant . | Target material . |
---|---|---|
B2F3 | H2O/TMA | B2O3187 |
[Al(mmp)3] | H2O | Al2O3189 |
[Al(NiPr)2{dmp}] | H2O | Al2O3190 |
[AlMe2{dmp}] | O3 | Al2O3190 |
[Ga{THD})3] | O3 | Ga2O3191 |
[GaMe2(OiPr)] | H2O | Ga2O3193 |
[InMe2(NMe2)] | H2O | In2O3194 |
[InMe2N(SiMe2)] | H2O | In2O3195 |
[In{(iPrN)2C(NR2)}3] (R=Me, Et) | H2O | In2O3199 |
[InMe2{dmp}] | H2O | In2O3196 |
[InMe2{edpa}] | O3, H2O | In2O3197,198 |
[In(dmamp)3] | O3 | In2O3200 |
Precursor . | Co-reactant . | Target material . |
---|---|---|
B2F3 | H2O/TMA | B2O3187 |
[Al(mmp)3] | H2O | Al2O3189 |
[Al(NiPr)2{dmp}] | H2O | Al2O3190 |
[AlMe2{dmp}] | O3 | Al2O3190 |
[Ga{THD})3] | O3 | Ga2O3191 |
[GaMe2(OiPr)] | H2O | Ga2O3193 |
[InMe2(NMe2)] | H2O | In2O3194 |
[InMe2N(SiMe2)] | H2O | In2O3195 |
[In{(iPrN)2C(NR2)}3] (R=Me, Et) | H2O | In2O3199 |
[InMe2{dmp}] | H2O | In2O3196 |
[InMe2{edpa}] | O3, H2O | In2O3197,198 |
[In(dmamp)3] | O3 | In2O3200 |
Despite the unparalleled simplicity, reactivity and efficiency of established boron-group precursors, a number of avenues of research are being directed towards the development of less reactive, more exclusive and importantly less toxic reagents for group 13 materials. This is particularly prevalent in the deposition of boron-containing films, with a standard process revolving around the highly toxic and flammable diborane.185 To this end, alkoxide compounds such as triisopropyl borate, [B(OiPr)3], have been shown to be effective in the deposition of boron as a film dopant, and are seemingly less reactive and toxic than their alkyl counterparts.185,186 Recently, B2F4 was shown for the first time to be used in the deposition of bismuth aluminate films with H2O or reducing agent Si2H6 (for elemental B deposition), alongside trimethylalumimium (TMA), as co-reagents. Film growth was found to quickly decline after substrate saturation with processes consisting of solely B2F4/H2O and B2F4/Si2H6. However, with alternate TMA/H2O pulses, repeatable growth of bismuth aluminate was obtained.187 Single-source boron containing precursors have also been used by Saly et al. to deposit calcium, strontium and barium borate with [M{Tpb}2] (M=Ca Sr & Ba) and water, as highlighted in previous alkaline earth discussions.71–73
In similar attempts to temper the reactivity of aluminium precursors, a number of heteroleptic species such as non-pyrophoric dimethylaluminium isopropoxide [Me2AlOiPr] have been developed and utilised in aluminium deposition, but are not recent enough developments to warrant more detailed discussion and can be found mentioned in a number of publications and the review described previously.148,188 The alkoxy ether complex aluminium tris(1-methoxy-2-methyl-2-propoxide), or [Al{mmp}3], (Fig. 13) was first reported to deposit Al2O3 films via solution injected ALD with H2O at growth rates of ∼0.6–7 Å per cycle (200–350 °C) in 2005 by Min et al.,189 and benefits from the enhanced stability of a chelating oxygen pendant moiety on each alkoxide ligand, though it is noted that the added bulk with respect to conventional monodentate ligand systems is the likely cause of the reduced growth per cycle cf. more established precursors.
A recent study by Mai et al. (2017),190 synthesised three prospective non-pyrophoric aluminium precursors based on an aluminium centre with two varied, singly coordinated amines in addition to a stabilising donor functionalised alkyl ligand, 3-(dimethylamino)-propyl, or {dmp}. The three complexes, of the formula [Al(NR2)2{dmp}], where R=Me, Et and iPr, exhibited good thermal stability and high volatility. In addition to the three amino precursors, a dimethyl aluminium derivative with {dmp}, [AlMe2{dmp}] was also investigated. (Fig. 13). Of the synthesised precursors, [Al(NiPr)2{dmp}] was shown to deposit Al2O3 films with H2O (100–180 °C, 1.0 Å per cycle), whilst the fully alkylated compound [AlMe2{dmp}] gave a GPC of 0.7 Å in a low-temperature (60–180 °C) oxygen plasma enhanced deposition.190
Recent investigations into novel gallium precursor chemistries are more limited than their aluminium analogues, though a 2014 study by Ramachandran and co-workers191 trialled the application of [Ga{THD}2] in the formation of Ga2O3 films. Although the precursor was shown to have enhanced volatility and thermal stability over its parent acetylacetonate, [Ga(acac)3] – previously used in gallium ALD192 – growth was only observed with an O3 oxidant and exhibited a very low growth per cycle of 0.1 Å within a temperature window of 100–400 °C.191 In a bid to decrease the pyrophoric nature of gallium alkyl precursors, an investigation by Lee et al.193 sought to apply the volatile, non-pyrophoric and reactive dimethylgallium isopropoxide (DMGIP), [GaMe2(OiPr)], with an H2O co-reagent to the ALD of gallium oxide. The process yielded a narrow temperature window (280–300 °C) with a low growth rate of ca. 0.3 Å per cycle. Interestingly, when compared to the analogous aluminium compound DMAIP, (Me2Al(OiPr)), mentioned briefly previously, this growth rate is exceedingly low taking into consideration that fact that the growth rate for the aluminium species is nearly comparable to the standard pyrophoric trimethylaluminium process.188,193
There is relatively more abundance within novel indium precursor development than there can be found within that of the earlier members of the group, though a wide variety of existing processes are covered in the 2013 review by Miikkulainen et al.148 Research of a similar vein to that taking place within aluminium and gallium ALD is being undertaken, with the varying of simple amido and alkyl substituents and their effect on deposition properties explored. Such a study utilised dimethylamino dimethyl indium [InMe2(NMe2)] and H2O, with deposition of In2O3 occurring between 250 °C and 400 °C, with a GPC of 0.61–65 Å per cycle between 300–350 °C.194 A similar investigation by Maeng et al.195 demonstrated growth of indium oxide films between 175 °C and 250 °C (GPC, 0.7 Å), with the similar aminosilyl precursor [InMe2{N(SiMe3)2}]. Further to this work, the volatile liquid [InMe2{dmp}], analogous to the aluminium precursor previously discussed, has been trialled with H2O in depositions found to afford indium oxide at 275 °C with a significant GPC of 0.6 Å.196
Investigations into novel precursors by Kim et al.197 in 2016 saw the low temperature (90–180 °C) deposition of In2O3 with O3 at a rate of ca. 0.5 Å per cycle with the precursor dimethyl(N-ethoxy-2,2-dimethylcarboxylicpropanamide)indium, [InMe2{edpa}] (Fig. 13). A subsequent study by Agbenyeke and co-workers198 saw deposition of the same material using H2O at 200–300 °C (∼0.8 Å per cycle).
Two novel indium guanidinates of the form [In{(iPrN)2C(NR2)}3] (R=Me, Et), were described by Gebhard et al.199 (Fig. 13) and were applied to the ALD of indium oxide with H2O. The monomeric compounds were found to deposit In2O3 at a growth rate of 0.4–0.5 Å per cycle at 230–300 °C. A later study into novel indium precursor systems yielded 1-dimethylamino-2-methyl-2-propoxy indium, or [In{dmamp}3], which, when reacted with O3 at temperatures between 175–200 °C was found to deposit carbon-negligible In2O3 at a GPC of 0.27 Å.200
11 New precursor chemistry of Group XIV elements
As has been highlighted previously, carbon incorporation by design in the form of carbonates has long been achieved with the use of CO2 or by plasmolysis or ozonolysis of ligands. Another well-established route towards incorporation of carbon into thin films is via the reaction of alkylated ligand systems with H2 plasma.201 The following overview will attempt to avoid these processes and focus on recent developments in the specific atomic layer deposition of carbon. Examples of ALD precursors used to generate Group XIV materials are summarised in Table 9.
Precursor . | Co-reactant . | Target material . |
---|---|---|
DSBAS | O3/O2 (plasma) | SiO2212,214 |
[H2Si(NRR′)2] | O3/O2 (plasma) | SiO2212,214 |
(R=tBu, R′=H/R=Et, R=Et) | ||
TEOS | H2O/NH3 | SiO2210 |
tBOS | H2O, TMA, N2 (plasma) | SiO2,220–222 silicon nitride217 |
TPS | H2O/TMA | SiO2223 |
Cyclic azasilanes | O3 | SiO2227 |
[Ge{HMDS}2] | MeOH/Te(SiMe3)2 | GeTe228 |
[Ge(i-C4H9)4] | H2 (plasma) | Ge2Sb3Te5229 |
[Ge{CAMD}2] | H2S | GeS230 |
[Ge(dpp-BIAN)] | O3 | GeO2231 |
[Ge{RNCH2CH2NR}2(NMe2)2], R=tBu,iPr | O3 | GeO2232 |
[Sn{acac}2] | O3, O3/Ti(OiPr)4 | SnOx,233 SnxTi1−xOy234 |
[Sn{CAMD}] | H2O2, NO, H2S | SnO2,235,236 SnS230 |
[Sn{(iPrN)2CMe}2] | H2S | SnS237 |
[Sn{dmamp}2] | H2O, O3 | SnO,238 SnO2239 |
[Sn{HMDS}2] | H2O, O3 | SnO,240 SnO2240 |
[Pb{dmamp}2] | H2O | PbOx241,242 |
[Pb{THD}2] | Te(SiMe3)2 | PbTe243 |
[Pb{THD}2] | Se(SiEt3)2 | PbSe244 |
[Pb{THD}2] | H2S | PbS245,246 |
[Pb{THD}2] | O3 | PbO2247 |
Precursor . | Co-reactant . | Target material . |
---|---|---|
DSBAS | O3/O2 (plasma) | SiO2212,214 |
[H2Si(NRR′)2] | O3/O2 (plasma) | SiO2212,214 |
(R=tBu, R′=H/R=Et, R=Et) | ||
TEOS | H2O/NH3 | SiO2210 |
tBOS | H2O, TMA, N2 (plasma) | SiO2,220–222 silicon nitride217 |
TPS | H2O/TMA | SiO2223 |
Cyclic azasilanes | O3 | SiO2227 |
[Ge{HMDS}2] | MeOH/Te(SiMe3)2 | GeTe228 |
[Ge(i-C4H9)4] | H2 (plasma) | Ge2Sb3Te5229 |
[Ge{CAMD}2] | H2S | GeS230 |
[Ge(dpp-BIAN)] | O3 | GeO2231 |
[Ge{RNCH2CH2NR}2(NMe2)2], R=tBu,iPr | O3 | GeO2232 |
[Sn{acac}2] | O3, O3/Ti(OiPr)4 | SnOx,233 SnxTi1−xOy234 |
[Sn{CAMD}] | H2O2, NO, H2S | SnO2,235,236 SnS230 |
[Sn{(iPrN)2CMe}2] | H2S | SnS237 |
[Sn{dmamp}2] | H2O, O3 | SnO,238 SnO2239 |
[Sn{HMDS}2] | H2O, O3 | SnO,240 SnO2240 |
[Pb{dmamp}2] | H2O | PbOx241,242 |
[Pb{THD}2] | Te(SiMe3)2 | PbTe243 |
[Pb{THD}2] | Se(SiEt3)2 | PbSe244 |
[Pb{THD}2] | H2S | PbS245,246 |
[Pb{THD}2] | O3 | PbO2247 |
Carbon precursors, simple silanes, aminosilanes and chlorosilanes, and variants of silanol have not been included in this table.
The selection of appropriate methodology for a novel carbon containing precursor with sole aim of introducing carbon into a film is desirable despite the well-established nature of carbon incorporation by other means. This is of particular interest in the deposition of silicon carbide (SiC), an important material for semiconductor applications, with a low wet-etch rate and higher operating temperature than SiN and SiO2. Current thermal CVD processes for crystalline material require high temperatures >1000–1500 °C, only somewhat mitigated by plasma enhanced CVD, however deposition of amorphous material is possible at temperatures as low as 200 °C.202,203 To this end, computational studies have attempted to direct research towards this arena, highlighting ethylene (C2H4), carbon tetrachloride (CCl4) and trichloromethane (CHCl3) as most promising for use alongside silicon precursors silane (SiH4), disilane (Si2H6) and monochlorosilane (SiH3Cl), though it is acknowledged that the processes occurring at the surface are not likely to be self-limiting.203,204
As, once again, the problematic topic of self-limiting ALD reactions is probed, the most notable novel deposition focussed on carbon films is that of graphene, which was reported by Zhang et al. in 2014.205 As seen in conventional graphene CVD, copper foil was used as a substrate, with benzene (C6H6) as a pre-aromaticised carbon source. Depositions were carried out with sequential pulses of H2/Ar plasma at 400 °C. 10 ALD cycles were found to be sufficient to afford the formation of a monolayer graphene sheet of reasonable size.
It is perhaps unsurprising that, as the stalwart of the semiconductor industry, a number of established ALD processes exist towards the fabrication of silicon containing thin-films.1 A discussion of simple, conventional precursors such as silanes, amidosilanes and chlorosilanes206–208 is as such beyond the scope of this overview, however a thorough review by Meng et al.209 (2016) of SiN deposition processes provides a good baseline. The development of non-chlorinated precursors is desirable in the first instance, not least due to the corrosive by-products formed.210 A study into the impact of aminosilane precursor structure on SiO2 ALD was carried out by O'Neill et al.,211 which suggested that a higher degree of amino substitution of silanes resulted in compromised growth rates and ligand incorporation into films, a finding corroborated by Mallikarjunan and co-workers exploring the efficacy of the monoaminosilane, di-sec-butylaminosilane (H3SiNsBu2/DSBAS) with the diaminosilanes bis(tert-butylaminosilane) (BTBAS) and bis(diethylaminosilane), of the formula SiH2(NRR′)2 (R=tBu, R′=H and R=Et, R=Et).212,213 All three variants described have found use in SiO2 ALD, with the high growth rates (∼1.6 Å per cycle, O3) at low (150–300 °C) and moderate (250–300 °C) temperatures respectively for di-sec-butylaminosilane (DSBAS) and bis(tert-butylamino)silane (BTBAS) proving to be of significant interest.212,214 A number of other amiosilanes have been reported and a computational study by Huang et al.215 can be found on the optimisation of precursor design in this field. Other silicon precursors studied for SiO2/silcon nitride deposition include, among others, the derivatives of silanol or silyl ethers, such as tetraethoxysilane (TEOS), tri(tert-butoxyl)silanol (tBOS), and tris(tert-pentoxy)silanol (TPS), which have been used with H2O or N2/NH3 (plasma) as traditional co-reactants,216–219 or with trimethylaluminium in a catalysed process known as “rapid ALD”, where multiple layers of material are deposited with each cycle.210,220–223 Interestingly, despite the multilayer growth observed within “rapid ALD”, conformity of high aspect ratio substrates was conserved. These catalysed processes, undertaken in the presence of Lewis acid (TMA)221,224 or Lewis base (ammonia/pyridine)225,226 catalysts serve to allow the low temperature deposition of silicon containing films, without the need for aggressive plasma or thermal conditions.215
Perhaps the most inventive silicon-based precursor development in recent years has been the design and synthesis of a range of cyclic azasilanes by Ju and Strandwitz in 2016227 (Fig. 15). A series of four ring-strained systems, with a variety of substituents, displayed high reactivity towards SiO2 substrates with an O3 co-reagent. The study offers insights into the reactivity of various silicon-based functional groups, with Si–OMe moieties displaying shorter O3 saturation pulses than their alkyl counterparts.
As group 14 is descended, the prevalence of the 2+ oxidation state increases and due to the demand for both Ge(ii) and Ge(iv) thin films in electronics, processes for the deposition of both valences are desirable. Of these Ge(ii) and Ge(iv) materials, many of the chalcogenides are of importance, as are ternary germanium materials. Germanium telluride (GeTe) and germanium-antimony-telluride (Ge2Sb3Te5) are of great interest as phase change memory materials and as such a number of ALD routes towards them have been developed. A recent review by Harmgarth et al.248 and other publications228,249–251 cover these precursors in greater depth than the scope of this overview, with many of the precursors discussed based on simple germanium halides, alkoxides and amides. There is a relative scarcity of novel precursor development for the atomic layer deposition of germanium chalcogenides, particularly within divalent species, though work in this area is expanding.
Germanium alkoxide precursors have been investigated extensively by Eom et al.252,253 and others254 for the formation of germanium-antimony-tellurides, with the general conclusion that, though the process is not fully understood, interaction between the germanium precursors and substrate surface was limited as long purge times resulted in little to no deposition, and the preferential formation of Ge(iv)Te2 as opposed to Ge(ii)Te was a significant hindrance. In an attempt to circumvent this, a combined ALD/CVD process was developed using [Ge(i-C4H9)4] and a reductive H2 plasma which has been shown to be effective in the deposition of Ge2Sb3Te5.229 Another recent method towards GeTe films utilises the Ge(ii) amide precursor [Ge{HMDS}2] and [Te(SiMe3)2], which when pulsed alongside methanol to create reactive intermediates in the gas-phase, afforded GeTe thin films. It was found to be necessary to induce gas-phase reactions with methanol due to the lack of reactivity between [Ge{HMDS}2] and [Te(SiMe3)2].228
An interesting precursor for the deposition of both SnS and GeS was developed by Kim and co-workers in 2014, which consists of a single dianionic, racemic cyclic amide ligand with sufficient steric bulk to singly satisfy the coordination sphere of Sn(ii) and Ge(ii). The cyclic amide complex, (Fig. 15) [Ge{(ButN)CH(CH3)CH(CH3)(NBut)}2], [Ge{CAMD}2], was found to deposit with a growth rate of ca. 0.21–0.28 Å per cycle between 50–100 °C after which a decrease in growth was observed.230
GeO2 was deposited with a divalent, two-coordinate precursor by Perego et al.,231 1,2-bis[(2,6-diisopropylphenyl)imino]acenapthalene germanium(ii), [Ge{dpp-BIAN}] (Fig. 15), and O3 at temperatures between 185 °C and 225 °C (∼0.5 Å per cycle).231,255 Further study into novel systems for the deposition of GeO2 focussed on tetravalent systems, through the somewhat simpler Ge(NMe2)-chelates (Fig. 15), where coordinative saturation was achieved with N,N′-di-tert-butyl-ethylenediamine or its isopropyl analogue, with growth rates ca. 0.40 Å and 0.31 Å per cycle respectively between 200–330 °C.232
A wide variety of tin(ii) and tin(iv) materials are of increasing interest, with the tin chalcogenides the primary focus of investigations. A 2015 review by Nazarov et al.,256 in addition to a review by Miikkulainen148 and co-workers, covers many precursors for the deposition of SnO2 and SnS. Such common and well-established precursors include simple Sn(iv) alkyls, amides and chlorides. With the intermediate position of tin within the carbon group, distinction must be made between the various valences of tin(ii) and (iv) materials, with oxidative control of respective states most difficult to establish in oxygen-rich and highly oxidising environments. This is particularly pertinent in the production of tin oxide materials, with the pure SnO phase considerably less stable than the SnO2 and SnOx phases.
Testament to this fact, a number of divalent tin precursors have been trialled in the deposition of tin oxide films, yet have failed to retain the divalent nature of tin within the material. An example of this can be seen with the β-diketonate precursor [Sn{acac}2], previously used in the production of SnS.233,234,257 Sn(acac)2 was found to deposit films of SnOx within a temperature window of 175–300 °C (∼1 Å per cycle), whilst was used recently by Chang et al.234 to obtain good stoichiometric control over the production of SnxTi1−xOy films with titanium isopropoxide and ozone. The same precursor was used with H2S in further work on SnS deposition to obtain phase control over cubic and orthorhombic SnS films via ALD.258
The somewhat subtle nature of tin oxide deposition with respect to oxidation is again exemplified with the N-heterocyclic stannylene precursor investigated by Kim et al.,230,235 [Sn(ii){(ButN)CH(CH3)CH(CH3)(NBut)}], the germanium analogue of which was also described. The stannylene was reacted with H2O2 to yield SnO2 films at temperature between 50 °C and 150 °C (∼1.8 Å per cycle),235 whilst a subsequent group publication sought to avoid the use of unstable H2O2, which was thought to decompose within high aspect-ratio substrates at higher temperatures, by using NO as an oxidant, achieving SnO2 films at a rate of ca. 1.4 Å per cycle at higher temperatures of 200–250 °C.236 The same precursor was shown in a later study to deposit SnS with an H2S co-reagent.230 Work by the same group investigated the application of a second Sn(ii) precursor, the acetamidinato complex [Sn{(iPrN)2CMe}2], which shows growth of single-phase SnS at 100–200 °C with a growth rates of ∼0.9 Å per cycle when used with H2S.237
Oxidative control over the deposition of tin(ii) oxide films was achieved by Han and co-workers238 using the Sn(ii) precursor bis(1-dimethylamino-2-methyl-2-propoxy)tin [Sn{dmamp}2]. SnO films grown with H2O were deposited between 90 °C and 210 °C with phase pure, crystalline films grown at 150 °C and above with a growth per cycle of 0.18 Å. Contrastingly, when ozone was used as co-reagent in a subsequent publication,239 SnO2 films were deposited at 100–230 °C exhibiting increasing growth rates with temperature (0.18–0.42 Å per cycle). Further work into the oxidative control of SnO deposition was undertaken by Tupala et al.,240 in a study that applied the long-established stannylene [Sn{HMDS}2] to atomic layer deposition processes with both H2O and O3. Films of SnO were grown with H2O displaying a variable growth rate (0.05–0.18 Å per cycle) at temperatures of 100–250 °C, whilst films grown with ozone showed lower, even more variable growth rates (0.05–0.11 Å per cycle) at temperatures of 80–200 °C. Importantly, films grown with water were shown to consist of crystalline SnO between 125 °C and 175 °C but contained significant quantities of SnO2, silicon and nitrogen, and films grown with O3 were found to consist of SnO2 and SiO2, with similar nitrogen contamination.240
Despite the use of lead in thin film materials becoming increasingly limited, the manifold oxides (simplified in this study to PbOx) and sulfides of lead(ii), in addition to ternary structures such as PbTiO3, have found use in a wide range of optoelectronic devices, solar and sensing applications.245,246,259–261 The heavier lead chalcogenides such as PbSe and PbTe are utilised in thermoelectric, photo optic and other semiconductor applications.243,262 As with other group 14 elements, simple alkyl and alkoxide complexes of lead and their long-standing uses in deposition are omitted.245,261,263 Very few recent advances within lead precursor chemistry pertinent to this overview exist, though a brief description of some relevant investigations is given herewith.
PbTe was deposited for the first time in 2014 by Zhang and co-workers243 via the atomic layer deposition of Pb(THD)2 and Te(SiMe3)2 at growth rates of 0.25 Å per cycle between 150–210 °C, whilst nanolaminates of PbSe/Te were deposited by the same author with the addition of Se(SiEt3)2.244 The Pb(THD)2 complex has also been shown to deposit films of PbS and PbO2 with H2S and O3 as oxidants respectively.245–247 Further to this, aminoalkoxide Pb(DMAMP)2, analogous with the Sn(ii) chelate of identical composition, has been successfully used with H2O on a number of occasions to deposit lead oxides.241,242
12 New precursor chemistry of Group XV elements
A number of nitrogen and phosphorous containing films have previously been touched upon in the discussion of advances in the precursor chemistry of lithium, group 13 elements and silicon, and the reader is directed towards more relevant literature in the case of further interest. The deposition of the metalloids and metallic pnictogens arsenic, antimony and bismuth materials are the primary focus of the following subsection, though little significant precursor development has taken place in recent years, in part due to the high efficacy of current technologies.
The evolution and widespread demand within the semiconductor industry for III–V (13–15) materials has seen the development of a range of similar, well-established ALD routes based around alkylated precursors, as mentioned in the group 13 discussions. Arsenide films are all routinely deposited through use of simple metalloid-alkyls, amides or hydrides, such as tBuAsH2,264 EtAsH2,265 As(NMe2)3,266 and, most commonly AsH3,267–269 whilst antimonide films often utilise antimony alkylsilyls such as Sb(SiEt3)3.270 Materials such as Sb2O3 and Sb2S3 are also of interest for battery and photovoltaic applications,271,272 with simple precursors such as Sb(NMe2)3 reacting with O3 or H2S to afford the desired chalcogenide film.273 Previous mention has been made of germanium antimony telluride as a phase-change material, with antimony chlorides and alkoxides being used to this end.228,274
A number of more varied precursors for the deposition of bismuth containing thin films have been investigated with a view to the deposition of a wide variety of materials, with Bi2O3 and Bi2S3 desirable due to tunable electronic and thermoelectric properties and uses in supercapacitors and gas sensors.275–277 More complex materials such as bismuth silicate, bismuth ferrate and bismuth titanates are also of interest for a range of electronic devices.277,278
The use of triphenylbismuth [Bi(C6H5)3] is relatively well established within CVD and ALD, readily oxidised to Bi2Ti2O7 with titanium isopropoxide and H2O, and Bi2O3 with O3 (∼0.23 Å per cycle, 250–320 °C).277–279 An interesting single-source precursor Bi(CH2SiMe3)3 for bismuth silicate was developed by Harjuoja et al.,280,281 acting as both bismuth and silicon source when reacted with O3. Growth rates of 0.4 Å per cycle between 250 °C and 350 °C were observed. For a number of years the chelating [Bi{mmp}3] system, tris(1-methoxy-2-methyl-2-propoxy)bismuth (Fig. 16), has been used to deposit both Bi2Ti2O7 and Bi1−x−yTixSiyOz (BTSO), with reactant pulses of [Ti(OiPr)4] and H2O,282 and [Ti(OiPr)4] or [Ti{mmp}4], [Si(OEt)4] and H2O283,284 or O3285 respectively.
As is the case for many other ALD processes, there are reports of the use of the β-diketonate ligand THD with H2O for bismuth thin film deposition, but the process exhibited low growth rates of ca. 0.1 Å per cycle between 270 °C and 300 °C.286 The same precursor, with the considerably more acidic H2S as a co-reagent, afforded polycrystalline p-type Bi2S3 at a GPC of 0.35 Å between temperatures of 175 °C and 250 °C.287 A 2004 study by Vehkamäki and co-workers288 sought to investigate the utility of a range of bismuth amide and thioamidate complexes; homoleptic silylamides [Bi{N(SiMe3)2}3] ([Bi{HMDS}3]), [Bi{N(SiMe2Et)2}3], [Bi{N(SiMe2nBu)2}3] and [Bi{N(SiMe2CHCH2)2}3]; homoleptic alkylamides [Bi(NEt2)3] and [Bi(NiPr2)3]; donor functionalised alkylamide [Bi{tBuNCH2CH2NMe2}3] and a single thioamidate [Bi{SC(Me)NiPr)3]. The presence of the thioamidate was intended to probe the effect of a conjugated soft sulfurous base on the soft acid bismuth. The commonplace silylamide, [Bi{HMDS}3] was taken forward to H2O-oxidised depositions based on thermogravimetric analysis. A narrow temperature window (190–200 °C) was found to produce binary oxide material with poor reproducibility and growth rates between 0.15 Å per cycle and 0.23 Å per cycle. Multicomponent bismuth tantalum oxide films were, however, significantly more reproducible with [Bi{HMDS}3] as a bismuth source, however the narrow temperature window between volatilisation and decomposition of precursor place large limitations on its further use.288 Findings by Rusek et al.289 (2017), during further studies into a range of bismuth amides as precursors for Bi2Te3, are consistent with proving [Bi{HMDS}3] to be the most promising bismuth source of the compounds studied, however reiterate that more thermally stable bismuth complexes are necessary to carry out >200 °C deposition. Alternative ALD routes based on the reaction of BiCl3 with [Te(Et3Si)2] have been shown to deposit bismuth telluride reliably.290,291
An investigation published in 2010 by Hatanpää et al.292 drew comparisons between a range of established and prospective bismuth precursors: Bi(OtBu)3, [Bi(OCMe2iPr)3], [Bi(OCiPr3)3], bismuth β-diketonate, [Bi{THD}3], and bismuth carboxylate, [Bi{O2CtBu}3]. ALD experiments were undertaken with an H2O oxidant with all five precursors, with the bismuth alkoxide [Bi(OCMe2iPr)3] (m.p. 40 °C) found to be the most successful when used at an evaporation temperature of 85 °C. A maximum growth rate of 0.36 Å per cycle was obtained on an Al2O3 surface at 150 °C.292 A summary of ALD precursors used for Group XV thin films is shown in Table 10.
Precursor . | Co-reactant . | Target material . |
---|---|---|
[Bi(C6H5)3]/[BiPh3] | O3 | Bi2O3277,278 |
[Bi(C6H5)3]/[BiPh3] | H2O | Bi2Ti2O7279 |
[Bi(CH2SiMe3)3] | O3 | Bismuth Silicate280,281 |
[Bi{mmp}3] | [Ti(OiPr)4]/H2O | Bi2Ti2O7282 |
[Bi{mmp}3] | [Ti(OiPr)4] or [ Ti(mmp)4], [Si(OEt)3], H2O or O3 | BTSO283–285 |
[Bi{THD}3] | H2O, H2S | Bi2O3,286 Bi2S3287 |
[Bi(HMDS)3] | H2O, [Te(Et3Si)2] | BiOx,288 Bi2Te3289 |
[BiCl3] | [Te(Et3Si)2] | Bi2Te3290,291 |
[Bi(OCMe2iPr)3] | H2O | BiOx292 |
Precursor . | Co-reactant . | Target material . |
---|---|---|
[Bi(C6H5)3]/[BiPh3] | O3 | Bi2O3277,278 |
[Bi(C6H5)3]/[BiPh3] | H2O | Bi2Ti2O7279 |
[Bi(CH2SiMe3)3] | O3 | Bismuth Silicate280,281 |
[Bi{mmp}3] | [Ti(OiPr)4]/H2O | Bi2Ti2O7282 |
[Bi{mmp}3] | [Ti(OiPr)4] or [ Ti(mmp)4], [Si(OEt)3], H2O or O3 | BTSO283–285 |
[Bi{THD}3] | H2O, H2S | Bi2O3,286 Bi2S3287 |
[Bi(HMDS)3] | H2O, [Te(Et3Si)2] | BiOx,288 Bi2Te3289 |
[BiCl3] | [Te(Et3Si)2] | Bi2Te3290,291 |
[Bi(OCMe2iPr)3] | H2O | BiOx292 |
13 Concluding remarks
It is hoped that this chapter has succeeded in its objective of providing an overview, which is by no means exhaustive, of developments within the design and development of ALD precursors. As is evident from not only the vast quantity of literature that seeks to develop and highlight new precursors, but the far wider application of ALD as a general process for the development of thin film technologies, research into ALD technologies is of increasing importance. Far more impressive is the volume of potential precursor systems described in the literature, which whilst possessing many of the fundamental properties of ALD precursors such as volatility, thermal stability and reactivity, have yet to be screened under specific ALD conditions.
As devices become increasingly more complex and a greater number of novel materials are developed, the need for precursors and processes to feed this growing demand is essential. With the drive for large volume production within the semiconductor industry and others, ALD offers scalability with the control required for miniaturisation of electronic devices, coupled with the sustainability benefits of ultrathin film deposition. However, in order to achieve this goal, there is the need to develop economically viable processes incorporating precursors that display the properties necessary for the high-throughput applications for which they are intended.293