Skip to main content
Top
Published in:

Open Access 01-03-2025 | Review

Mechanisms of ZDDP—An Update

Author: Hugh Spikes

Published in: Tribology Letters | Issue 1/2025

Activate our intelligent search to find suitable subject content or patents.

search-config
loading …

Abstract

The article updates the understanding of Zinc Dialkyldithiophosphate (ZDDP) as an antiwear additive in lubricants, focusing on recent research that has revealed new insights into its mechanisms of action. Key topics include the stages of ZDDP tribofilm formation, the impact of mechanochemistry on tribofilm growth, and the degradation and removal of tribofilms under various conditions. The article also discusses the role of other additives and contaminants, such as soot and water, in influencing ZDDP performance. Additionally, it explores the potential for computer-based modeling to describe the behavior of lubricant additives. The text is aimed at professionals with expertise in tribology and lubricant chemistry and provides a comprehensive overview of the latest developments in the field.
Notes

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Zinc dialkyldithiophosphate (ZDDP) was first patented as a lubricant additive in 1941 and has been a key component of crankcase engine lubricants for over 75 years, providing a combination of excellent antiwear action, extreme pressure protection and oxidation inhibition. In 2004 we reviewed research that had been carried out on ZDDP and summarised our understanding at that time of its mechanisms of action [1]. Since then, research has continued apace; indeed, there have been more than twice as many papers and patents published whose title mentions ZDDP, ZnDTP, or zinc dithiophosphate in the last 20 years than in all the years between 1941 and 2004. Several important new insights have emerged and the current paper reviews this recent research to update our understanding of the mechanisms of action of ZDDP.
The paper does not reiterate the history of ZDDP, or most research carried out before 2004 and for this the reader is referred to our previous review [1]. During the last 20 years research attention has focussed mainly on the ability of ZDDPs to form protective boundary films on rubbing surfaces, now widely termed tribofilms, to reduce wear and prevent scuffing. ZDDPs also possess powerful antioxidant properties that were shown in the 1960s and 1970s to originate predominantly from their ability to decompose peroxy- radicals and hydroperoxides [2, 3]. These compounds, having structure ROO. (peroxy free radical) and ROOH (hydroperoxide), respectively, where R is an alkyl group, are key intermediates in the hydrocarbon oxidation cycle [4]. This mechanism of antioxidancy was confirmed in the 1990s using 31P NMR to follow the sequence of reactions via which ZDDP decomposes hydroperoxide species [5]. Compared to work on tribofilm formation there has been very limited research since the 1990s on this aspect of ZDDPs behaviour, although a practical concern has been the extent to which blends of other antioxidants with ZDDP help retain latter’s effectiveness as an antiwear additive; the aim being to enable both extended engine oil drain intervals and lower initial ZDDP concentrations [68]. Also of interest has been the related issue of the extent to which ZDDP’s reactions as an antioxidant degrade its ability to form tribofilms and reduce wear [911].
This review first outlines key features of ZDDP and its current use in engine oils. Since our understanding of ZDDP antiwear behaviour depends very strongly on our ability to measure its interactions with surfaces and the composition and properties of the resulting tribofilms, the review then describes the main methods now being used to study this aspect of ZDDP’s behaviour. Then our current understanding of the various stages of ZDDP tribofilm development is discussed. This is important since it is now recognised that both the composition and mechanical properties of ZDDP tribofilms evolve during prolonged rubbing.
Over the last 20 years several undesirable features of ZDDP when used a lubricant additive have been recognised and these are then outlined, as are the ways that the behaviour of ZDDP is affected by other species present in lubricants, such as other additives, water and engine soot. The ability of ZDDP to form films and protect non-ferrous surfaces that have come into widespread use, such as aluminium alloys and diamond-like carbon coatings (DLC) is also examined. Finally, the last 20 years has seen rapid growth in the application of computer-based modelling to describe the behaviour of lubricant additives both at a macro-scale and at a molecular level, and the contribution of this to our current understanding of ZDDP is considered.

2 ZDDP Composition and Use

A simplified structural formula of ZDDP is shown in Fig. 1. Here the R groups are most commonly alkyl groups but a few ZDDPs in use contain alkyl-substituted aryl groups.
Fig. 1
Simplified structural formula of the monomeric form of ZDDP
Full size image
It should be noted that Fig. 1 shows the monomeric and neutral structure of ZDDP. As discussed in [1], at low temperatures ZDDPs exist in solution in an equilibrium between monomeric and dimeric forms; this has been quantified using NMR and more recently, has been modelled [12, 13]. Also most commercial ZDDPs contain a considerable proportion of a basic form [1].
In terms of ZDDP molecular structure, the most important distinction is between primary and secondary ZDDPs. Primary ZDDPs are produced from primary alcohols in which the -OH group is at the end of an alkyl chain, such as in butan-1-ol, CH3CH2CH2CH2OH. This means that the alkyl C atoms that are bonded directly to the O atoms in the ZDDP molecule (the α-carbons) are themselves also bonded to two H atoms and one C atom (or in the case of zinc dimethyldithiophosphate to three H atoms). Secondary ZDDP are made from secondary alcohols, such as butan-2-ol, CH3CH2CH(OH)CH3, where the -OH is attached to one of the internal carbon atoms of the alcohol. In consequence, the C atoms bonded directly to the O atoms in ZDDP are themselves also bonded to two C atoms and one H atom. Typical examples of primary and secondary ZDDPs based on C4 alcohols are shown schematically in Fig. 2. This apparently small difference results in a large change in reactivity so that secondary ZDDPs usually form tribofilms much faster than primary ones and are also less thermally stable. Figure 2 also shows an alkylaryl ZDDP, and a tertiary alkyl ZDDP. In the latter the C atom attached to O is bonded to three other C atoms. Attempts have been made to synthesise tertiary ZDDPs but they are thermally unstable at room temperature [14, 15] and so are not used.
Fig. 2
Examples of ZDDP alkyl and alkylaryl structures
Full size image
Because they are more thermally stable, primary ZDDPs tend to be employed in higher temperature applications, such as diesel engine oils while secondary ZDDPs are often preferred in lubricants that operate at lower temperatures. In practice however, many modern commercial ZDDPs are mixed primary/secondary, produced either by mixing primary and secondary ZDDPs, so that individual molecules in the blend will be wholly primary or secondary, or by synthesising ZDDPs from a mixture of primary and secondary alcohols, in which case some of the ZDDP molecules will contain both primary and secondary alkyl groups. Alkylaryl ZDDPs are thermally stable but relatively unreactive and are only used in high temperature applications, such as hydraulics and gas engines.
A second important feature of alkyl structure is alkyl chain length. In general, it is found that very short alkyl chain ZDDPs, such as prop-2-yl (C3) and but-2-yl (C4) form tribofilms considerably faster than those with longer chains, such as 4-methylpent-2-yl (C6).
Commercial ZDDPs are generally made by reacting P2S5 with an alcohol to form dialkyldithiophosphoric acid and then neutralising this with a zinc base, such as ZnO, Zn(OH)2, ZnCO3, or even, in academic research, Zn metal [1618]. This method can be employed in laboratory synthesis but it is also possible to prepare ZDDPs by double decomposition from an ammonium or alkali metal dialkyldithiophosphate and a zinc base or salt [14, 1921]. The zinc base employed and the pH, mixing, and temperature conditions determine the proportions of neutral and basic ZDDP and Yamaguchi et al. have described pure neutral and basic versions can be made using, respectively, zinc metal and ZnCl2/NaOH in the synthesis [22].
The alkyl structure and neutral/basic ratio of ZDDPs can be quite easily determined using a combination of 1H, 13C, and 31P NMR [12, 13].
As is evident from Fig. 1, ZDDPs contain the elements P, S, and Zn and these are essential in providing all the antiwear, extreme pressure, and antioxidant properties of the additive. However, in the 1980s it was realised that when they enter the engine exhaust as volatile compounds or particulates, these elements have the unfortunate effect of degrading exhaust aftertreatment catalysts and filters [23, 24]. Sulphur and phosphorus oxides poison or coat catalyst surfaces, while tiny inorganic metal salt particles (sulphated ash) block the filters used to remove emitted carbon particulates. Throughout the 1990s and 2000s this led to the permitted levels of P, S and metals in passenger car engine oils being progressively reduced, as shown for P and S in Table 1, adapted from [1]. In the mid-2000s it was not clear whether these levels would be reduced yet further and there was considerable research to develop “low SAPS” additives containing little or no S, P or metallic elements to replace ZDDP [25]. However, an alternative strategy was adopted, and in 2009, engine oil specifications that established a threshold phosphorus retention in the engine oil were introduced in GF-5. These maintained a maximum permitted phosphorus level of 0.08% wt. (800 ppm P) but required at least 79% of the phosphorus present in the fresh engine oil to be still present in the oil at the end of a specified engine test, thereby limiting the amount that could enter the exhaust to harm the aftertreatment system. In practice this requirement promoted the use of ZDDPs with lower volatility than some that were previously employed [2628]. Both the 0.08% wt. phosphorus limit and phosphorus-retention limit remained unchanged in GF-6.
Table 1
Phosphorus and sulphur limits in gasoline engine oils
1989
SG
No P, S limits
1994
SH, GF-1
 ≤ 0.12% wt. P
1997
SJ, GF-2
 ≤ 0.10% wt. P
2000
SL, GF-3
 ≤ 0.10% wt. P
2004
GF-4
 ≤ 0.08% wt. P, ≤ 0.50% wt. S
2009
GF-5
 ≤ 0.08% wt. P, ≤ 0.50% wt. S
Phosphorus retention, 79% minimum
(Sequence IIIGB, phosphorus retention)
2021
GF-6
 ≤ 0.08% wt. P, ≤ 0.50% wt. S
Phosphorus retention, 79% minimum
(Sequence IIIGB, phosphorus retention)
It should be noted that as well as the maximum permitted phosphorus levels shown in Table 1, not shown in this Table is that all the listed classifications include a minimum permitted level of 0.05% or 0.06% wt P. This reflects the fact that excessive wear can occur if too low a concentration of ZDDP or other P-containing additive is employed, and presumably informed the introduction of a phosphorus volatility limit rather than a further reduction in allowed phosphorus level in 2009.
Progress in reducing phosphorus content in heavy duty diesel engine oils has been less rapid. A limit of 0.12% wt. phosphorus was introduced in CJ-4 in 2006, and this was maintained in the most recent category, CK-4 in 2016. However, it seems probable, based on current work on the next proposed category (PC12), that in 2027 a maximum phosphorus limit of somewhere between 0.08% and 0.10% wt. will be introduced in the next category of heavy-duty engine oils.
The introduction of a phosphorus retention limit in 2009 has tended to stifle interest in the development of alternative low SAPS antiwear additives for crankcase engine oils, and almost all modern gasoline and diesel engine oils are still based on ZDDP as the predominant antiwear additive. However, with respect to ZDDP’s role as an antioxidant, there has been increasing use of supplementary antioxidants, to compensate for the reduction in ZDDP concentration, to manage higher engine temperatures, and to enable extended drain intervals. There has also been growing interest in the potential for other engine oil additives, such as overbased detergents, some polymeric additives and friction modifiers to enhance the effectiveness of ZDDP as an antiwear and extreme pressure additive.
In practical terms, the lubricated contact conditions in crankcase engines are becoming increasingly severe, with greater power densities coupled with the use of lower viscosity base oils to reduce hydrodynamic friction. Despite this, the author does not envisage the demise of ZDDP as the principal engine oil antiwear additive over the next few years. Within the energy transition, the crankcase engine itself appears to have a limited future and may well become extinct within the next few decades, at least in the developed world. Significant investment in new engine oil additives is thus unlikely, so the reign of ZDDP as the antiwear additive of choice in crankcase engines will probably continue until these engines themselves are no more. Also, it should be noted that ZDDPs have quite widespread use in hydraulic fluids, gearbox oils, and greases and there is little reason to suppose that their use in these applications will not continue for the foreseeable future.

3 ZDDP Tribofilm Characterisation

When metallic surfaces are rubbed together in a ZDDP-containing oil or grease, the ZDDP reacts chemically to form a typically 50 to 150 nm thick film on the rubbing surfaces. Figure 3 shows the structure of a ZDDP tribofilm as understood in 2004 [1]. It consists of a thin metal sulphide film, overlain with much thicker pads of primarily iron and zinc phosphate and polyphosphate, separated by deep valleys in which there is negligible film. The pads are typically 1 to 5 μm in diameter and up to 150 nm high.
Fig. 3
Schematic of cross section of ZDDP tribofilm in 2004. Reprinted with permission from ref [1]. Copyright (2004) Springer
Full size image
This overall picture has changed relatively little in the intervening years, but, as outlined in Sect. 5 below, we now know more about the sequence of reactions involved in forming this structure, and far more about the properties of the tribofilm.
In our 2004 review it was shown how the growth in understanding of ZDDP has been largely controlled by the application of new analytical methods to study ZDDP itself, its behaviour in solution, and the nature of films that it forms on tribological surfaces. Table 2, which is adapted from [1] by extending it to the present day, lists some of the many experimental techniques that have been applied in ZDDP research. In 2004, as described in [1], two analytical methods in wide use in research to study tribofilms were X-ray absorption near edge structure, XANES, to measure the composition of ZDDP films, and AFM to map their morphology at very high resolution. These have continued to be important and have been augmented by others, notably the spacer layer imaging interferometry method (SLIM) [29] and focussed ion beam milling-based analysis (FIB) [30].
Table 2
Introduction of experimental tools to study ZDDPs
1950s
Wet chemistry
Radiotracer techniques, 32P, 35S 65Zn
Optical interference microscopy
1960s
GC, GC/MS
Infrared spectroscopy
XRF
1970s
XPS
Auger
SIMS
1H and 31P NMR
1980s
STEM
EELS
EXAFS
1990s
Cryogenic surface analysis
XANES
AFM and other nanoprobes
2000s
X-PEEM
Raman
Spacer layer interferometry (SLIM)
Scanning white light interferometry (SWLI)
Nanoindentation
FIB-based methods (TEM/SEDX/XRD)
Synchrotron XPS
2010s + 
Quartz crystal microscopy (QCM)
In situ AFM
Atom probe tomography (APT)
Another important advance has been the development of in situ methods that are able to characterise ZDDP tribofilms on surfaces within a rubbing test rig [31, 32] These have the advantage of avoiding cooling or cleaning the ZDDP tribofilm prior to analysis, which might change its properties or composition. They also enable the evolution of the tribofilm over time to be studied without having to perform a series of tests of different duration.
Additionally, of course, general improvements to instrumentation and data processing have meant that analytical techniques used in the past have become more sensitive, powerful, and easier to use in recent years, as well as, often, less expensive.

3.1 Measurement of ZDDP Tribofilm Thickness and Morphology

Three experimental techniques have become quite widely used to measure and map the topography of ZDDP, and indeed other additive tribofilms, over the last 20 years—atomic force microscopy (AFM), spacer layer interferometry (SLIM), and scanning white light interferometry (SWLI).
Atomic force microscopy (AFM) was first applied to study the tribofilms formed by ZDDP in 1997 [33] and in recent years it has become routinely employed to map the film thickness, topography, and friction of these films [3439]. In 2006 Topolovec et al. used AFM to show that the ZDDP tribofilms form only on rubbed tracks and not on the surrounding surface, as demonstrated in Fig. 4, which also illustrates clearly the pad-like structure of the film material [34]. These authors also pioneered the use of ethylenediamine-tetraacetic acid (EDTA) solution to remove ZDDP films from a selected part of the rubbed track to reveal the substrate beneath the film, both to measure wear and to provide a reference surface to determine accurately the ZDDP tribofilm thickness. In 2015 Kalin et al. used a combination of AFM and lateral force microscopy to map the topography, stiffness, and surface shear resistance of ZDDP tribofilms on both steel and DLC coatings [37]. They confirmed that the tribofilms had a pad-like structure and found that they formed only on the asperities and had lower friction and adhesion properties than the inter-pad regions. Rydel et al. employed AFM to study surfaces before and after ZDDP tribofilm removal to correlate microstructure with ZDDP film formation [38]. They found less tribofilm on carbides than on the martensitic matrix and suggested that this might originate from less adsorption of ZDDP on the carbides or preferential loss of film from the carbides due to weaker adhesion.
Fig. 4
AFM topography map and profile of secondary ZDDP tribofilm on an MTM disc track. Reprinted with permission from ref. [34]. Copyright (2007) Springer
Full size image
An important development in 2015 was the use by Gosvami et al. of a liquid cell AFM to simultaneously generate and monitor ZDDP tribofilm formation in situ [40]. Rastering of the AFM tip back and forth in contact mode against a substrate immersed in ZDDP solution generated a tribofilm whose topography could be monitored continuously from the tip’s vertical displacement. The authors employed this to show the exponential dependence of film growth rate on contact pressure and explained this in terms of a mechanochemical reaction rate model. This important concept will be further discussed later in this review. Several researchers have subsequently applied this in situ AFM method to monitor ZDDP film formation on various substrates, e.g. [4148].
AFM normally employs a cantilever with a built-in, very small tip, typically of Si3N4 or Si. This gives very high spatial resolution but high contact pressure. To produce less severe conditions somewhat closer to macro-scale contacts, colloid probe AFM can be used, This uses a tiny sphere attached to the cantilever and was employed in 2005 by Topolovec et al. with a 10 μm diameter glass sphere to measure the friction of individual ZDDP tribofilm pads and also normal force curves during tip approach and withdrawal from a ZDDP tribofilm [49, 50]. The latter suggested that secondary ZDDP might have a thin outer crust of material harder than the film below it. More recently, Gosvami et al. have used AFM cantilever-mounted 50 μm diameter steel balls against a steel flat to study ZDDP tribofilm growth in situ [51]. They found similar tribofilm build up during rubbing in ZDDP solution to that seen using a small Si AFM tip.
Spacer Layer Interferometry (SLIM) was first employed in 2000 to measure and monitor ZDDP tribofilm thickness in rolling-sliding tests in a minitraction machine (MTM) [52]. In the MTM a ball is loaded and rubbed against the flat surface of a disc immersed in lubricant. To apply SLIM, the ball is periodically unloaded from the disc and loaded upward against an optical film-coated, transparent window. An interference image of the separation between window and the reflective steel ball surface is then captured. This separation originates from the presence on the ball surface of a transparent tribofilm, so calibration of interference colours can be used to obtain a map of the thickness of the ZDDP tribofilm [53, 54]. Since this is employed in situ, without cooling or removing the rubbed surface from the test rig, SLIM enables the development of the film to be monitored during a rubbing test as shown in Figs. 5 and 6. Figure 5 shows sets of typical SLIM interference images from the ball obtained during tests with a PCS Instruments MTM with different ZDDPs [34]. Figure 6 shows the effect of slide-to-roll ratio on mean tribofilm growth by a primary ZDDP obtained from SLIM images, in this case using a Wedeven Instruments WAM ball-on-flat-ring tribometer rather than an MTM [55].
Fig. 5
SLIM interference images of ZDDP tribofilms formed from five ZDDP solutions on MTM balls during 4 h sliding-rolling tests; ZDDP1 is a primary alkyl, ZDDP2s are secondary alkyls and ZDDPar an alkylaryl ZDDP. Reprinted with permission from ref [34]. Copyright (2007) Springer
Full size image
Fig. 6
Influence of slide-roll ratio (SRR) on ZDDP tribofilm formation in ball on ring tests measured using SLIM with a WAM tribology test machine. SRR is negative, i.e. ball travels slower than counterface. Reprinted with permission from ref [55]. Copyright (2011) Sage Publications
Full size image
The durability of the film and its interaction with other additives can also be studied by replacing the ZDDP-containing lubricant with base oil or a formulated oil after the ZDDP tribofilm has formed and then continuing rubbing [5658]. As described later, SLIM has shown an evolution in the strength of ZDDP tribofilms during prolonged rubbing [59, 60]. It has also revealed that ZDDP films are often susceptible to partial or complete removal by additives that contain amine groups [56, 57, 61, 62]. Over the last decade the use of SLIM in an MTM (MTM-SLIM) has become a very widely used tool to study the kinetics of formation of ZDDP and other lubricant additive tribofilms [34, 39, 43, 59, 6367].
In the 2000s there was growing interest in applying interference microscopy, such as scanning white light interferometry (SWLI), to measure surface topography [68] and towards the end of this decade some researchers started to apply it to surfaces rubbed in ZDDP solutions, with the aim of measuring the wear beneath any tribofilm present. However, in 2009 Benedet et al. showed that this approach could suggest, misleadingly, a high level of wear that actually originated from internal reflections within the transparent tribofilm [36]. The authors also showed how this problem could be addressed by (i) removing part of the tribofilm with EDTA and (ii) coating the whole surface with a 25 nm thick reflective gold film. This enabled the ZDDP tribofilm thickness and topography as well as any wear beneath the film to be quantified.
Recently Dawcyzk et al. have compared the above three methods of measuring ZDDP tribofilm thickness and roughness and discussed their advantages and disadvantages [69]. Based on the many studies carried out using all the above methods, it is evident that ZDDP films grow over a period of typically one to four hours of rubbing to form solid tribofilms that, for simple ZDDP solutions on steel surfaces, stabilise at between 100 and 200 nm mean thickness. These films are generally very rough and consist of pads typically 1 to 5 μm in diameter separated by deep valleys, as can be seen in Fig. 4 and schematically in Fig. 3. In reciprocating contact, the pads can be circular, but in unidirectional sliding or rolling-sliding conditions they are often elongated in the sliding direction. In formulated engine oils, ZDDP tribofilms tend to be thinner, in the 20–60 nm range, are often smoother, and may consist of bands of different thicknesses oriented along the sliding direction.
It should be noted that the AFM and MTM/SLIM methods of monitoring the formation of ZDDP tribofilms described above are very different. The AFM has a single, very high pressure contact, sliding at very low speeds, while MTM/SLIM contacts are rolling/sliding, with much higher sliding speeds and generally lower pressure contact conditions; they are much more realistic of lubricated machine components. There appear to have been no systematic attempts to relate the two methods in terms of ZDDP tribofilm formation, except to note that both appear to form rough tribofilms films and give tribofilm formation rates that increase exponentially with temperature and shear stress. In in situ AFM experiments, films are generally formed by a tip rastering over an area between 1 and 100 μm square, and the morphology of the tribofilm appears to be scaled to this size, with primarily smaller pads than the 1–5 μm ones formed in the MTM. While some research has studied the durability of the tribofilms formed in the MTM, the extent to which film removal takes place during film growth in AFM contacts has not been established and may be considerable. The very great differences between the two methods makes it difficult to relate them but one possible way forward might be to carry out further work using colloid probe AFM as outlined above [51]. This forms a much larger contact, with lower pressure and less likelihood of indenting a tribofilm surface than the conventional AFM tip; it may provide a bridge to link the AFM and MTM methods. For comparison, SLIM could then be applied to monitor film formation on a macro-scale ball in pure sliding mode, as has been used previously both in the MTM [70] and in an in situ Raman study of tribofilm formation [71].

3.2 Measurement of Mechanical Properties of ZDDP Tribofilms

ZDDP tribofilms are too thin for their mechanical properties to be measured by macro-scale indentation methods, but since the early 1990s they have been studied by nanocontact techniques, such as AFM [33], interfacial force microscopy [72, 73], and surface forces apparatus [74]. In recent years there has been growing availability of commercial nanoindentation instruments and these have also been quite widely employed to map the elastic modulus and hardness of ZDDP tribofilms [7582]. A surprising variety of values of reduced elastic modulus, E/(1-v2), have been measured, ranging from about 40 GPa [77] [82] to between 80 and 150 GPa [47, 73, 75, 78, 80, 81]. One study has even determined a value of around 200 GPa on the top of ZDDP tribofilm pads, comparable to steel [73]. Some of this variation may represent differences in elastic moduli at different regions of the tribofilm [72] and varying conditions and times of rubbing, but some may originate from the depth of measurement into the film being studied and interactions with the substrate stiffness. Hardness measurements also span a quite wide range from 1.5 up to ca 6 GPa and from AFM measurements, Demmou et al. have suggested that both hardness and elastic modulus of the tribofilm increase linearly towards the film’s interface with the substrate [77].
One limitation of early work was that measurements were only made at room temperature, and it was not known whether the mechanical properties of ZDDP tribofilms might be quite different at elevated temperature, especially if the films had a glassy structure. More recent studies using commercial indenters with variable temperature capabilities have shown that the elastic modulus of ZDDP tribofilms is quite constant from room temperature up to 200 °C [77, 78, 80], though a few have shown that hardness may slightly decrease with increasing temperature [77].
All the above measurements were made at end of a rubbing test, after a ZDDP had fully formed. Recent work has used an AFM to form a tribofilm and study its properties in situ as it develops [48]. This indicates that, when initially formed some ZDDP tribofilms can flow slowly during sliding or under static squeeze, suggesting they are highly viscous, with viscosity ca 100 to 1000 GPa.s, comparable to a molten glass. This is consistent with the concept, discussed later in this review, that fast-forming ZDDP tribofilms are initially relatively weak with an amorphous structure, but become more crystalline and wear-resistant during prolonged rubbing.
It should be noted that accurate determination of the mechanical properties of sub-100 nm films on rubbed surfaces is problematic, even with modern instruments, and this may be the origin of some of the discrepancies in the literature. There are a variety of problems, including possible pile-up, limited depth resolution, extreme variations in thickness of film across the surfaces and decoupling the stiffness and hardness of the substrate from that of the film [83]. In general, it has been found that oscillatory methods, such as continuous stiffness measurement are more effective than individual loading/unloading cycles since they allow continuous measurement of stiffness along the loading curve [84, 85].

3.3 Measurement of Chemical Composition of ZDDP Tribofilms

As indicated in Table 2, many analytical methods have been applied to study the composition of ZDDP tribofilms. Two key advances in the last fifteen or so years have been (i) the application of methods to study how tribofilm composition varies with depth, notably using focussed ion beam milling (FIB), and (ii) the development of ways to observe ZDDP tribofilm composition in situ, i.e. as the film is formed within a tribometer.
A limitation of most early ZDDP film analysis was that measurements were averages over the penetration depth of the technique. An exception was XANES that could provide information from two depth scales by measuring either emitted photons or electrons [86]. Another approach was to analyse a tribofilm while progressively etching away the surface, typically using an argon ion sputtering [87]. Both these approaches showed that the composition of ZDDP tribofilms often varies with depth, but both have limitations. XANES probes the average of two quite poorly defined depths, while argon etching may change the composition of the surface being probed [88]. Also, unless very small regions are etched and analysed, the roughness of ZDDP tribofilms complicates interpretation of depth information.
In 2007 Heuberger et al. used variable angle XPS to investigate the local variation of tribofilm structure through the first few nanometres of secondary C3/C6 ZDDP tribofilms and explore the degree of phosphate polymerisation as a function of depth, as discussed later in this review [89]. This approach avoids potential complications arising from sputtering and was extended in 2010 by Zhou et al., who employed synchrotron source-based XPS with variable energy photons to probe different depths of secondary C4/C6 ZDDP tribofilms [90].
A very significant advance in providing information as to how ZDDP tribofilms vary through their thickness has been to analyse lamellar samples milled and extracted from rubbed surfaces using a focussed ion beam (FIB). This was first employed to examine ZDDP tribofilms in 2004 [91], but in recent years has become quite extensively applied in tribology as the technique has become more widely available. In most applications of FIB, the surface of interest is first protected by a coating, typically of platinum, and then parallel vertical trenches are milled using a beam of high energy gallium ions to leave a thin wafer between them, as shown in Fig. 7A. The wafer is then extracted and thinned to provide a sample representing a cross section through the protective coating, the tribofilm, and the underlying substrate. This can then be analysed by techniques, such as transmission electron microscopy (TEM), X-ray diffraction (XRD), and energy dispersive X-ray spectroscopy (EDX) to determine how the structure or elemental composition of the film varies with depth. A typical TEM image of the cross section of a ZDDP tribofilm together with EDX analysis across the film is shown in Fig. 8 [92].
Fig. 7
ZDDP tribofilm FIB specimen preparation. Reprinted from ref. [93]
Full size image
Fig. 8
a TEM image of the cross section of a secondary C6 ZDDP tribofilm, b EDX analysis showing the distribution of elements across the film. Reprinted with permission from ref. [92]. Copyright (2006) Elsevier
Full size image
Two limitations of using EDX analysis of a FIB wafer to obtain elemental distribution are (i) the wafer thickness and thus signal can vary across its width and (ii) the focussed electron beam excites a finite volume within the wafer with greater diameter than that of the beam itself and this limits the depth resolution to tens of nanometers [94].
One quite new method of obtaining higher resolution elemental distribution in ZDDP tribofilms is atom probe tomography (APT) [95, 96]. In this, a needle-shaped specimen is obtained using FIB, either from an extracted wafer or by annular milling of the rubbed surface [97]. This is then cooled to cryogenic temperatures and bombarded by a laser to emit ions that are collected by a mass spectrometer. Based on the sample tip geometry, a 3D map of the spatial origin of the detected ions can be constructed as illustrated in Fig. 9 [95]. To date APT has been applied only sparingly in tribology and has served to confirm compositional information obtained from other methods. The method is so new that its limitations are not yet fully determined. One is that very stringent solvent cleaning of the sample is required prior to analysis in vacuum, and this may remove some of the tribofilm [96]. It is also not known whether ion bombardment during sample preparation may result in changes to the tribofilm, an issue that is also relevant to other FIB-based analysis methods [93].
Fig. 9
3D APT reconstruction of the tribofilm formed by secondary ZDDP. Reprinted with permission from ref. [95]. Copyright (2017) American Chemical Society
Full size image
The second major advance in the chemical analysis of ZDDP tribofilms has been the use of in situ methods, where the composition of the film on a rubbed surface is monitored either periodically or continuously during a rubbing test without removing the rubbed sample from the tribometer.
In 2003 Piras et al. studied ZDDP tribofilm growth on an iron coating on a germanium crystal using attenuated total reflection infrared spectroscopy (ATR-FTIR) to probe the contact through the germanium. This enabled IR spectra of the film to be obtained in situ but was limited both by the fragility of the iron coating and because the ATR method measures an average spectrum over a large area. A more versatile approach has recently been developed by Okubo et al., who extended the MTM-SLIM approach to obtain Raman spectra from the rubbed track on a steel ball [71]. In MTM-SLIM the ball is periodically uploaded against a coated glass disc to capture an optical interference image, but Okubo also used a Raman spectrometer to obtain a spectrum from the ball track.
In general, X-ray and electron-based analytical methods cannot be applied to study tribofilms in situ due to adsorption of the exciting and emitted species by air or windows. One recent exception is the use of XANES to monitor tribofilm composition from a rubbed disc within a pin on disc tribometer [98, 99]. To prevent the relatively low energy x-rays required to obtain spectra from low atomic number elements of interest, such as P and S being absorbed by air, the tribometer was enclosed in a helium atmosphere during tribofilm analysis.
For all the above analyses, both ex situ and in situ, it is important to emphasise that mechanical and chemical analysis of ZDDP tribofilms is not straightforward. These films are thin, rough, inhomogeneous into their depth and possibly also across their area. They are also often fragile, at least in terms of conditions imposed by bombardment with ions, x-rays, or focussed laser beams, as used some analysis methods. Modern methods of mechanical and chemical surface analysis can address some of these problems, but it is important that the researcher be aware with their limitations as noted above.

4 Mechanochemical Formation of ZDDP Tribofilms

It has long been recognised that ZDDP can form both thermal films and tribofilms [100102]. Thermal films are formed on surfaces simply immersed in ZDDP solutions at high temperatures, typically above 140 °C [86, 103105], while tribofilms are generated on rubbing surfaces and can form at much lower temperatures [56], even as low as 25 °C [103]. A considerable amount of research has compared the two types of film and it is now generally considered that they are chemically broadly similar, at least initially, although, as discussed in Sect. 5.5, tribofilms undergo a reduction in phosphate chain length during extended rubbing and appear to be more wear-resistant than pre-formed thermal ones [106].
Almost all chemical reactions are very strongly dependent on temperature, with reaction rate generally increasing exponentially with temperature as described by the Arrhenius equation. It has been shown that the initial growth in thicknesses of both ZDDP thermal and ZDDP tribofilms is indeed exponential with temperature [103, 107]. However, in recent years there has been increasing realisation that some chemical reactions rates are equally strongly dependent on applied stress and models have been developed that extend the Arrhenius equation to include a stress term in the exponential. The resulting rate constant is shown in Eq. 1 where A is a pre-constant, Ea is the activation energy, N is Avogadro’s number, τ is the applied stress, ∆v is the stress activation volume, R is the universal gas constant and T is the absolute temperature. The Arrhenius term is simply e−Ea/RT, while the applied stress contributes the term eNτ∆v/RT
$$k = Ae^{{ - \left( {E_{a} - N\tau \Delta v} \right)/RT}} = Ae^{{ - \frac{{E_{a} }}{RT}}} e^{{\frac{N\tau \Delta v}{{RT}}}}$$
(1)
This dependence of chemical reactions on applied stresses is now called mechanochemistry and has become a quite important branch of chemical research and application over the last two decades [108110].
In 2015 Gosvami et al. showed that the rate of tribofilm formation by ZDDP increased exponentially with applied pressure, suggesting that ZDDP tribofilm formation is mechanochemically controlled [40]. They used an AFM to rub a tip against a flat immersed in a solution of a mixture of primary and secondary ZDDPs in a base oil and simultaneously employed the tip vertical displacement to monitor tribofilm thickness. Figure 10 shows how ZDDP tribofilm thickness growth rate increased with pressure over 2000 sliding cycles at 100 °C. A similar exponential dependence on temperature at constant pressure was also observed, to give temperature and stress dependence in accord with Eq. 1.
Fig. 10
Dependence of mixed primary and secondary ZDDP tribofilm growth rate in an AFM at 2000 cycles and 100 °C. Reprinted with permission from ref [40]. Copyright (2015) The American Association for the Advancement of Science
Full size image
Since tribofilm formation increased exponentially with the pressure at the tip/flat contact, Gosvami et al. assumed that pressure was the stress driving the reaction. However, in 2016, Zhang and Spikes showed that a ZDDP tribofilm could form from mixed primary/secondary ZDDP solution in full film EHD lubricated conditions in an MTM where there was no asperity-asperity contact [111]. By using base fluids with different EHD friction properties at the same pressure, they showed that ZDDP tribofilm formation rate depended on the shear stress rather than the pressure in the contact [111]. This separation of the effect of pressure from that of shear stress was not possible in boundary lubrication conditions where shear stress is generally broadly proportional to pressure.
Several further studies have confirmed that the initial rate of growth of ZDDP tribofilms depends exponentially on applied shear stress [42, 112]. Very recently Fang et al. extended Zhang and Spikes’ approach by using blends of high and low EHD friction base oils to deconvolute the effect of shear stress and pressure on a secondary ZDDP tribofilm growth rate [113]. They found that the growth rate increased rapidly with shear stress but also decreased very slightly with pressure, as indicated in Eq. 2.
$$k = Ae^{{ - \left( {E_{a} - N\tau \Delta v + Np\Delta v2} \right)/RT}} = Ae^{{ - \frac{{E_{a} }}{RT}}} e^{{\frac{N\tau \Delta v}{{RT}}}} e^{{ - \frac{Np\Delta \Omega }{{RT}}}}$$
(2)
The stress activation constant, ∆v was approximately 100 times as large as the pressure activation constant ∆Ω.
It has been suggested that small molecules must be tethered to a surface to experience sufficient mechanical force to undergo mechanochemistry [114], but it is not yet known whether the critical stress that applies this force is a shear stress or some other associated asymmetric stress such a tensile one. In other mechanochemical research it has been found that tensile applied forces can pull covalent bonds apart, as, for example in the permanent degradation of polymers under shear [115], but it has also been shown that applied stresses can change bond angles and thus destabilise molecules [116].
The insight that ZDDP tribofilm formation is driven by shear stress explains many features of ZDDP tribofilm formation, for example why films are not formed in pure rolling conditions as seen in Fig. 6 above, since shear stress is negligible in the absence of sliding, and why the films have a pad-like structure. As a ZDDP tribofilm develops, presumably initially on asperities [37], the resulting regions will bear more and more of the load (and thus shear stress) in mixed and boundary lubrication and so will experience fastest film growth [111]. In the absence of other additives, ZDDP films can grow up to about 150 to 200 nm thickness, but their growth then levels out. The reason for this levelling has not been fully explained but may result when the relatively low elastic modulus or hardness of the tribofilm compared to the substrate starts to reduce the effective shear stress present at the top of the film [111]. This is further discussed in Sect. 11.1 on modelling tribofilm formation later in this review.
It is tempting to conclude that thermal and tribofilms are formed by essentially the same underlying mechanism, the difference being simply that thermal films form slowly at zero applied stress while tribofilms form much faster and at lower temperatures because of a finite stress contribution. However, this has not been proved, and it remains possible that there may be different film-forming molecular reactions at low and at elevated temperature.

5 Stages of ZDDP Tribofilm Formation

The structure of a typical ZDDP tribofilm was shown in Fig. 3. We now know that its formation involves essentially five stages [89]
(i)
adsorption of ZDDP molecules on surfaces
 
(ii)
reaction to form an iron sulphide or sulphate
 
(iii)
possible transfer of alkyl groups from O to S
 
(iv)
polymerisation of adsorbed phosphate/thiophosphate to form an amorphous polyphosphate film on top of the S-based film
 
(v)
depolymerisation of polyphosphate to form a nanocrystalline and thus more wear-resistant zinc and iron phosphate film.
 
Our current knowledge of each of these stages is outlined below.

5.1 Adsorption of ZDDP on Ferrous-Based Surfaces

For ZDDP to form tribofilms, its molecules or those of its decomposition products must first adsorb on the rubbing surfaces, and several researchers have studied the initial adsorption of ZDDP on a variety of surfaces. Early work used radiotracers based on 13C- and 65Zn- labelled ZDDP [117123] and microcalorimetry [20], but recently quartz crystal microbalance (QCM) has been also applied [124126].
Dacre and Bovington used 65Zn- and 13C- labelled isopropyl ZDDP and found that ZDDP adsorbed to form a monolayer (stated to be ca 0.9 nm2/molecule) on both iron powder and 1% Cr steel [120]. At low temperatures, the ratio of adsorbed 13C and 65Zn atoms implied complete molecules on the surfaces, but above a critical temperature of about 60 °C, Zn was lost from the surface. The authors suggested that this corresponded to the splitting of ZDDP molecules into individual, adsorbed (RO)2PSS ions. Early work using XANES provided some support for this loss of Zn [127]. Most other adsorption studies have also found monolayer adsorption of ZDDP on metals, with Langmuir-type dependence on solution concentration, although there are few discrepancies. For example one study found strong adsorption of a primary ZDDP on iron [122]. but another found no adsorption whatsoever of ZDDPs on very high purity iron powder [20].
Yamaguchi et al. used inelastic tunnelling spectroscopy to study the adsorption of both commercial and synthesised ZDDP on Al2O3 surface at cryogenic temperature and concluded that adsorption was via the S atoms, with the rest of the molecule oriented away from the surface [22, 128]. For an alkylaryl ZDDP, dissociation of the adsorbing molecules to form cresols was identified, but this was not seen with primary and secondary alkyl ZDDPs. Because of the use of dithiophosphates as mineral flotation agents, considerable research has also been carried out on their adsorption on metal sulphides and this confirms that the sulphur atoms in dithiophosphate ions bond initially to the metal ions in the sulphides [129, 130]
Overall, it appears that ZDDP adsorbs readily on surfaces and that monolayer adsorption is effectively complete at low concentrations (< 500 ppm P), so that tribofilm formation rate becomes independent of concentration above this value [125]. Recent QCM work has shown that ZDDP also adsorbs on ceramics, but at a lower coverages than on metal oxides at comparable concentrations, and it has been suggested that the extent of adsorption depends on a ceramic’s ionicity [125].
One outstanding question is what happens after initial, physical adsorption to form a monolayer? No recent studies appear to have progressed beyond the important observation by Dacre and Bovington of a chemical reaction at elevated temperature that releases Zn and possibly forms a ferrous dithiophosphate. This type of cation exchange has been observed in solution [131133] and might well occur on surfaces and, since FeDDP is known to be much less chemically stable than ZDDP [131]. It may be an important precursor in the subsequent reaction described below to form a ferrous sulphide. As considered later in this review, molecular modelling may help resolve this question in future.

5.2 Reaction to Form Sulphide/Sulphate

Several early studies identified a sulphur-rich layer at the tribofilm/substrate interface [74, 134], but others found sulphur to be distributed throughout the tribolayer [87, 91]. Quite recently Soltanahmadi et al. used FIB/EDX to analyse ZDDP-derived tribofilms on rubbed tracks from a micropitting rig. They identified a sulphur-rich layer of thickness 5 to 10 nm between the steel substrate and the phosphate film [135]. Since this layer was not associated with a zinc-rich layer they ascribed it to iron sulphide. A sulphur-rich layer between the substrate and the main tribofilm formed from a secondary ZDDP has also been very recently detected using atom probe analysis [96].
In 2016, Shimizu et al. employed AFM to measure the morphology of the film formed by a primary C8 ZDDP on an MTM ball during the first minute of rubbing in pure sliding conditions [39, 70]. After 30 s rubbing, they observed large lumps of a sulphur-rich material, typically 10 μm diameter and up to 700 nm high. After a few minutes rubbing these lumps broke up into smaller fragments and then became mixed with, and overlain by, a P-rich tribolayer. These sulphur-rich particles were only formed on the stationary ball that was continually in contact and not on the disc, nor were they formed in rolling-sliding conditions where both surfaces move through the contact. The authors suggested that their formation was favoured under severe conditions when a phosphate film that might prevent their growth was unable to form rapidly on the surface. This ability of ZDDP to form a sulphide film most easily when a phosphate film could not develop was also noted by Kontou et al., who studied the tribofilm formed by ZDDP in lubricants that also contained carbon black particles and dispersant in rolling-sliding conditions [136]. As discussed later in this review, soot and carbon black can rapidly abrade ZDDP tribofilms, and using FIB and EDX, Kontou found that with some dispersants this resulted in the formation of surfaces having no phosphate tribofilm but instead a quite thick iron sulphide layer, as shown in Fig. 11.
Fig. 11
TEM/EDX showing a thick iron sulphide tribofilm formed by a base oil containing secondary ZDDP, carbon black, and dispersant [136]. The steel substrate is on the right and a protective gold coating is on the immediate left. A platinum coating overlays the gold coating. Reprinted from ref [136]
Full size image
Recently Dorgham et al. have used XAS to monitor tribofilm formation on the disc of a pin-on-disc apparatus by a secondary ZDDP using a synchrotron X-ray beam. Their results indicated the initial formation of a zinc or iron sulphate, but then, as rubbing continued, mainly sulphide was produced. The authors suggested that the sulphate might have been formed only until the initial iron oxide layer on the steel was consumed [98, 137].
From the above it is evident that iron sulphide can form very rapidly on ferrous surfaces rubbed in ZDDP. It appears that sulphide generation is more prevalent in pure sliding conditions and before the development of a phosphate film, or when such a film cannot develop. As discussed later in this review, this may illustrate the extreme pressure role of ZDDP – to form a metal sulphide in extreme conditions, such as the sliding cam-follower interface and thereby prevent scuffing even when a phosphate-based antiwear film cannot be maintained.

5.3 Transfer of Alkyl Groups from O to S

An outstanding question concerning tribofilm formation by ZDDPs is whether, prior to phosphate film formation, alkyl groups in the dialkyldithiophosphate are transferred from O to S to form the structural isomer di(thioalkyl)phosphate via the reaction;
https://static-content.springer.com/image/art%3A10.1007%2Fs11249-025-01968-3/MediaObjects/11249_2025_1968_Figb_HTML.png
(3)
and whether this reaction plays a key role in driving tribofilm formation.
Thiophosphates and dithiophosphates are known to be strong alkylating agent and to be able to self-alkylate as shown in Eq. 3 [138]. In 1981, Coy and Jones studied the thermal decomposition of primary and secondary C4 ZDDPs at 180 °C and used 1H and 31P NMR to identify a range of intermediates that contained RS groups bonded to P [139, 140]. It is quite appealing to consider that similar transfer might occur in tribofilm formation at lower temperatures, since a thioalkyl RS- group bonded to P might be expected to be more easily displaced than an alkoxy group RO-, leading to easier phosphate polymerisation. In 1998 Fuller et al. applied XANES (XAS) to study the formation of tribofilms on surfaces from mixed primary/secondary ZDDP solutions preheated at 200 °C for several hours and observed peaks that they tentatively ascribed to the di(thioalkyl) phosphate species shown in Eq. 3 [141]. No reference samples were, however, available to confirm this.
At present there is no direct experimental evidence that the reaction shown in Eq. 3 helps drive ZDDP tribofilm formation in low temperature rubbing contacts (ca 20–120 °C). However, there is some indirect evidence that this type of rearrangement is not essential for tribofilm formation, since, as indicated later in this review, sulphur-free zinc dialkylphosphates (ZDPs) form thick polyphosphate and phosphate tribofilms on rubbing surfaces at a similar rate, to a similar thickness and with a similar dependence of alkyl structure to their zinc dialkyldithiophosphate analogues [142, 143].
We do know, however, that most of the sulphur in ZDDP is lost during tribofilm formation, so it is probable that the type of exchange shown in Eq. 3 must occur at some stage to facilitate the loss of alkyl sulphides. The outstanding question is whether this precedes and enables the formation of the phosphate film or occurs only during and after film formation. Modelling work by Mosey and Woo has suggested that it occurs relatively slowly and at higher temperature than other potential phosphate tribofilm-forming reactions [144]. In situ Raman analysis may help resolve this question in future research.

5.4 Polymerisation

It has been known since the 1950s that the bulk of a ZDDP tribofilm consists of zinc phosphate-based material [100, 101], and since the 1980s much research has been devoted to determining the precise composition of this film and thereby its mechanism of formation. Metal phosphates have been widely studied outside of tribology because of their use as phosphate glasses [145] and they are well known to form a range of molecular structures based on P-O-P bonds [146]. The various possible phosphate structures are shown in Fig. 12 below, taken from [1]
Fig. 12
Phosphate structures. a orthophosphate, b pyrophosphate, c a polyphosphate, d metaphosphate, e ultraphosphate. Reprinted with permission from ref. [1]. Copyright (2004) Springer
Full size image
In a ZDDP tribofilm, the negatively charged oxygen atoms will generally be balanced by positively charged Zn or Fe cations. The polyphosphate shown in Fig. 12c has four P atoms, but much longer chains are possible. In metaphosphates, ring structures are formed so that there is only one anionic charge on each phosphorus group, while in ultraphosphates some phosphorus groups have no anionic charge since three of their O atoms form bridges to other phosphorus atoms. In the extreme case, the ultraphosphate has no PO moieties at all and thus no overall charge; so forming an allotrope of P2O5. The extent of polymerisation is often described by the ratio of the number of bridging oxygen atoms (BO) that link two P atoms to the number of oxygens that are connected to just one P and are thus non-bridging (NBO). This ratio varies between zero (orthophosphate) and 1.5 (extreme ultraphosphate), and in tribofilms it has been quantified using XANES [147], XPS [89], infrared [148], and Raman [149] surface analysis.
It was gradually realised during the 1990s that the phosphate material in ZDDP tribofilms was a mixture of short chain phosphates and longer chain polyphosphates, with generally a higher polyphosphate content close to the outer surface of the film and short chain phosphate closer to the metal substrate [147]. Very little, if any, sulphur was present except close to the steel surface. Since polyphosphate is found in the upper layers of the tribofilm it is presumed to be formed initially during rubbing and its formation to be mechanochemically driven by the high shear stresses present at asperity contacts.
There is growing evidence that the initial structure and properties of ZDDP tribofilms depend on how rapidly they are formed. Some initial ZDDP tribofilms can be quite soft and may even flow at elevated temperatures, as described in Sect. 3.2. This might explain the elongated island shape seen in some conditions [48]. Several studies using SLIM have shown that when ZDDPs form tribofilms very quickly, part of the film can be lost during rubbing. This can be seen in Fig. 6 at high SRR, and was also noted by Fujita et al. [56] and, very recently, by Ueda et al. [150]. Ueda et al. used XPS to analyse the surfaces of tribofilms formed very rapidly by short chain, secondary ZDDPs and found that these had a significant ultraphosphate component, as also noted in previous work using similar ZDDPs [89, 90]. However, when the tribofilm was partially lost, the remaining film was predominately short chain polyphosphate. It thus appears that fast-forming ZDDP tribofilms are initially long chain poly- and ultra-phosphates with few Zn cations and are thereby quite soft and easily worn. However, as described in the next section, during extended rubbing (or in slow growth rate conditions) they progressively depolymerise, taking up more Zn and eventually Fe cations, to become harder and more wear resistant.
One of the least well understood aspects of the formation of polyphosphate tribofilms concerns the molecular mechanisms by which dialkyldithiophosphate ions or their structural isomer di(thioalkyl)phosphate ions (Eq. 3) link together to form P-O-P bridges and thence ultra- or polyphosphate. The general process is presumed to involve attack by an oxygen from one phosphate ion on the P atom of another, but we still have very limited information about how this occurs.
Most of our insights come from quite early research studying the volatile products evolved during thermal degradation of ZDDPs using gas chromatography. This detected alkenes and alkylsulphides and proposed three main decomposition mechanisms to form the phosphate ion as summarised in Fig. 13. Two involve loss of alkene, either equation (a) by direct cleavage of O-C bonds to form carbocations [16, 151] or equation (b) by extraction of a H atom from the second carbon atom of the alkyl group [14]. In the third, equation (c), there is transfer of an alkyl group from O to S as shown in Eq. 3 in the previous section [122]. Equations (a) to (c) in Fig. 13 all generate phosphate ions that may then displace OR or SR groups from a neighbouring ZDDP molecule to form a P-O-P bridge.
Fig. 13
Possible reactions to form phosphate anions during ZDDP tribofilm formation
Full size image
Unfortunately, all the above reactions are based on high temperature thermal decomposition studies and there is, as yet, no direct experimental evidence that they occur in tribofilm formation at much lower temperatures. However, one molecular modelling study has proposed that a primary ZDDP may react via equation (b) in Fig. 13 while secondary ones, that should form more stable carbocations, may undergo equation (a) [144].

5.5 Depolymerisation

Quite recently it has been determined that, while the main initial process of tribofilm formation is polymerisation to polyphosphate and even meta- and ultraphosphate, this is followed by a much slower depolymerisation of most of the film to form short chain polyphosphates and eventually pyro- and orthophosphate, driven by Zn cation take-up and then, if available, Fe(III) ions. This process results in a structural change that causes the tribofilm to become more durable during rubbing.
In 2005, Fujita et al. used MTM-SLIM to show that once thick tribofilms had formed from both primary C8 and a secondary C3/C6 ZDDP solutions, these films remained largely unchanged even when the ZDDP solution was replaced by base oil and rubbing was resumed [56]. This suggested that the films were able to withstand the rolling-sliding rubbing condition used, and that the limiting film thicknesses did not represent a balance between continuous film formation and removal. The films were partially removed when the ZDDP solution was replaced by an aminic solution but this was ascribed to chemical attack of the amine on the films, possibly involving sequestration of the Zn ions [56].
Recently, however, Parsaeian et al. found using MTM-SLIM that if ZDDP solution was replaced by base oil before a thick tribofilm had fully developed, then the tribofilm film that had formed was partially removed during subsequent rubbing, as illustrated in Fig. 14 [59]. This implied that the tribofilm formed initially was relatively soft, but that it became harder and more wear-resistant during prolonged rubbing. The authors measured the ratio of BO to NBO in the tribofilm using XPS and found that the observed increase in strength of the film during rubbing was associated with a decrease in this ratio, suggesting that it resulted from shortening of the polyphosphate chains.
Fig. 14
Tribofilm thickness results for the early stage durability test. Reprinted with permission from ref. [59]. Copyright (2017) Elsevier
Full size image
This finding was further explored by Ueda et al. in 2019, who used FIB/TEM to determine the structure of tribofilms formed by primary C8 and secondary C6 ZDDPs at various stages of their formation in MTM tests [60]. They confirmed Parsaeian’s work and also found that the initially formed tribofilm was largely polyphosphate and had an amorphous structure. However, during rubbing this gradually converted to a shorter chain, nanocrystalline and thus much more wear-resistant film. This conversion began close to the steel substrate but progressed through almost the whole film during prolonged rubbing. Based on EDX measurements on FIB samples the authors proposed that the conversion was driven largely by diffusion of iron into the film. This is supported by a study by Berkani et al. who used zinc metaphosphate particles dispersed in base oil to lubricate a rubbing ball on disc contact. They found when the ball and disc were steel, the particles were depolymerised to orthophosphate during rubbing. No such conversion took place in a sapphire/sapphire contact. However if FeOOH (goethite) particles were included in the base oil, depolymerisation of the metaphosphate occurred even in a sapphire/sapphire rubbing system [152].

5.6 Breakdown/Removal of ZDDP Tribofilms

All the above suggests that a ZDDP tribofilm after prolonged rubbing consists of a ferrous sulphide sublayer covered by pads of nanocrystalline ferric/zinc phosphate with the top of the pads, that represents material most recently formed, being a thin layer of polyphosphate.
One question still outstanding is when and how such a ZDDP tribofilm is eventually lost from the rubbing surfaces to allow wear; is it removed mechanically via abrasive/adhesive wear or fatigue, or chemically via corrosive wear, or some combination of the two? Jahanmir found that while a primary C4/C5 ZDDP reduced wear in low load, sliding contacts, it could increase wear rate compared to the base oil at very high loads [153]. He related this to an absence of phosphorus and loss of sulphur from the rubbed track and suggested that sulphide particles were delaminating from the rubbed surface. Mourhatch and Aswath examined ZDDP tribofilm surfaces after scuffing using FIB and identified regions where patches of film had been torn off [154]. Both the above studies were based on pure sliding contacts, which are of limited relevance to the rolling-sliding conditions present in components such as gears and rolling bearings. MTM-SLIM studies have shown that once a film has formed in rolling-sliding conditions it persists with thickness almost unchanged for long periods of rubbing, so long as film-removing species, such as dispersants and soot are not introduced [56]. Some caveats should be noted however. One is that no systematic, very long duration test studies appear to have been made to determine whether the ZDDP tribofilms eventually fail. A second is that most ZDDP research has used very smooth test specimens, so contact stress conditions are relatively mild at the asperity level.
It has been shown that the phosphorus compounds formed when ZDDP acts as an antioxidant are unable to generate tribofilms, so that when all the ZDDP initially present is used up a tribofilm can no longer form [9]. In this case it was found that the tribofilm thickness fell to zero and high wear ensued, promoted by the presence of corrosive sulphur species from the reacted ZDDP. However, the mechanism by which the ZDDP tribofilm was removed was not identified.
From the above it is evident that considerable further work is needed to determine the conditions at which, and mechanisms by which, ZDDP tribofilms are lost from rubbing surfaces. Clearly such information is needed to complete our understanding of the antiwear action of ZDDP and it is also required to develop reliable models of the tribofilm formation of ZDDP, as discussed later in this review.

6 ZDP

In 1979 Dorinson suggested that zinc dibutylphosphate, in which the four sulphur atoms of a ZDDP shown in Fig. 1 are replaced by oxygen atoms, was almost as effective at reducing wear as the corresponding ZDDP, zinc dibutyldithiophosphate [155]. Little notice was taken of this since the dibutylphosphate compound was considerably less soluble in base oil than the corresponding ZDDP and it also lacked the extreme pressure properties and much of the antioxidant capability of the ZDDP. However, in the 2000s, the need to reduce the level of sulphur in engine oils led to renewed interest in the use of zinc dialkylphosphates (ZDPs) [156, 157]. It was shown that when blended with appropriate mild EP additives and antioxidants they could be used to formulate effective engine oils with antiwear properties comparable to those of ZDDP-based oils. Research showed that the ZDPs formed thick tribofilms at similar rates to their ZDDP counterparts, and, like the ZDDPs, the secondary ZDPs were much more reactive than the primary ones [142, 158]. Analysis of the tribofilms formed by ZDPs using XPS also indicated that their tribofilms were based on phosphates with similar BO and NBO ratios to those in ZDDP tribofilms [143].
As noted earlier in this paper, this response of ZDPs, as well as having practical application, casts some doubt on the significance of the proposed O/S exchange mechanism during ZDDP tribofilm formation. It does not directly refute this exchange process but implies that it is not a necessary step in the formation of the polyphosphate film.

7 How does ZDDP Work?

ZDDP is an antioxidant, corrosion inhibitor, antiwear agent, and extreme pressure additive. As mentioned in the Introduction, its antioxidant properties originate from its ability to react with and decompose peroxy-radicals and hydroperoxides, both of which are key intermediates in the hydrocarbon oxidation cycle [2, 5]. The effectiveness of ZDDP in inhibiting bearing corrosion was recognised from the beginning [159], but has received little or no attention from researchers. It probably originates from a combination of ZDDP limiting the formation of corrosive organic acids and peroxides via its antioxidant response, and its formation of tribofilms that prevent access of these corrosive agents to the bearing surfaces [160].
ZDDP is commonly described as an AW/EP additive. Its extreme pressure (EP) properties are generally assumed to originate primarily from its sulphur component and thus its ability to form iron sulphide, with its antiwear capabilities coming mainly from its phosphorus and hence formation of phosphate. This assumption is, however, largely based on analogy, since S-based additives are known to have EP properties while P-based ones are more generally regarded predominantly as AW additives.
The most obvious mechanism of ZDDP reducing wear is that the formation of a thick and robust phosphate film on the rubbing surfaces prevents direct metal or metal oxide asperity contact in boundary and mixed lubrication conditions, thereby preventing adhesive wear. Also, because the film is based, at least initially, on zinc phosphate rather than iron phosphate it does not need ferrous material to form it. This is the most widely accepted mechanism but two alternative or additional mechanisms have also been suggested. One is that iron oxide wear particles are “digested” within a ZDDP tribofilm by reaction with zinc polyphosphate to form iron phosphate. This process, it is suggested, removes hard and abrasive particles that would otherwise cause abrasive wear of the surfaces [161, 162]. The second is that ZDDP, acting in its role as an antioxidant, reacts with and thus removes peroxy- free radicals that would otherwise react with ferrous surfaces to produce easily worn oxides [163, 164].
It should be noted that the considerable body of research outlined above to measure ZDDP tribofilm thickness carries the implicit assumption that the formation of thick tribofilms by ZDDPs is responsible for their excellent wear-reducing properties. However, this is by no means proven. As shown in Fig. 15, in a detailed study of many classes of potential antiwear additives in 2012, Benedet identified several additives that gave superior wear performance to ZDDP despite forming much thinner tribofilms [165]. Recently Konicek et al. noted negligible tribofilm on a valve train tappet from an engine test despite the latter suffering very low wear [166]. Based on this and parallel bench tests they suggested that wear control may not require a thick tribofilm but instead originate from non-adherent, particulate triboproducts on the surfaces. It is, of course, quite possible that in firing engine conditions ZDDPs form protective tribofilms to provide initial protection and that these are lost during prolonged rubbing, but only after the surfaces have been chemically smoothed. This type of effect, of formation of a rough tribofilm on valve train parts in engine tests followed by its loss and the generation of smooth surfaces was indeed noted in very early work on ZDDP [167]. In a recent study of the degradation of oil in a gasoline engine in service, negligible ZDDP was detected in the oil after 6000 km operation, but the used oil was still able to form a thin tribofilm in a reciprocating tribometer at 100 °C [11]. Further research to study the evolution of tribofilm (if any) on actual rubbing parts in firing engines would be very desirable.
Fig. 15
Comparison of wear coefficients determined after 16 h rubbing in reciprocating MTM with MTM-SLIM images taken after 2 h of rubbing for several antiwear additives. Reproduced from [165]. ZDDP1 is a primary ZDDP. Details of the other additives are provided in the reference
Full size image
Two recent studies have measured the scuffing performance of ZDDP solutions and compared this with the base oil. Chern et al. used a 4-ball machine with step-loading and found that ZDDP markedly increased scuffing load and that the latter increased with increase in ZDDP concentration [168]. Ueda et al. used an MTM and extreme traction machine (ETM) to study scuffing in step-sliding speed tests while also using SLIM to monitor tribofilm thickness [169]. They found that for both primary and secondary ZDDPs, to prevent scuffing the surfaces had to be run-in at relatively mild conditions to allow a tribofilm to form; if a high sliding speed was imposed without this, scuffing occurred immediately. However, once a tribofilm was able to form, the contact survived without scuffing up to very high sliding speeds. Eventually at a combination of very high sliding speed and very high contact pressure, the tribofilm gradually became thinner and scuffing occurred when the thickness fell below about 5 nm.
It thus seems that the anti-scuffing properties of ZDDP originate from a combination of the S and P moieties. The sulphur forms an underlying ferrous sulphide and may also play a role in preventing scuffing as parts of the film are lost. The phosphorus builds a relatively thick phosphate film that protects the sulphide up to more extreme conditions than would otherwise be possible.

8 Downsides of ZDDP

In addition to their deleterious impact on exhaust aftertreatment systems outlined earlier in this review, in recent years several other problems associated with using ZDDP have been noted.

8.1 Friction

It has been known for many years that ZDDP can contribute to a decrease in engine efficiency [170], and in the late 1990s friction measurements in mixed rolling/sliding conditions showed that as ZDDP formed a tribofilm this led to a progressive increase in friction at intermediate entrainment speed conditions [171, 172]. This is illustrated in Fig. 16a, which shows a series of Stribeck (friction coefficient versus entrainment speed) curves taken during a two hour MTM test using a ZDDP-containing lubricant in mixed lubrication conditions [69]. It is evident that as rubbing progresses and a tribofilm forms, a higher and higher entrainment speed is needed to enter mixed and, ultimately, full EHD film lubrication. In a crankcase engine this implies that as a tribofilm develops more and more of the piston/liner and cam-follower cycle will operate in high friction boundary and mixed lubrication. This is now realised to be due to the roughness of ZDDP tribofilms. As was shown in Fig. 3, ZDDP tribofilms have a pad-like structure, with regions of thick film separated by deep valleys where almost no tribofilm is present. This means that as the pad height grows, the surface roughness increases. Dawcyzk et al. measured Stribeck friction curves at different stages in MTM tests and used AFM to measure ZDDP tribofilm roughness at each stage, from which they determined the lambda ratio, the ratio of EHD film thickness to composite surface roughness [69]. Figure 16b shows that all the measured Stribeck curves collapse on to a single line when plotted against the prevailing lambda ratio, indicating that the increase in friction as the test progresses originates entirely from the impact of ZDDP tribofilm roughness on EHD film formation. Figure 16 shows results for a primary ZDDP but similar friction behaviour and dependence of lambda ratio was also observed for a secondary ZDDP.
Fig. 16
a Friction versus entrainment speed curves during a test with a primary ZDDP. b Corresponding friction versus lambda ratio (ratio of EHD film thickness to surface roughness) curves. Reprinted from ref [69]
Full size image
Dawycz et al. found that although for a given ZDDP the friction curves collapsed onto single curves when plotted against lambda ratio, the boundary friction coefficient (ca 0.14 in Fig. 16b) varied from ZDDP to ZDDP. Several other studies have also shown that ZDDP boundary friction depends on alkyl structure, with linear alkyl groups giving lower friction that branched or cyclic ones [142, 173, 174], analogous to organic friction modifiers.

8.2 Micropitting

Micropitting is a form of damage found in gears and rolling bearings when these operate in mixed boundary/EHD conditions. The high stresses that occur as asperities rub against one another lead to the formation of numerous, asperity-scale, shallow-angle fatigue cracks. As these cracks develop, they undermine the material above them, which is then easily removed to form tiny pits and eventually rapid wear known as micropitting wear. Two key, related drivers of micropitting are the roughness of the surfaces and the lambda ratio, and it has been shown that ZDDP can promote micropitting because it forms protective tribofilms so rapidly that these prevent smoothing of surfaces due to running-in. In consequence asperity peaks are not removed and severe asperity stresses continue during normal operation, leading to micropitting [175177]. This is illustrated in Fig. 17 where a test in secondary C6 ZDDP solution gives much more micropitting wear, measured as loss of diameter of a cylindrical specimen, than in the corresponding base oil, or, indeed, where surfaces were initially rubbed in base oil to enable running-in, followed by replacement of the base oil by ZDDP solution.
Fig. 17
Microptting wear with base oil PAO, ZDDP solution, and PAO followed by ZDDP solution. Reprinted with permission from ref. [175]. Copyright (2004) Elsevier
Full size image

8.3 Soot Wear

A third issue concerning the use of ZDDP in engine oils is that it can lose much of its antiwear ability when high levels of engine soot are present. This has grown in importance due to longer oil drain intervals, exhaust gas recirculation, and the widespread use of direct injection gasoline engines, all of which increase soot content in the lubricant. Throughout the 1970s to 2000s many possible mechanisms for this effect were suggested [178], but since the 2010s it has become widely accepted that much, if not all of it originates from a corrosive-abrasive mechanism [65, 179181]. Soot is not generally hard enough to abrade hardened steels, but it is hard enough to abrade a ZDDP tribofilm. This means that when a high concentration of soot particles is present in an oil, any ZDDP tribofilm that forms is immediately worn off. In consequence, a blend of ZDDP and soot can give much higher wear rate than the corresponding blend in which ZDDP (or soot) is absent, as illustrated in Fig. 18 [182]. Recently it has been shown that this wear process is much faster with some steel alloys than others [183], and that it can be alleviated, at least to some extent, by appropriate choice of dispersants and steel hardness [136, 184]. It has also been shown that similar corrosive-abrasive wear due to combined soot/ZDDP occurs in fully formulated heavy duty diesel engine oils during use [185].
Fig. 18
Influence of ZDDP (mixed primary/secondary), dispersant, and carbon black on mean disc wear scar diameter from HFR tests using hard discs (adapted from [182])
Full size image

9 Interaction with Other Species

ZDDPs are used in formulated oils and so must coexist and continue to be effective in the presence of other lubricant additives as well as, in engine oils, contaminants, such as soot and water. In early work, Rounds explored the impact of a range of lubricant additives on the antiwear properties of ZDDP and found that some of these, notably a rust inhibitor, an aminic friction modifier and S and Cl-based EP additives were very detrimental to ZDDP’s wear control properties [186]. Considerable research followed and in 2005 Nicholls et al. reviewed the interactions of ZDDP with other engine oil additives, especially detergents and dispersants [187]. These have continued to be of concern but, driven by the quest to reduce friction and thereby improve efficiency, there has also been growing interest in how ZDDPs respond to the presence of friction modifier additives.
Most research has interpreted the impact of other additives on ZDDP tribofilm formation in terms of the following.
(i)
Interaction of the other additive molecules with ZDDP molecules in solution or at surfaces to reduce the latters’ chemical activity (e.g. polyamine dispersants, [188]).
 
(ii)
Adsorption/reaction of other additive molecules on a ZDDP tribofilm (e.g. some OFMs, [61]).
 
(iii)
Competition between other additive molecules and those of ZDDP for access to the rubbing surface (e.g. detergents [189]).
 
(iv)
Chemical or physical participation by other additives in the formation or removal of the ZDDP tribofilm, (e.g. amines, overbased detergents, [61, 190]).
 

9.1 Interaction with Detergents

The influence of overbased detergents on ZDDP friction, wear, and tribofilm formation is complex since these detergents are colloidal, based on nano-scale particles of Ca or Mg carbonate or hydroxide base. In the absence of ZDDP the latter accumulate on the rubbed tracks in both rolling [191] and rolling-sliding [192] conditions to form solid tribofilms that can reach more than 100 nm thickness. In rolling-sliding conditions they may form such films considerably faster than ZDDP. This means that as well as chemically interacting with ZDDP, overbased detergents can complement or compete with ZDDP’s tribofilm formation at a purely physical level. Thus XANES and XPS analysis has indicated that the Ca from Ca sulphonate and salicylate is present in ZDDP tribofilms both as calcium phosphate and as CaCO3 and that when detergent is present, ZDDPs form shorter polyphosphate chains [190, 193196]. In terms of chemical effect, Huq et al. showed that overbased detergents increased the thermal decomposition temperature of ZDDP [197], while Yu et al. found that they reduced the rate of thermal ZDDP film formation [198]. Most studies have found that overbased detergents degrade the antiwear performance of ZDDPs, and in the past most research has focussed on Ca-based detergents with limited study of Mg-based ones. The growing importance of the latter may change this and recent work has suggested that Mg detergents may damage ZDDP tribofilms more than Ca-based ones [199]. In terms of friction, early work suggested that a Ca-sulphonate detergent increased the boundary friction of ZDDP tribofilms [200] but it has subsequently been found that the friction of sulphonate detergents depends very strongly on the linearity of their alkyl chains [201], so this deleterious response may not occur with linear alkyl chain detergents.

9.2 Interaction with Dispersants

There has been considerable research on the interaction of succinimide polyamine dispersants with ZDDPs [187], mostly using IR and 31P NMR analysis to detect changes in P-S bonds and the P electronic environment. The structure of these is illustrated in Fig. 19 from [124]. They are the most commonly used engine oil dispersants and act to sterically stabilise colloidal soot particles.
Fig. 19
Structures of succinimide polyamines a mono- and b bis-succinimide. PIB is a polyisobutene chain. From ref [124]
Full size image
Early work showed that succinimide molecules interact in solution with ZDDP molecules [202] and it soon became evident that Zn ions combine with NH and NH2 groups of the dispersant to form coordination complexes [188, 203205]. Kapur et al. showed that this association was stronger with mono-succinimides, that have free NH2 groups, than bis-succinimides, and that primary ZDDPs interacted more strongly than a secondary one [182]. This complexation, it has been suggested, may explain why dispersants reduce the rate of ZDDP tribofilm formation [206]. However more recently, MTM-SLIM has been widely used to monitor the impact of dispersant on ZDDP tribofilms [56, 62, 124, 178], and, while this has confirmed that dispersants do significantly reduce the rate of tribofilm formation, it has shown that dispersants also actively and very rapidly remove pre-formed ZDDP tribofilms during rubbing. As can be seen in Fig. 20, only part of the tribofilm is removed and, based on recent understanding of the evolution of ZDDP tribofilms during rubbing [60], it is likely that the dispersant removes Zn-based polyphosphate film by complexing out Zn ions, but leaves any underlying Fe phosphate film intact. Recently Tabibi has explored the influence of a range of succinimide structures and found quite large differences in their effect on ZDDP, confirming the difference between mono- and bis-succinimides [124].
Fig. 20
Primary ZDDP tribofilm formation and removal by succinimide dispersant. Reprinted with permission from ref. [56]. Copyright (2005) Taylor & Francis
Full size image
From the above would appear difficult to disentangle the two main mechanisms by which dispersants might suppress ZDDP tribofilm growth; does the dispersant tie up the ZDDP molecules in solution or at surfaces, rendering them less reactive and thereby reducing tribofilm growth rate? Or does the dispersant simply extract Zn from the tribofilm as the latter forms, weakening it, or even preventing its growth altogether? Or, indeed, both?
Although there has been much study of the interactions between dispersants and ZDDPs, there has been surprisingly little published on the actual impact of dispersants on wear rate. Zhang et al. showed that the addition of dispersant could increase wear rate by a factor of between 2 and 8 depending on dispersant type and concentration [62]. Of course, in the context of engine oils, the intrinsic influence of dispersants on wear may well be minor compared to their influence on the wear promoted by soot particles, since it has been found that some dispersant types are much more effective at suppressing soot wear than others [136].

9.3 Interaction with Friction Modifiers

In 2007 Topolovec et al. explored the effect of a range of friction modifier additives on ZDDP friction and tribofilm formation in pure sliding and rolling-sliding contact conditions [61]. They studied both the impact of FMs on secondary ZDDP tribofilm growth rate and the impact of FMs on pre-formed ZDDP tribofilms. For surfactant-type organic friction modifiers (OFMs), response was very varied; OFMs having basic functional groups, such as amine and amide were effective in reducing boundary friction while others were much less so. However, the amine and amide both supressed ZDDP film formation and in the case of amine, rapidly removed the tribofilm. This is reminiscent of the impact of succinimide dispersant on ZDDPs and suggests that basic nitrogen functionalities, while adsorbing strongly on a ZDDP tribofilm, may also weaken it. Matsui et al. studied the impact of organic friction modifiers as well as functionalised polymers and a succinimide dispersant on primary C8 ZDDP tribofilm formation and friction [207]. As shown in Fig. 21, electron probe microanalysis was used to measure the elemental composition of the rubbed surfaces and showed that different OFMs had a quite selective effect on the composition of the tribofilm with some, notably those with acid and alcohol groups, reducing sulphur content, while those with aminic groups reducing phosphorus and zinc. Matsui et al. also used 31P NMR to assess the interaction of the OFMs with ZDDP in solution and found that additives with acid groups increased the neutral to basic ZDDP ratio in solution while those with amino groups had the opposite effect. It is thus possible that some of the effect of other additives on ZDDP might originate from their influence on the chemical nature of the ZDDP in solution.
Fig. 21
Chemical images of phosphorus, sulphur, oxygen, and zinc obtained by EPMA after the triboexperiments with the optical microscope images at the same position. Reprinted with permission from ref. [207]. Copyright (2018) Taylor & Francis
Full size image
Several other recent studies have compared the effect of different amine structures on tribofilm build up and removal [57, 208, 209], but findings have been contradictory, with some research showing reductions in film formation while others not. All studies have, however shown that amine surfactants are very effective at reducing boundary friction of ZDDP tribofilms. As well as amines, some other OFMs and also polymeric friction modifiers have been shown to reduce boundary friction in combination with ZDDP, without overly inhibiting tribofilm formation [210, 211].
Since the 1990s, driven by engine fuel economy requirements, there has been a great deal of research concerning the frictional properties of combinations of ZDDP with the friction modifier additive molybdenum dialkyldithiocarbamate (MoDTC) [212]. MoDTC reduces friction by forming nanocrystals of low shear strength MoS2 on rubbed surfaces and tends to be relatively indifferent to the solid surface on which it forms, including ZDDP tribofilms [212]. FIB/TEM analysis of surfaces rubbed in solutions containing both ZDDP and MoDTC has shown the presence of MoS2 nanocrystals both within and close to the surface of ZDDP tribofilms [213, 214] and MoDTC has been found to reduce friction both when blended with ZDDP and when a pre-formed ZDDP tribofilm is rubbed in MoDTC solution [61]. While MoDTC is able to form MoS2 and reduce friction in the absence of ZDDP, it has been suggested that the presence of ZDDP helps protect the MoS2 from wear, limits the formation of MoO3, sulphurises MoDTC, and extends the useful life of MoDTC by acting as an alternative peroxide decomposer [132, 133, 213, 215, 216]. One Raman-based study has found MoS2 formation by MoDTC only when ZDDP was present [71]. All of the above describes the effect of ZDDP on MoDTC; in terms of the opposite, MoDTC appears to produce a slight reduction in the thickness of tribofilm formed by ZDDPs, leading to partial loss of ZDDP antiwear performance [215, 216].
While ZDDP appears to improve the friction-reducing performance of MoDTC, or at least its longevity, its effect on other sulphur-free Mo-based additives is much more striking. In principle, without any sulphur content the latter cannot form MoS2, but it has been found that when ZDDP is present, they are able to form MoS2 and produce a consequent reduction in boundary friction [217, 218]. It appears that the sulphur atoms present in ZDDP are sufficiently labile that they can contribute to MoS2 formation.

9.4 Interaction with Water

As well as other lubricant additives, ZDDP also encounters and must retain effectiveness in the presence of two important contaminants, engine soot and water. The levels of both these have been increasing in recent years, the latter due to growth in biofuel use and the introduction of hybrid engines that often operate intermittently and at relatively low temperatures. The impact of soot on ZDDP was outlined in the previous section. Systematic study of the influence of water on ZDDP tribofilm formation is relatively recent. In 2011 Nedelcu et al. compared the behaviour of a primary C4 ZDDP with and without low levels of added water (0.5 to 2%) in a rolling four-ball tester and found that the water reduced tribofilm thickness and also the length of polyphosphate chains in the film [219]. This was extended by Cen et al. who used a sliding pin-on-disc tribometer and showed that increase of water level, either from an increase in relative humidity (RH) or by adding water to the lubricant directly, reduced tribofilm thickness and also increased wear rate [220]. Several other researchers have come to similar conclusions [221223]. Parsaeian et al. used MTM-SLIM to show that increase of RH did not significantly affect the rate of ZDDP film growth but rather the thickness at which the ZDDP tribofilm stabilised [221]. Costa et al. compared the response of blends of anhydrous and hydrated ethanol (the latter contains 5% water) in both a solution of ZDDP in base oil and a commercial engine oil containing secondary ZDDP. They found that both ethanols removed pre-formed tribofilm, but the hydrated ethanol did so much faster and more completely [224]. Various mechanisms have been suggested for the deleterious effect of water on ZDDP film formation and wear, including water molecules blocking the surface, direct corrosive-abrasive wear of steel by water, and Fe ions produced by corrosion depolymerising the tribofilm.
Overall, from the above it appears ZDDP is remarkably tolerant of the presence of other additives at the concentrations generally used in engine oil, perhaps just as well considering the melange of species present in crankcase engine oils. The main exception appears to be some additives containing primary aminic groups, though even here ZDDP appears to function effectively so long as the molar ratio of amine to ZDDP remains low [208].

10 Non-ferrous Surfaces

Until quite recently almost all research on ZDDP studied its interaction with ferrous metal alloy surfaces. Indeed, most research was limited to ZDDP’s response on the 1% Cr steel AISI 52100 (SUJ2, 100Cr6, 535A99), since this is the most common ball-bearing steel and many friction and wear bench tests require the use one or more ball-shaped test specimens. A question that obviously arises is whether ZDDP is equally effective on other types of metal or, indeed, ceramic surfaces?

10.1 ZDDP on Non-ferrous Metals

A recent study compared the film-forming behaviour of antiwear additives on four different ferrous alloys and concluded that a mixed primary/secondary ZDDP’s response (though not that of ashless antiwear additives) was very similar regardless of the alloying elements present [225]. This is, perhaps, not surprising if we believe that the initially formed polyphosphate tribofilm is based largely on zinc cations, with iron cations coming along later, as discussed in Sect. 5 of this review. Ito et al. also showed that a secondary C6 ZDDP formed tribofilms readily on oxidised steel surfaces consisting predominantly of Fe3O4 [92], and most ferrous alloys will have some iron oxide at the surface. A few previous studies have looked at the behaviour of ZDDP on non-ferrous metals. These have been mainly been confined to measuring adsorption [96, 97] and thermal film formation [187], probably because the hardness of different metals varies greatly, and makes comparison of their tribological interactions with ZDDP problematic. This was addressed recently by Ueda et al. who used ion implantation to produce thin films rich in a range of non-ferrous metals on steel surfaces; so thin as not to affect the overall stiffness or hardness and thus contact pressure [67]. They found that secondary C6 ZDDP formed tribofilms faster on Ni-rich surfaces than Fe-rich ones but formed films only slowly on Cr- and Mo-rich surfaces. The rate of film formation on Cr and Mo surfaces increased with ZDDP concentration, suggesting that the differences originated from different levels of ZDDP adsorption on the various metallic or metal oxide surfaces. Other researchers have also noted that ZDDP adsorbs less easily on Cr-rich surfaces [100]. Using fully formulated engine oils Zhu et al. compared the films formed on cast iron and Cr-plated cylinder liners rubbed against CrN -coated rings [226] in a fully formulated engine oil and found S/P/Zn-containing films on the cast iron cylinders but none on the plateaux of the Cr-plated liners.

10.2 ZDDP on Ceramics

The ability of ZDDP to form tribofilms on engineering ceramics, such as Si3N4 and WC has also been investigated. Most studies employed a ceramic tribo-element rubbing against a ferrous one, which, while of practical relevance, makes it difficult to isolate the behaviour of ZDDP on the ceramic itself because of the possibility of transfer of metal ions or tribofilm from one surface to the other. However, Sheasby et al. studied the behaviour of a secondary ZDDP in Si3N4/ Si3N4 and ZrO2/ZrO2 contacts and noted the presence of a pad-like film on Si3N4, comparable to the formed on steel [227]. In a recent study, Ueda et al. used an MTM to study Si3N4, SiC and WC ball and disc tribopairs [125]. They found that a secondary ZDDP formed a thick ZDDP tribofilm on the Si3N4 and WC surfaces in both mixed and full EHD film conditions, but that negligible film was generated on SiC. The films formed on Si3N4 and WC were, however, quite weak compared to those formed on steel, being easily removed by rubbing in base oil. They concluded that, while ZDDP could, in principle, form tribofilms via mechanochemical action on all surfaces, in the absence of ferrous ions this film remained amorphous and thus relatively soft. The lack of a film on SiC was ascribed to very limited adsorption of ZDDP on this essentially metal-free material.
In the context of engine lubrication, two surfaces of strong interest in recent years have been hypereutectic Al/Si alloy, used to cast lightweight engine blocks, and diamond-like carbon (DLC) coatings that are becoming ubiquitous on many machine components including those in crankcase engines.

10.3 ZDDP on Al/Si

Hypereutectic Al/Si alloys contain typically 16–18% Si (more Si than the eutectic level of 12.6%), with the pro-eutectic silicon dispersed in Al/Si eutectic in the form of large Si particles up to tens of micron across. These Si particles are very hard compared to the matrix and so provide surfaces that are wear-resistant enough to be used directly as tribologically active aluminium cylinder surfaces, without the need for separate steel liners. In 2005 Nicholls et al. used AFM and spatially resolved XAS to study the tribofilm-forming properties of a mixed primary/secondary (C8/C4) ZDDP on hypereutectic Al/Si alloy rubbing against a steel counterface [228]. They found polyphosphate present initially on both Si particles and the surrounding matrix, but that the eutectic matrix was rapidly worn away to leave protruding Si particles. ZDDP tribofilm was present as pads on the Si particles, just as on steel, with a similar polyphosphate composition. These findings have been broadly confirmed by several other studies [229231]. One question that arises is whether a ZDDP tribofilm can form on Al or eutectic matrix, both of which are relatively soft. The above studies certainly identified tribofilm on the matrix, but it is possible that this was transferred from the Si particles or the ferrous counterface. This was resolved by Gosvami et al. who used an AFM to rub an Al2O3 colloid probe against an Al flat and observed tribofilm formation from a mixed primary/secondary ZDDP on both surfaces. They suggested that extensive plastic deformation and consequent strain hardening of the soft Al enabled local stresses to become high enough to promote mechanochemical reaction of the ZDDP [44].

10.4 ZDDP on DLC

Research on the behaviour of ZDDP on DLC surfaces is complicated by the fact that there are several different structures of DLC coatings in use. Distinction must be drawn between those containing largely graphitic, sp2 C–C bonding (e.g. a-C) as compared to those based mainly on diamond-like sp3 bonding (ta-C), and also between DLCs that are largely hydrogenated (e.g. a-C:H) and those that are not (e.g. a-C). In addition, some DLCs contain significant proportions of atoms other than carbon, for example Si, W, and WC. As with ceramics, it is also important to distinguish between studies where DLC is rubbed against steel, where Fe ions or ZDDP tribofilm might transfer from steel to the DLC counterface, and studies of DLC against DLC.
In 2008 the behaviour of DLC/DLC lubricated by ZDDP solution was studied in both sliding [232] and rolling-sliding [192] contact. In both studies, patchy tribofilms were observed though, unlike the films formed on steel, these were easily rinsed off by cyclohexane [232]. A systematic study of the behaviour of lubricant additives on a wide range of types of DLC surface was carried out in 2011 by Vengudusamy et al. [66]. Like previous studies, he found that ZDDP formed a patchy film on most types of DLC surface but a pad-like structure similar to that formed on steel was only observed on DLCs containing a significant proportion of W, and was present only on the W inclusions [233, 234]. Different types of DLC gave very different wear behaviour in both DLC/DLC and DLC/steel contacts as shown in Fig. 22 [66].
Fig. 22
Influence of primary ZDDP on wear of various DLCs in a DLC/DLC and b DLC/steel tribopairs. Reprinted with permission from ref. [66]. Copyright (2011) Elsevier
Full size image
Vengudusamy also assessed the durability of the ZDDP tribofilms by rubbing pre-formed films in base oil and found partial film removal in all cases [235]. Interestingly he found that, unlike in ZDDP tribofilms on steel, most of the sulphur in the films on DLC was close to the upper surface, with almost none at the DLC/tribofilm interface, where instead phosphate was present.
Kalin et al. studied the mechanical properties of ZDDP tribofilms formed on DLC surfaces rubbed against steel [79, 236]. They found that these films had a dual structure with some regions that were hard and stiff, while others were extremely soft.
Okubo et al. studied the influence of ZDDP alkyl structure on friction and tribofilm formation on two types of DLC surface in both DLC/DLC and DLC/steel contacts [237, 238]. As with steel, they found that secondary ZDDPs formed tribofilms much faster than primary ZDDPs on DLC and the effect of alkyl structure on boundary friction was also similar, with linear chains giving lower friction than branched ones. They found a difference in behaviour of a highly basic and a more neutral ZDDP on one of the DLCs. Ueda et al. also studied a DLC/DLC tribopair [125] and found no measurable ZDDP film formation in either mixed or full film rolling/sliding conditions, nor any adsorption of ZDDP on a carbon coating using QCM.
From the above, it appears that ZDDP can form a tribofilm on a range of material surfaces so long as (i) the ZDDP adsorbs on the surface and (ii) the contact pressure and thus shear stress is large enough to induce a mechanochemical reaction. However, if no metallic ions are present, as is the case of most DLCs, the film remains relatively soft and easily worn since it is based only on the initially formed amorphous zinc polyphosphate and cannot progress to a more crystalline phosphate. Interpretation is confused by the fact that many studies rubbed a DLC coated surface against a steel one and then studied the tribofilm formed on the DLC, and it is not possible to discount the possibility that tribofilm formation was controlled by metal or metal oxide transferred to DLC from the steel. The role of the S in ZDDP during its tribofilm formation on non-metallic surfaces does not appear to have been much, if at all, explored.

11 Modelling

Research on modelling ZDDP can be divided into two types, macro-scale modelling of ZDDP tribofilm formation and removal and molecular-scale modelling of ZDDP chemical reactions.

11.1 Macro-scale Modelling

The first model of ZDDP tribofilm formation appears to have been by So and Lin in 1994 [107]. They assumed that ZDDP film formation was a balance between film formation and removal and that the rate of formation was controlled by a combination of Arrhenius kinetics and diffusion, while film removal was due to asperity indentation and consequent scaping away of film material. This early work was constrained by lack of experimental data on the kinetics of ZDDP tribofilm growth, but the development of the SLIM method described in Sect. 3.1 has made such data quite readily accessible and, in recent years, has supported the development of a series of models [239248]. All these incorporate; (i) a solid contact model to determine the pressure (and thus the shear stress and temperature) across a contact and (ii) expressions for film formation and film removal rate. Tribofilm thickness then evolves across the contact during rubbing over a series of time steps from the difference between film growth and film removal rates.
Most models have employed a numerical contact mechanics model with elastic-perfectly plastic deformation, although recently Lyu et al. used an analytical rough surface model [248]. Most studies have assumed boundary lubrication, with all the applied load supported by contacting asperities, but a few have included fluid film pressure [243, 247]. Initial work in 2012 that preceded general appreciation of the importance of mechanochemistry was based solely on flash temperature-controlled, Arrhenius film formation [239], but all subsequent work has included a stress-activated film growth model. Generally this has used the Eyring equation (Eq. 1), but one group [240, 243, 245] have employed a model by Bulgarevich based on an inverted Boltzmann assumption, where the surface atoms are activated by rubbing to much higher energy than those in the subsurface [249].
As described earlier in this review, experimental work shows that ZDDP tribofilm is generally initially rapid and then levels out to an asymptotic value, but that an overshoot is seen where some of the film is suddenly lost, if the film growth rate is initially very large. To develop models that show a similar response, various approaches have been adopted. One is to use a film growth rate equation where the rate diminishes steadily to zero as a critical film thickness is approached [239, 240]. Another is simply to impose a critical film thickness at which film growth abruptly stops [247]. Alternatively, the rate of film removal can been required to increase with film thickness [248]. Many models [239, 240, 242, 243, 248] also assume that the hardness decreases linearly away from its interface with the substrate, as suggested by one experimental work [77]. To achieve the temporary overshoot in film thickness seen in some experiments, one study has employed a time lag in the film growth response [242].
Most researchers have assumed that tribofilm removal rate follows an Archard wear equation, though two models have been based on a removal rate equation developed experimentally [240, 243], A few models have also allowed wear of the metallic substrate to occur [240, 248].
Overall, the various models have been found to predict the rate of average ZDDP tribofilm growth quite well, as might be expected since they have all been based largely on equations and rate constants derived experimentally. Most also show the expected growth of a pad-like structure on pre-existing surface asperities due to stress activation. However, few, if any, of the models so far developed seem reliable enough to teach us much about ZDDP tribofilm formation beyond information that is already provided from experiment. One study has shown a possible impact of different forms of surface roughness on tribofilm formation, but this still needs to be experimentally confirmed [245].
The main stumbling-block in developing useful models of ZDDP film formation appears to lie in our as-yet limited understanding of both the tribofilm removal process and the evolving nature of the tribofilm mechanical properties during rubbing. Until these have been fully resolved and related to the contact conditions and initial ZDDP structure, macro-scale models may remain of little practical value.

11.2 Molecular Modelling

Progress in modelling the tribofilm-forming behaviour of ZDDP at the atomic and molecular level has been controlled to a very large extent by progress in developing quantum mechanics approximations to describe the electronic processes involved, together with the growth in performance of computer systems required for such modelling. Two recent reviews provide general backgrounds to the application of computational molecular modelling to tribochemistry [250, 251]. Martini et al. identified two main approaches, one based on density functional theory (DFT) that considers the electronic structure of whole systems consisting of atoms and molecules, and the other reactive force fields that focuses on bonds between individual pairs of atoms. They concluded that there was still a lack of useful force fields for ZDDP. This is unfortunate since modelling using reactive force fields is much faster in terms of computing time than DFT, and so can be included in molecular dynamics simulations (MD), where the motion of a large ensemble of molecules of interest is simulated over time based on intermolecular interactions and Newtonian mechanics. Reactive force fields can then determine whether bonds break or form in response to these interactions [252]. With DFT, individual molecules and pairs can be simulated using molecular dynamics, but it is generally impractical to consider larger ensembles.
ZDDP was first modelled in the mid-1990s using ab initio quantum chemistry to test possible structures and simulate the adsorption of ZDDP molecules on surfaces [253, 254]. This confirmed the structures of both neutral and basic forms determined by crystallography and spectroscopy, and that the two P-S bonds in the dithiophosphate are equivalent. Jiang et al. compared the bonding energy of different alkyl ZDDPs on iron oxide and correlated this with their wear performance [255].
In the early 2000s, Mosey and Woo applied quantum chemistry (QC) and DFT to model ZDDP reactivity, including the number of S atoms coordinated with Zn, the monomer/dimer equilibrium and P-O and O-C cleavage processes [144, 256, 257]. Initially they proposed homolytic C-O bond cleavage to form free radicals [256], but subsequently modified this to heterolytic cleavage to form ions [144]. They evaluated the energetics of the various reaction mechanisms developed from thermal decomposition studies which were summarised in Fig. 13 earlier in this review and concluded that the most likely reaction based solely on energetics was unimolecular Ei elimination for primary ZDDPs and that alkyl transfer from O to S was unlikely at low temperatures.
In 2005 Mosey et al. applied ab initio molecular dynamics (AIMD) to model the structural response to pressure of a triphosphate molecule together with one or two zinc phosphate molecules in a cell [258]. Their simulation found that the number of O atoms coordinated with a Zn atom grew from between 2 and 4 at low pressure to 6 at a pressure of 17 GPa and that this resulted in a large increase in bulk modulus and hardness. The authors suggested that the implied pressure-induced cross-linking was practically important and could explain why ZDDP was relatively ineffective on soft metals, such as aluminium. Various researchers questioned whether such very high pressures were possible in metal–metal contacts and a series of experimental studies followed using IR, Raman, and XPS to observe zinc phosphate at very high pressure [259261]. Some evidence of irreversible structural distortions of the phosphate structure at high pressure was found but none of an increase in coordination number, or indeed of polyphosphate chain length. As well as identifying cross-linking via the zinc atoms as described above, Mosey and Woo also used ab initio DFT to model the interactions between metathiophosphate molecules, which they proposed were formed by high temperature decomposition of ZDDP [262]. They found that these all dimerised readily but that one structure, RSP(= O)2, could link together to form polyphosphate. They suggested that this pointed towards a pathway for the formation of polyphosphates from ZDDP.
Onodera et al. modelled the behaviour under shear of a zinc metaphosphate film and how this interacted with small particles of iron oxide [263265]. The objective was to test the conjecture that ZDDP tribofilms reduce wear by digesting abrasive iron oxide. Initially they used classical molecular dynamics simulation but later augmented this with quantum chemical MD to model the chemical interactions between phosphate and metal oxide more realistically. They concluded that iron and some other metal oxides could be digested by phosphate under a combination of high pressure and shear.
In an ingenious paper in 2009, Onodera et al. combined MD with finite element analysis to study diffusion of iron atoms from an Fe2O3 substrate into a ZDDP tribofilm and the effect of this on the mechanical properties of the film [266]. They found that when the tribofilm was subject to shear force, iron atoms diffused rapidly to form a graded composition in the tribofilm, in agreement with experimental work on ZDDP tribofilms described in Sect. 5.4 [60]. The researchers then derived a relationship between iron concentration and elastic properties using quantum chemistry-based MD simulations and employed this to determine the elastic modulus and hardness profile through the tribofilm. As might be expected, iron atom diffusion resulted in a harder and stiffer interfacial region.
In 2021, Salinas Ruiz et al. carried out a combined experimental and QC modelling study of the behaviour of ZDDP in contacts between various different types of DLC/DLC [267]. Their modelling work found that for hard ta-C DLC, the very high pressure and shear stress present at asperities resulted in sulphur from the ZDDP penetrating and weakening the DLC. This did not occur for less hard DLCs. For all DLCs, the ZDDP molecules were found to break down into atoms and small fragments within the sliding contact.
As outlined in Sect. 5.1 earlier, the first stage in ZDDP tribofilm formation must be surface adsorption, and early experiments have revealed intriguing processes, such as loss of zinc atoms into the solution. Recently DFT has been applied to investigate this initial stage. In 2022 Peeters et al. explored the adsorption of ZDDP molecules on Al/Mg alloy surfaces [268]. They confirmed that ZDDP’s sulphur atoms bond to the Al substrate but found that, while ZDDP adsorbed on pure Al and Mg surfaces, it only dissociated under pressure to release dithiophosphate units on a Mg17Al12 alloy surface.
Very recently, Benini et al. have applied DFT to model the role of adsorption and consequent effect of shear stress on ZDDP reaction on iron, partially oxidised iron, and iron oxide surfaces, with a focus on the role of the sulphur atoms [269]. They found that for non-adsorbed molecules, molecular dissociation was endothermic but that it became exothermic and thus favoured when ZDDP molecules were adsorbed on iron surfaces. When ZDDP molecules were adsorbed on an iron substrate, applied shear stress promoted the detachment of organophosphorus groups from the central Zn–S units, as a presumed first stage of ZDDP tribofilm formation. Interestingly, they found that surface oxidation of the iron substrate significantly reduced adsorption of ZDDP molecules on the substrate. This finding is debateable since experimental work has shown that ZDDP tribofilms form readily on an iron oxide surface, albeit with very little sulphur content [92]. However the study is important in being the first to model the initial stages of ZDDP tribofilm formation on ferrous surfaces in a systematic fashion.
Overall, in the author’s view, atomic and molecular modelling of ZDDP reactions to date are tantalising and serve to confirm, as least in part, deductions from experimental work. However, they do not appear to be reliable enough to predict beyond what is already surmised. In particular, they are not yet able, or at least have not been applied, to explore reaction pathways beyond one or two steps. In the ideal world we would enjoy a full simulation of the sequence of reactions by which ZDDP molecules adsorb, react initially to form a sulphide film, and then progress to form thereon a polyphosphate tribofilm. Perhaps one day this will be possible …

12 Some Outstanding Questions

It is evident from the above that our understanding of the role of ZDDP in controlling wear and scuffing has advanced greatly in the last 20 years, especially with respect to our appreciation of the role of mechanical stress on tribofilm formation. Based on the latter we now have reliable equations that describe the rate of formation of tribofilms.
This review reveals three main areas where we still have very limited understanding of ZDDP’s behaviour. One concerns the first stage of reaction of ZDDP with rubbing surfaces to form sulphides. This is important with respect to ZDDP’s extreme pressure properties but also may control if and how ZDDP forms, or fails to form, films on non-ferrous surfaces. We know that a ZDDP molecule adsorbs on a metal surface via its S atoms; we know it forms a sulphate/sulphide film very rapidly; but the molecular mechanism of this formation and indeed, whether this reaction varies with ZDDP structure is not yet known. Further work on this initial film formation, possibly using in situ spectroscopic methods, such as Raman and XPS with different ZDDP structures might help us learn more about this.
A second gap in our knowledge concerns the precise molecular reaction sequence by which ZDDP molecules combine to form a polyphosphate structure and how and when they lose most of their sulphur atoms. Our current understanding was outlined in Sect. 5 and is based almost entirely on very early thermal decomposition studies. Attempts were made at the time to relate this to wear performance, but this was a relatively crude instrument compared to our current ability to monitor film growth rate using AFM and SLIM. There is clearly scope for revisiting work on the impact of ZDDP molecular structure of tribofilm growth using modern methods of surface film monitoring and analysis, combined with advanced molecular modelling. Ideally, we might even detect low MWt species, such as alkenes and alkyl sulphides emitted as a tribofilm develops, as was done in thermal decomposition studies. However, the very small quantities involved makes this extremely challenging.
The above gaps in our knowledge are predominantly of academic interest, though work to fill them may have benefits in terms of ZDDP molecular design. A third, much more practically important area where research is needed is to understand the mechanical process or processes by which ZDDP films are lost to eventually fail. Do they fail by fatigue, by abrasion or some other process? Are they still present on rubbing surfaces in engines after extended service – and if so, how do they compare with freshly formed ZDDP tribofilms? Such understanding is needed to build useful models of ZDDP tribofilm formation/removal but also to help design formulations that optimise ZDDP performance in the various applications in which ZDDP is currently employed, especially if longer lubricant lives are desired.
There is also clearly much scope for molecular-scale modelling simulation to take place alongside experiment, using ZDDPs of different structure and perhaps also ZDPs, to help deduce the different stages of ZDDP tribofilm formation, from adsorption through to phosphate film generation.
Finally, it should be noted that recent studies of ZDDP tribofilm growth rate and chemical and mechanical properties have highlighted considerable differences depending on ZDDP molecular structure, both with respect to primary and secondary structure, and also alkyl chain size. This emphasises how important it is for publications on ZDDP research to state the actual alkyl structure of the ZDDPs studies. In this review, the nature of the ZDDP used, and especially whether it is primary, secondary, or mixed, has been included where this is available and may be relevant. But sadly, far too many publications still fail to include this crucial information. This not only greatly detracts from their value, but also makes it impossible to test and verify any findings described.

13 Conclusion

This paper has reviewed our current understanding of the lubricant additive ZDDP, with a particular focus on research that has been carried out over the last 20 years and on the ability of ZDDP to form wear-reducing tribofilms. This research has taken place in the context of reduced ZDDP concentrations in engine oils and ever more severe contact conditions present in lubricated machine components due to widespread reductions in the viscosities of liquid lubricants in use.
A key advance over this period has been our discovery that ZDDP tribofilm formation is driven by applied stress as well as by temperature and is thus mechanochemical in origin. This explains many of the observed features of ZDDP tribofilms including their ability to form at low temperatures, the fact that they form only on rubbing surface and their high roughness. We have also learnt that the composition and thus mechanical properties of ZDDP tribofilms change during prolonged rubbing so that they become harder and more resistant to wear. Several disadvantages of ZDDP as a lubricant additive have also been highlighted in the last 20 years, their deleterious effect on friction, their promotion of micropitting wear, and high levels of wear produced when both ZDDP and soot are present in an engine lubricant. The origins of all of these have been resolved by research.
ZDDPs are now being used with a wider range of tribological surfaces than was common 20 years ago, especially with Al/Si and DLC surfaces, and research has shown that their behaviour on these surfaces is quite different from that on ferrous surfaces. Indeed, it is possible that if these surfaces had predominated from the outset, ZDDP might never have become as important an additive as it now is.
There are still important outstanding questions concerning the tribological behaviour of ZDDP. These include the precise molecular reactions that occur during tribofilm formation and how these are influenced by ZDDP molecular structure, and the mechanisms by which ZDDP tribofilms eventually fail.
To date, almost all the recent advances described in this review have come from experimental research using in situ methods and/or advanced surface analysis methods. Some computer-based modelling research has been carried out both at the macro-scale and to explore ZDDP’s chemical reactions in rubbing conditions at a molecular level. This has provided some valuable insights but, in the author’s view, it has not yet fully fulfilled its potential promise. Perhaps, in future, a coordinated, combined experimental and molecular modelling approach using different ZDDP structures offers the best route to revealing fully the tribological behaviour and properties of ZDDP.
Finally in view of the energy transition and since the predominant use of ZDDP is in crankcase engines, can we assume that ZDDP research will continue at its current very rapid pace in future? The answer in the short term is probably yes, since ZDDP is almost certain to be the antiwear additive of choice in hydrogen-fuelled crankcase engines. It is also generally being employed as a reference additive in the development of new antiwear additives for other applications, such as electric vehicle (EV) transmission oils. However, unless ZDDP can somehow be made to function reliably in EV oils, which is very unlikely, the author fears that funding for further research on ZDDP may dwindle in the next decade and, sadly, some of the above outstanding questions may never be fully resolved.

Acknowledgements

The author thanks Dr Ksenija Topolovec-Miklozic for the excellent AFM image of a ZDDP tribofilm used as the Graphical Abstract.

Declarations

Competing Interests

The authors declare no competing interests.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literature
1.
go back to reference Spikes, H.: The history and mechanisms of ZDDP. Tribol. Lett. 17, 469–489 (2004) Spikes, H.: The history and mechanisms of ZDDP. Tribol. Lett. 17, 469–489 (2004)
2.
go back to reference Burn, A.J.: The mechanism of the antioxidant action of zinc dialkyl dithiophosphates. Tetrahedron 22, 2153–2161 (1966) Burn, A.J.: The mechanism of the antioxidant action of zinc dialkyl dithiophosphates. Tetrahedron 22, 2153–2161 (1966)
3.
go back to reference Howard, J.A., Ohkatsu, Y., Chenier, J.H.B., Ingold, K.U.: Metal complexes as antioxidants.I. The reactions of zinc dialkyldithiophosphates and related compounds with peroxy radicals. Can. J. Chem. 51, 1543–1553 (1973) Howard, J.A., Ohkatsu, Y., Chenier, J.H.B., Ingold, K.U.: Metal complexes as antioxidants.I. The reactions of zinc dialkyldithiophosphates and related compounds with peroxy radicals. Can. J. Chem. 51, 1543–1553 (1973)
4.
go back to reference Newley, R.A., Spikes, H.A., Macpherson, P.B.: Oxidative wear in lubricated contact. Trans ASME J. Lubr. Technol. 102, 539–544 (1980) Newley, R.A., Spikes, H.A., Macpherson, P.B.: Oxidative wear in lubricated contact. Trans ASME J. Lubr. Technol. 102, 539–544 (1980)
5.
go back to reference Yagishita, K., Igarashi, J.: 31P NMR and mass spectrometric studies of the reaction of zinc dialkyldithiophosphates with cumene hydroperoxide (Part 1) kinetics and mechanisms of the initial homolytic reaction. J. Japan Pet. Inst. 38, 374–383 (1995) Yagishita, K., Igarashi, J.: 31P NMR and mass spectrometric studies of the reaction of zinc dialkyldithiophosphates with cumene hydroperoxide (Part 1) kinetics and mechanisms of the initial homolytic reaction. J. Japan Pet. Inst. 38, 374–383 (1995)
6.
go back to reference Wang, L., Zhang, D., Wu, H., Xie, Y., Dong, G.: Effects of phosphorus-free antioxidants on oxidation stability and high-temperature tribological properties of lubricants. Proc. Inst. Mech. Eng. Part J 231, 1527–1536 (2017) Wang, L., Zhang, D., Wu, H., Xie, Y., Dong, G.: Effects of phosphorus-free antioxidants on oxidation stability and high-temperature tribological properties of lubricants. Proc. Inst. Mech. Eng. Part J 231, 1527–1536 (2017)
7.
go back to reference Jin, Y., Duan, H., Cheng, B., Wei, L., Tu, J., Liu, J., Li, J.: Synthesis of a multi-phenol antioxidant and its compatibility with alkyl diphenylamine and ZDDP in ester oil. Tribol. Lett. 67, 58 (2019) Jin, Y., Duan, H., Cheng, B., Wei, L., Tu, J., Liu, J., Li, J.: Synthesis of a multi-phenol antioxidant and its compatibility with alkyl diphenylamine and ZDDP in ester oil. Tribol. Lett. 67, 58 (2019)
8.
go back to reference Wang, S., Yu, S., Huang, B., Feng, J., Liu, S.: Unique synergism between zinc dialkyldithiophosphates and Schiff base bridged phenolic diphenylamine antioxidants. Tribol. Int. 145, 106134 (2020) Wang, S., Yu, S., Huang, B., Feng, J., Liu, S.: Unique synergism between zinc dialkyldithiophosphates and Schiff base bridged phenolic diphenylamine antioxidants. Tribol. Int. 145, 106134 (2020)
9.
go back to reference Masuko, M., Ohkido, T., Suzuki, A., Ueno, T.: Fundamental study of changes in friction and wear characteristics due to ZnDTP deterioration in simulating engine oil degradation during use. In Proc Leeds Lyon Symposium on Tribology, Transient Processes in Tribology. Tribology Series, Elsevier, 43, 359–366, (2003). Masuko, M., Ohkido, T., Suzuki, A., Ueno, T.: Fundamental study of changes in friction and wear characteristics due to ZnDTP deterioration in simulating engine oil degradation during use. In Proc Leeds Lyon Symposium on Tribology, Transient Processes in Tribology. Tribology Series, Elsevier, 43, 359–366, (2003).
10.
go back to reference Dörr, N., Brenner, J., Ristić, A., Ronai, B., Besser, C., Pejaković, V., Frauscher, M.: Correlation between engine oil degradation, tribochemistry, and tribological behavior with focus on ZDDP deterioration. Tribol. Lett. 67, 62 (2019) Dörr, N., Brenner, J., Ristić, A., Ronai, B., Besser, C., Pejaković, V., Frauscher, M.: Correlation between engine oil degradation, tribochemistry, and tribological behavior with focus on ZDDP deterioration. Tribol. Lett. 67, 62 (2019)
11.
go back to reference Agocs, A., Besser, C., Brenner, J., Budnyk, S., Frauscher, M., Dörr, N.: Engine oils in the field: a comprehensive tribological assessment of engine oil degradation in a passenger car. Tribol. Lett. 70, 28 (2022) Agocs, A., Besser, C., Brenner, J., Budnyk, S., Frauscher, M., Dörr, N.: Engine oils in the field: a comprehensive tribological assessment of engine oil degradation in a passenger car. Tribol. Lett. 70, 28 (2022)
12.
go back to reference Harrison, P.G., Kikabhai, T.: Proton and phosphorus-31 nuclear magnetic resonance study of zinc(II) O, O′-dialkyl dithiophosphates in solution. J. Chem. Soc. Dalt. Trans. 4, 807–814 (1987) Harrison, P.G., Kikabhai, T.: Proton and phosphorus-31 nuclear magnetic resonance study of zinc(II) O, O′-dialkyl dithiophosphates in solution. J. Chem. Soc. Dalt. Trans. 4, 807–814 (1987)
13.
go back to reference Harrison, J.J., Chan, C.Y., Onopchenko, A., Pradhan, A.R., Petersen, M.: Neutral zinc(II) O, O-di-alkyldithiophosphates - variable temperature 31P NMR and quantum chemical study of the ZDDP monomer-dimer equilibrium. Magn. Reson. Chem. 46, 115–124 (2008)PubMed Harrison, J.J., Chan, C.Y., Onopchenko, A., Pradhan, A.R., Petersen, M.: Neutral zinc(II) O, O-di-alkyldithiophosphates - variable temperature 31P NMR and quantum chemical study of the ZDDP monomer-dimer equilibrium. Magn. Reson. Chem. 46, 115–124 (2008)PubMed
14.
go back to reference Dickert, J.J., Rowe, C.N.: The thermal decomposition of metal O, O-dialkylphorodithioates. J. Org. Chem. 32, 647–653 (1967) Dickert, J.J., Rowe, C.N.: The thermal decomposition of metal O, O-dialkylphorodithioates. J. Org. Chem. 32, 647–653 (1967)
15.
go back to reference Brazier, A.D., Elliott, J.S.: The thermal stability of zinc dithiophosphates. J. Inst. Pet. 53, 63–76 (1967) Brazier, A.D., Elliott, J.S.: The thermal stability of zinc dithiophosphates. J. Inst. Pet. 53, 63–76 (1967)
16.
go back to reference Ashford, J.S., Bretherick, L., Gould, P.: The thermal decomposition of zinc di-(4-methylpentyl-2) dithiophosphate. J. Appl. Chem. 15, 170–178 (1965) Ashford, J.S., Bretherick, L., Gould, P.: The thermal decomposition of zinc di-(4-methylpentyl-2) dithiophosphate. J. Appl. Chem. 15, 170–178 (1965)
17.
go back to reference Yamaguchi, E.S., Ruhe Jr, W.R.: Lower ash lubricating oil having ultra-neutral zinc dialkyldithiophosphates. US Patent 5728656 (1998) Yamaguchi, E.S., Ruhe Jr, W.R.: Lower ash lubricating oil having ultra-neutral zinc dialkyldithiophosphates. US Patent 5728656 (1998)
18.
go back to reference Rigdon, O.W., Edwards, R.S., Powers III, W.J.: Process for preparation of zinc dialkyl dithiophosphates. US Patent 3686243, (1972) Rigdon, O.W., Edwards, R.S., Powers III, W.J.: Process for preparation of zinc dialkyl dithiophosphates. US Patent 3686243, (1972)
19.
go back to reference Wystrach, V.P., Hook, E.O., Christopher, G.L.M.: Basic double salts of O, O-dialkyl phosphorodithioate acids. J. Org. Chem. 21, 705–707 (1956) Wystrach, V.P., Hook, E.O., Christopher, G.L.M.: Basic double salts of O, O-dialkyl phosphorodithioate acids. J. Org. Chem. 21, 705–707 (1956)
20.
go back to reference Jones, W.G., Pope, M.I.: The heats of adsorption of zinc di n-alkyl dithiophosphates onto pure iron, cast iron and ferric oxide surfaces. Thermochim. Acta C130, 141–148 (1988) Jones, W.G., Pope, M.I.: The heats of adsorption of zinc di n-alkyl dithiophosphates onto pure iron, cast iron and ferric oxide surfaces. Thermochim. Acta C130, 141–148 (1988)
21.
go back to reference Willermet, P.A., Mahoney, L.R., Haas, C.M.: The effects of antioxidant reactions on the wear behavior of a zinc dialkyldithiophosphate. ASLE Trans. 22, 301–306 (1979) Willermet, P.A., Mahoney, L.R., Haas, C.M.: The effects of antioxidant reactions on the wear behavior of a zinc dialkyldithiophosphate. ASLE Trans. 22, 301–306 (1979)
22.
go back to reference Yamaguchi, E.S., Ryason, P.R., Labrador, E.Q.: Inelastic electron tunneling spectra of neutral and basic zinc dithiophosphates on native aluminum oxide surfaces. Tribol. Trans. 38, 243–250 (1995) Yamaguchi, E.S., Ryason, P.R., Labrador, E.Q.: Inelastic electron tunneling spectra of neutral and basic zinc dithiophosphates on native aluminum oxide surfaces. Tribol. Trans. 38, 243–250 (1995)
23.
go back to reference Korcek, S., Nakada, M.: Engine oil performance requirements and reformulation for future gasoline engines and systems. SAE Trans. 105, 909–915 (1996) Korcek, S., Nakada, M.: Engine oil performance requirements and reformulation for future gasoline engines and systems. SAE Trans. 105, 909–915 (1996)
24.
go back to reference Uy, D., O’Neill, A.E.: Raman studies of automotive catalyst deactivation. SAE Trans. 115, 136–143 (2006) Uy, D., O’Neill, A.E.: Raman studies of automotive catalyst deactivation. SAE Trans. 115, 136–143 (2006)
25.
go back to reference Spikes, H.: Low- and zero-sulphated ash, phosphorus and sulphur anti-wear additives for engine oils. Lubr. Sci. 20, 103–136 (2008) Spikes, H.: Low- and zero-sulphated ash, phosphorus and sulphur anti-wear additives for engine oils. Lubr. Sci. 20, 103–136 (2008)
26.
go back to reference Bardasz, E.A., Schiferl, E., Nahumck, W., Kelley, J., Williams, L., Riley, M.J., Hubbard, C.P.: Low volatility ZDDP technology: part 1 - engines and lubricant performance in field applications. SAE Tech. Pap. 01, 19 (2007) Bardasz, E.A., Schiferl, E., Nahumck, W., Kelley, J., Williams, L., Riley, M.J., Hubbard, C.P.: Low volatility ZDDP technology: part 1 - engines and lubricant performance in field applications. SAE Tech. Pap. 01, 19 (2007)
27.
go back to reference Bardasz, E.A., Schiferl, E., Nahumck, W., Kelley, J., Williams, L., Hubbard, C.P., Thanasiu, E., Jagner, M., O’Neill, A., Uy, D.: Low volatility ZDDP technology: Part 2 - exhaust catalysts performance in field applications. SAE Tech. Pap. 01, 41 (2007) Bardasz, E.A., Schiferl, E., Nahumck, W., Kelley, J., Williams, L., Hubbard, C.P., Thanasiu, E., Jagner, M., O’Neill, A., Uy, D.: Low volatility ZDDP technology: Part 2 - exhaust catalysts performance in field applications. SAE Tech. Pap. 01, 41 (2007)
28.
go back to reference Wang, L. Wang, L., Li, G., Xu, X.: The effect of ZDDP type on phosphorus volatility: Engine oil performance on sequence IIIG and field tests. SAE Tech. Pap. no. 2013–01–25, (2013) Wang, L. Wang, L., Li, G., Xu, X.: The effect of ZDDP type on phosphorus volatility: Engine oil performance on sequence IIIG and field tests. SAE Tech. Pap. no. 2013–01–25, (2013)
29.
go back to reference Cann, P.M., Spikes, H.A., Hutchinson, J., J.: The development of a spacer layer imaging method (SLIM) for mapping elastohydrodynamic contacts. Tribol. Trans. 9, 915–921 (1996) Cann, P.M., Spikes, H.A., Hutchinson, J., J.: The development of a spacer layer imaging method (SLIM) for mapping elastohydrodynamic contacts. Tribol. Trans. 9, 915–921 (1996)
30.
go back to reference Giannuzzi, L.A., Stevie, F.A. (eds.): Introduction to focused ion beams instrumentation, theory, techniques and practice. Springer, New York (2005) Giannuzzi, L.A., Stevie, F.A. (eds.): Introduction to focused ion beams instrumentation, theory, techniques and practice. Springer, New York (2005)
31.
go back to reference Sawyer, W.G., Wahl, K.J.: Accessing inaccessible interfaces : in situ approaches to materials tribology. MRS Bull. 33, 1145–1150 (2008) Sawyer, W.G., Wahl, K.J.: Accessing inaccessible interfaces : in situ approaches to materials tribology. MRS Bull. 33, 1145–1150 (2008)
32.
go back to reference Jacobs, T.D.B., Greiner, C., Wahl, K.J., Carpick, R.W.: Insights into tribology from in situ nanoscale experiments. MRS Bull. 44(June), 478–486 (2019) Jacobs, T.D.B., Greiner, C., Wahl, K.J., Carpick, R.W.: Insights into tribology from in situ nanoscale experiments. MRS Bull. 44(June), 478–486 (2019)
33.
go back to reference Pidduck, A.J., Smith, G.C.: Scanning probe microscopy of automotive anti-wear films. Wear 212, 254–264 (1997) Pidduck, A.J., Smith, G.C.: Scanning probe microscopy of automotive anti-wear films. Wear 212, 254–264 (1997)
34.
go back to reference Topolovec-Miklozic, K., Forbus, T.R., Spikes, H.A.: Film thickness and roughness of ZDDP antiwear films. Tribol. Lett. 26, 161–171 (2007) Topolovec-Miklozic, K., Forbus, T.R., Spikes, H.A.: Film thickness and roughness of ZDDP antiwear films. Tribol. Lett. 26, 161–171 (2007)
35.
go back to reference Li, Y.R., Pereira, G., Lachenwitzer, A., Kasrai, M., Norton, P.R.: Studies on ZDDP thermal film formation by XANES spectroscopy, stomic force microscopy, FIB/SEM and 31P NMR. Tribol. Lett. 29, 11–22 (2008) Li, Y.R., Pereira, G., Lachenwitzer, A., Kasrai, M., Norton, P.R.: Studies on ZDDP thermal film formation by XANES spectroscopy, stomic force microscopy, FIB/SEM and 31P NMR. Tribol. Lett. 29, 11–22 (2008)
36.
go back to reference Benedet, J., Green, J.H., Lamb, G.D., Spikes, H.A.: Spurious mild wear measurement using white light interference microscopy in the presence of antiwear films. Tribol. Trans. 52, 841–846 (2009) Benedet, J., Green, J.H., Lamb, G.D., Spikes, H.A.: Spurious mild wear measurement using white light interference microscopy in the presence of antiwear films. Tribol. Trans. 52, 841–846 (2009)
37.
go back to reference Oblak, E., Kalin, M.: Relationship between the nanoscale topographical and mechanical properties of tribochemical films on DLC coatings and their macroscopic friction behavior. Tribol. Lett. 59, 1–16 (2015) Oblak, E., Kalin, M.: Relationship between the nanoscale topographical and mechanical properties of tribochemical films on DLC coatings and their macroscopic friction behavior. Tribol. Lett. 59, 1–16 (2015)
38.
go back to reference Jelita Rydel, J.K., Pagkalis, K., Kadiric, A., Rivera-Díaz-del-Castillo, P.E.J.: The correlation between ZDDP tribofilm morphology and the microstructure of steel. Tribol. Int. 113, 13–25 (2017) Jelita Rydel, J.K., Pagkalis, K., Kadiric, A., Rivera-Díaz-del-Castillo, P.E.J.: The correlation between ZDDP tribofilm morphology and the microstructure of steel. Tribol. Int. 113, 13–25 (2017)
39.
go back to reference Shimizu, Y., Spikes, H.A.: The influence of slide–roll ratio on ZDDP tribofilm formation. Tribol. Lett. 64, 19 (2016) Shimizu, Y., Spikes, H.A.: The influence of slide–roll ratio on ZDDP tribofilm formation. Tribol. Lett. 64, 19 (2016)
40.
go back to reference Gosvami, N.N., Bares, J.A., Mangolini, F., Konicek, A.R., Yablon, D.G., Carpick, R.W.: Mechanisms of antiwear tribofilm growth revealed in situ by single-asperity sliding contacts. Science 348, 102–106 (2015)PubMed Gosvami, N.N., Bares, J.A., Mangolini, F., Konicek, A.R., Yablon, D.G., Carpick, R.W.: Mechanisms of antiwear tribofilm growth revealed in situ by single-asperity sliding contacts. Science 348, 102–106 (2015)PubMed
41.
go back to reference Liu, X., Tang, R.C.Y., Hao, R., Walsh, K., Zhou, C., Dillon, S.J.: Local chemo-mechanical insights into the efficacy of ZDDP additives from in situ single asperity growth and mechanical testing. Tribol. Int. 112, 103–107 (2017) Liu, X., Tang, R.C.Y., Hao, R., Walsh, K., Zhou, C., Dillon, S.J.: Local chemo-mechanical insights into the efficacy of ZDDP additives from in situ single asperity growth and mechanical testing. Tribol. Int. 112, 103–107 (2017)
42.
go back to reference Dorgham, A., Parsaeian, P., Azam, A., Wang, C., Morina, A., Neville, A.: Single-asperity study of the reaction kinetics of P-based triboreactive films. Tribol. Int. 133, 288–296 (2019) Dorgham, A., Parsaeian, P., Azam, A., Wang, C., Morina, A., Neville, A.: Single-asperity study of the reaction kinetics of P-based triboreactive films. Tribol. Int. 133, 288–296 (2019)
43.
go back to reference Dorgham, A., Azam, A.M.: On the transient decomposition and reaction kinetics of zinc dialkyldithiophosphate. ACS Appl. Mater. Interfaces 10, 44803–44814 (2018)PubMed Dorgham, A., Azam, A.M.: On the transient decomposition and reaction kinetics of zinc dialkyldithiophosphate. ACS Appl. Mater. Interfaces 10, 44803–44814 (2018)PubMed
44.
go back to reference Gosvami, N.N., Lahouij, I., Ma, J., Carpick, R.W.: Nanoscale in situ study of ZDDP tribofilm growth at aluminum-based interfaces using atomic force microscopy. Tribol. Int. 143, 106075 (2020) Gosvami, N.N., Lahouij, I., Ma, J., Carpick, R.W.: Nanoscale in situ study of ZDDP tribofilm growth at aluminum-based interfaces using atomic force microscopy. Tribol. Int. 143, 106075 (2020)
45.
go back to reference Mittal, P., Maithani, Y., Singh, J.P., Gosvami, N.N.: In situ microscopic study of tribology and growth of ZDDP antiwear tribofilms on an Al–Si alloy. Tribol. Int. 151, 106419 (2020) Mittal, P., Maithani, Y., Singh, J.P., Gosvami, N.N.: In situ microscopic study of tribology and growth of ZDDP antiwear tribofilms on an Al–Si alloy. Tribol. Int. 151, 106419 (2020)
46.
go back to reference Yamashita, N., Hirayama, T.: Effect of atmospheric gas on ZDDP tribofilm formation. Tribol. Int. 193, 109400 (2024) Yamashita, N., Hirayama, T.: Effect of atmospheric gas on ZDDP tribofilm formation. Tribol. Int. 193, 109400 (2024)
47.
go back to reference Sato, K., Watanabe, S., Sasaki, S.: High friction mechanism of ZDDP tribofilm based on in situ AFM observation of nano-friction and adhesion properties. Tribol. Lett. 70, 94 (2022) Sato, K., Watanabe, S., Sasaki, S.: High friction mechanism of ZDDP tribofilm based on in situ AFM observation of nano-friction and adhesion properties. Tribol. Lett. 70, 94 (2022)
48.
go back to reference Dorgham, A., Azam, A., Parsaeian, P., Wang, C., Morina, A., Neville, A.: Nanoscale viscosity of triboreactive interfaces. Nano Energy 79, 105447 (2021) Dorgham, A., Azam, A., Parsaeian, P., Wang, C., Morina, A., Neville, A.: Nanoscale viscosity of triboreactive interfaces. Nano Energy 79, 105447 (2021)
49.
go back to reference Mikložič, K.T., Spikes, H.A., H.A.: Application of atomic force microscopy to the study of lubricant additive films Trans ASME. J. Tribol. 127, 405–415 (2005) Mikložič, K.T., Spikes, H.A., H.A.: Application of atomic force microscopy to the study of lubricant additive films Trans ASME. J. Tribol. 127, 405–415 (2005)
50.
go back to reference Topolovec Miklozic, K.: Study of boundary lubricating films using atomic force microscopy. PhD Thesis, University of London, London (2003) Topolovec Miklozic, K.: Study of boundary lubricating films using atomic force microscopy. PhD Thesis, University of London, London (2003)
51.
go back to reference Gosvami, N.N., Ma, J., Carpick, R.W.: An in situ method for simultaneous friction measurements and imaging of interfacial tribochemical film growth in lubricated contacts. Tribol. Lett. 66, 154 (2018) Gosvami, N.N., Ma, J., Carpick, R.W.: An in situ method for simultaneous friction measurements and imaging of interfacial tribochemical film growth in lubricated contacts. Tribol. Lett. 66, 154 (2018)
52.
go back to reference Taylor, L., Dratva, A., Spikes, H.A.: Friction and wear behavior of zinc dialkyldithiophosphate additive. Tribol. Trans. 43, 469–479 (2000) Taylor, L., Dratva, A., Spikes, H.A.: Friction and wear behavior of zinc dialkyldithiophosphate additive. Tribol. Trans. 43, 469–479 (2000)
53.
go back to reference Choo, J.W., Olver, A.V., Spikes, H.A.: Influence of surface roughness features on mixed-film lubrication. Lubr. Sci. 15, 219–232 (2003) Choo, J.W., Olver, A.V., Spikes, H.A.: Influence of surface roughness features on mixed-film lubrication. Lubr. Sci. 15, 219–232 (2003)
54.
go back to reference MacLaren, A., LaMascus, P., Carpick, R.W.: Enhancing range and reliability of the spacer layer imaging method. Tribol. Lett. 72, 95 (2024) MacLaren, A., LaMascus, P., Carpick, R.W.: Enhancing range and reliability of the spacer layer imaging method. Tribol. Lett. 72, 95 (2024)
55.
go back to reference Naveira-Suarez, A., Tomala, A., Grahn, M., Zaccheddu, M., Pasaribu, R., Larsson, R.: The influence of base oil polarity and slide-roll ratio on additive-derived reaction layer formation. Proc. Inst. Mech. Eng. Part J 225, 565–576 (2011) Naveira-Suarez, A., Tomala, A., Grahn, M., Zaccheddu, M., Pasaribu, R., Larsson, R.: The influence of base oil polarity and slide-roll ratio on additive-derived reaction layer formation. Proc. Inst. Mech. Eng. Part J 225, 565–576 (2011)
56.
go back to reference Fujita, H., Glovnea, R.P., Spikes, H.A.: Study of zinc dialkydithiophosphate antiwear film formation and removal processes, part I: experimental. Tribol. Trans. 48, 558–566 (2005) Fujita, H., Glovnea, R.P., Spikes, H.A.: Study of zinc dialkydithiophosphate antiwear film formation and removal processes, part I: experimental. Tribol. Trans. 48, 558–566 (2005)
57.
go back to reference Dawczyk, J., Russo, J., Spikes, H.A.: Ethoxylated amine friction modifiers and ZDDP. Tribol. Lett. 67, 106 (2019) Dawczyk, J., Russo, J., Spikes, H.A.: Ethoxylated amine friction modifiers and ZDDP. Tribol. Lett. 67, 106 (2019)
58.
go back to reference Zhang, Z., Yamaguchi, E.S., Kasrai, M., Bancroft, G.M.: Interaction of ZDDP with borated dispersant using XANES and XPS. Tribol. Trans. 47, 527–536 (2004) Zhang, Z., Yamaguchi, E.S., Kasrai, M., Bancroft, G.M.: Interaction of ZDDP with borated dispersant using XANES and XPS. Tribol. Trans. 47, 527–536 (2004)
59.
go back to reference Parsaeian, P., Ghanbarzadeh, A., Van Eijk, M.C.P., Nedelcu, I., Neville, A., Morina, A.: A new insight into the interfacial mechanisms of the tribofilm formed by zinc dialkyl dithiophosphate. Appl. Surf. Sci. 403, 472–486 (2017) Parsaeian, P., Ghanbarzadeh, A., Van Eijk, M.C.P., Nedelcu, I., Neville, A., Morina, A.: A new insight into the interfacial mechanisms of the tribofilm formed by zinc dialkyl dithiophosphate. Appl. Surf. Sci. 403, 472–486 (2017)
60.
go back to reference Ueda, M., Kadiric, A., Spikes, H.A.: On the crystallinity and durability of ZDDP tribofilm. Tribol. Lett. 67, 123 (2019) Ueda, M., Kadiric, A., Spikes, H.A.: On the crystallinity and durability of ZDDP tribofilm. Tribol. Lett. 67, 123 (2019)
61.
go back to reference Topolovec-Miklozic, K., Forbus, T.R., Spikes, H.A.: Performance of friction modifiers on zddp-generated surfaces. Tribol. Trans. 50, 328–335 (2007) Topolovec-Miklozic, K., Forbus, T.R., Spikes, H.A.: Performance of friction modifiers on zddp-generated surfaces. Tribol. Trans. 50, 328–335 (2007)
62.
go back to reference Zhang, J., Yamaguchi, E.S., Spikes, H.A.: The antagonism between succinimide dispersants and a secondary zinc dialkyl dithiophosphate. Tribol. Trans. 57, 57–65 (2014) Zhang, J., Yamaguchi, E.S., Spikes, H.A.: The antagonism between succinimide dispersants and a secondary zinc dialkyl dithiophosphate. Tribol. Trans. 57, 57–65 (2014)
63.
go back to reference Fujita, H., Spikes, H.A.: Study of zinc dialkyldithiophosphate antiwear film formation and removal processes, part II: Kinetic model. Tribol. Trans. 48, 567–575 (2005) Fujita, H., Spikes, H.A.: Study of zinc dialkyldithiophosphate antiwear film formation and removal processes, part II: Kinetic model. Tribol. Trans. 48, 567–575 (2005)
64.
go back to reference Kapadia, R., Glyde, R., Wu, Y.: In situ observation of phosphorous and non-phosphorous antiwear films using a mini traction machine with spacer layer image mapping. Tribol. Int. 40, 1667–1679 (2007) Kapadia, R., Glyde, R., Wu, Y.: In situ observation of phosphorous and non-phosphorous antiwear films using a mini traction machine with spacer layer image mapping. Tribol. Int. 40, 1667–1679 (2007)
65.
go back to reference Olomolehin, Y., Kapadia, R., Spikes, H.A.: Antagonistic interaction of antiwear additives and carbon black. Tribol. Lett. 37, 49–58 (2010) Olomolehin, Y., Kapadia, R., Spikes, H.A.: Antagonistic interaction of antiwear additives and carbon black. Tribol. Lett. 37, 49–58 (2010)
66.
go back to reference Vengudusamy, B., Green, J.H., Lamb, G.D., Spikes, H.A.: Tribological properties of tribofilms formed from ZDDP in DLC/DLC and DLC/steel contacts. Tribol. Int. 44, 165–174 (2011) Vengudusamy, B., Green, J.H., Lamb, G.D., Spikes, H.A.: Tribological properties of tribofilms formed from ZDDP in DLC/DLC and DLC/steel contacts. Tribol. Int. 44, 165–174 (2011)
67.
go back to reference Ueda, M., Kadiric, A., Spikes, H.A.: Influence of steel surface composition on ZDDP tribofilm growth using ion implantation. Tribol. Lett. 69, 62 (2021) Ueda, M., Kadiric, A., Spikes, H.A.: Influence of steel surface composition on ZDDP tribofilm growth using ion implantation. Tribol. Lett. 69, 62 (2021)
68.
go back to reference Taşan, Y.C., De Rooij, M.B., Schipper, D.J.: Measurement of wear on asperity level using image-processing techniques. Wear 258, 83–91 (2005) Taşan, Y.C., De Rooij, M.B., Schipper, D.J.: Measurement of wear on asperity level using image-processing techniques. Wear 258, 83–91 (2005)
69.
go back to reference Dawczyk, J., Morgan, N., Russo, J., Spikes, H.: Film thickness and friction of ZDDP tribofilms. Tribol. Lett. 67, 34 (2019) Dawczyk, J., Morgan, N., Russo, J., Spikes, H.: Film thickness and friction of ZDDP tribofilms. Tribol. Lett. 67, 34 (2019)
70.
go back to reference Shimizu, Y., Spikes, H.A.: The tribofilm formation of ZDDP under reciprocating pure sliding conditions. Tribol. Lett. 64, 46 (2016) Shimizu, Y., Spikes, H.A.: The tribofilm formation of ZDDP under reciprocating pure sliding conditions. Tribol. Lett. 64, 46 (2016)
71.
go back to reference Okubo, H., Tadokoro, C., Sasaki, S.: In situ Raman-SLIM monitoring for the formation processes of MoDTC and ZDDP tribofilms at steel/steel contacts under boundary lubrication. Tribol. Online 15, 105–116 (2020) Okubo, H., Tadokoro, C., Sasaki, S.: In situ Raman-SLIM monitoring for the formation processes of MoDTC and ZDDP tribofilms at steel/steel contacts under boundary lubrication. Tribol. Online 15, 105–116 (2020)
72.
go back to reference Warren, O.L., Graham, J.F., Norton, P.R., Houston, J.E., Michalski, T.A.: Nanomechanical properties of films derived from zinc dialkyldithiophosphate. Tribol. Lett. 4, 189–198 (1998) Warren, O.L., Graham, J.F., Norton, P.R., Houston, J.E., Michalski, T.A.: Nanomechanical properties of films derived from zinc dialkyldithiophosphate. Tribol. Lett. 4, 189–198 (1998)
73.
go back to reference Graham, J.F., McCague, C., Norton, P.R.: Topography and nanomechanical properties of tribochemical films derived from zinc dialkyl and diaryl dithiophosphates. Tribol. Lett. 6, 149–157 (1999) Graham, J.F., McCague, C., Norton, P.R.: Topography and nanomechanical properties of tribochemical films derived from zinc dialkyl and diaryl dithiophosphates. Tribol. Lett. 6, 149–157 (1999)
74.
go back to reference Bec, S., Tonck, A., Georges, J.M., Coy, R.C., Bell, J.C., Roper, G.W.: Relationship between mechanical properties and structures of zinc dithiophosphate anti-wear films. Proc. R. Soc. A Math. Phys. Eng. Sci. 455, 4181–4203 (1999) Bec, S., Tonck, A., Georges, J.M., Coy, R.C., Bell, J.C., Roper, G.W.: Relationship between mechanical properties and structures of zinc dithiophosphate anti-wear films. Proc. R. Soc. A Math. Phys. Eng. Sci. 455, 4181–4203 (1999)
75.
go back to reference Aktary, M., McDermott, M.T., McAlpine, G.A.: Morphology and nanomechanical properties of ZDDP antiwear films as a function of tribological contact time. Tribol. Lett. 12, 155–162 (2002) Aktary, M., McDermott, M.T., McAlpine, G.A.: Morphology and nanomechanical properties of ZDDP antiwear films as a function of tribological contact time. Tribol. Lett. 12, 155–162 (2002)
76.
go back to reference Ye, J., Kano, M., Yasuda, Y.: Evaluation of local mechanical properties in depth in MoDTC/ZDDP and ZDDP tribochemical reacted films using nanoindentation. Tribol. Lett. 13, 41–47 (2002) Ye, J., Kano, M., Yasuda, Y.: Evaluation of local mechanical properties in depth in MoDTC/ZDDP and ZDDP tribochemical reacted films using nanoindentation. Tribol. Lett. 13, 41–47 (2002)
77.
go back to reference Demmou, K., Bec, S., Loubet, J.L., Martin, J.M.: Temperature effects on mechanical properties of zinc dithiophosphate tribofilms. Tribol. Int. 39, 1558–1563 (2006) Demmou, K., Bec, S., Loubet, J.L., Martin, J.M.: Temperature effects on mechanical properties of zinc dithiophosphate tribofilms. Tribol. Int. 39, 1558–1563 (2006)
78.
go back to reference Pereira, G., Munoz-Paniagua, D., Lachenwitzer, A., Kasrai, M., Norton, P.R., Capehart, T.W., Perry, T.A., Cheng, Y.T.: A variable temperature mechanical analysis of ZDDP-derived antiwear films formed on 52100 steel. Wear 262, 461–470 (2007) Pereira, G., Munoz-Paniagua, D., Lachenwitzer, A., Kasrai, M., Norton, P.R., Capehart, T.W., Perry, T.A., Cheng, Y.T.: A variable temperature mechanical analysis of ZDDP-derived antiwear films formed on 52100 steel. Wear 262, 461–470 (2007)
79.
go back to reference Kalin, M., Oblak, E.: Akbari, S: Evolution of the nano-scale mechanical properties of tribofilms formed from low- and high-SAPS oils and ZDDP on DLC coatings and steel. Tribol. Int. 96, 43–56 (2016) Kalin, M., Oblak, E.: Akbari, S: Evolution of the nano-scale mechanical properties of tribofilms formed from low- and high-SAPS oils and ZDDP on DLC coatings and steel. Tribol. Int. 96, 43–56 (2016)
80.
go back to reference Landauer, A.K., Barnhill, W.C., Qu, J.: Correlating mechanical properties and anti-wear performance of tribo films formed by ionic liquids, ZDDP and their combinations. Wear 354–355, 78–82 (2016) Landauer, A.K., Barnhill, W.C., Qu, J.: Correlating mechanical properties and anti-wear performance of tribo films formed by ionic liquids, ZDDP and their combinations. Wear 354–355, 78–82 (2016)
81.
go back to reference Mourhatch, R., Aswath, P.B.: Tribological behavior and nature of tribofilms generated from fluorinated ZDDP in comparison to ZDDP under extreme pressure conditions — part II : morphology and nanoscale properties of tribofilms. Tribol. Int. 44, 201–210 (2011) Mourhatch, R., Aswath, P.B.: Tribological behavior and nature of tribofilms generated from fluorinated ZDDP in comparison to ZDDP under extreme pressure conditions — part II : morphology and nanoscale properties of tribofilms. Tribol. Int. 44, 201–210 (2011)
82.
go back to reference Ueda, M., Wainwright, B., Spikes, H., Kadiric, A.: The effect of friction on micropitting. Wear 488–489, 204130 (2022) Ueda, M., Wainwright, B., Spikes, H., Kadiric, A.: The effect of friction on micropitting. Wear 488–489, 204130 (2022)
83.
go back to reference Pharr, G.M., Strader, J.H., Oliver, W.C.: Critical issues in making small-depth mechanical property measurements by nanoindentation with continuous stiffness measurement. J. Mater. Res. 24, 653–666 (2009) Pharr, G.M., Strader, J.H., Oliver, W.C.: Critical issues in making small-depth mechanical property measurements by nanoindentation with continuous stiffness measurement. J. Mater. Res. 24, 653–666 (2009)
84.
go back to reference Li, X., Bhushan, B.: A review of nanoindentation continuous stiffness measurement technique and its applications. Mater Charact 48, 11–36 (2002) Li, X., Bhushan, B.: A review of nanoindentation continuous stiffness measurement technique and its applications. Mater Charact 48, 11–36 (2002)
85.
go back to reference Bhushan, B., Li, X.: Nanomechanical characterisation of solid surfaces and thin films. Intern. Mat. Rev. 48, 125–164 (2003) Bhushan, B., Li, X.: Nanomechanical characterisation of solid surfaces and thin films. Intern. Mat. Rev. 48, 125–164 (2003)
86.
go back to reference Fuller, M., Yin, Z., Kasrai, M., Bancroft, G.M., Yamaguchi, E.S., Ryason, P.R., WilIermet, P.A., Tan, K.H.: Chemical characterization of tribochemical and thermal films generated from neutral and basic ZDDPs using X-ray absorption spectroscopy. Tribol. Int. 30, 305–315 (1997) Fuller, M., Yin, Z., Kasrai, M., Bancroft, G.M., Yamaguchi, E.S., Ryason, P.R., WilIermet, P.A., Tan, K.H.: Chemical characterization of tribochemical and thermal films generated from neutral and basic ZDDPs using X-ray absorption spectroscopy. Tribol. Int. 30, 305–315 (1997)
87.
go back to reference Martin, J.M., Grossiord, C., Le Mogne, T., Bec, S., Tonck, A.: The two-layer structure of Zndtp tribofilms, part I: AES XPS and XANES analyses. Tribol. Int. 34, 523–530 (2001) Martin, J.M., Grossiord, C., Le Mogne, T., Bec, S., Tonck, A.: The two-layer structure of Zndtp tribofilms, part I: AES XPS and XANES analyses. Tribol. Int. 34, 523–530 (2001)
88.
go back to reference Lewin, E., Counsell, J., Patscheider, J.: Spectral artefacts post sputter-etching and how to cope with them – a case study of XPS on nitride-based coatings using monoatomic and cluster ion beams. Appl. Surf. Sci. 442, 487–500 (2018) Lewin, E., Counsell, J., Patscheider, J.: Spectral artefacts post sputter-etching and how to cope with them – a case study of XPS on nitride-based coatings using monoatomic and cluster ion beams. Appl. Surf. Sci. 442, 487–500 (2018)
89.
go back to reference Heuberger, R., Rossi, A., Spencer, N.D.: XPS study of the influence of temperature on ZnDTP tribofilm composition. Tribol. Lett. 25, 185–196 (2007) Heuberger, R., Rossi, A., Spencer, N.D.: XPS study of the influence of temperature on ZnDTP tribofilm composition. Tribol. Lett. 25, 185–196 (2007)
90.
go back to reference Zhou, J.G., Thompson, J., Cutler, J., Blyth, R., Kasrai, M., Bancroft, G.M., Yamaguchi, E.S.: Resolving the chemical variation of phosphates in thin ZDDP tribofilms by X-ray photoelectron spectroscopy using synchrotron radiation: evidence for ultraphosphates and organic phosphates. Tribol. Lett. 39, 101–107 (2010) Zhou, J.G., Thompson, J., Cutler, J., Blyth, R., Kasrai, M., Bancroft, G.M., Yamaguchi, E.S.: Resolving the chemical variation of phosphates in thin ZDDP tribofilms by X-ray photoelectron spectroscopy using synchrotron radiation: evidence for ultraphosphates and organic phosphates. Tribol. Lett. 39, 101–107 (2010)
91.
go back to reference Minfray, C., Martin, J.M., Esnouf, C., Le Mogne, T., Kersting, R., Hagenhoff, D.: A multi-technique approach of tribofilm characterisation. Thin Solid Films 447–448, 272–277 (2004) Minfray, C., Martin, J.M., Esnouf, C., Le Mogne, T., Kersting, R., Hagenhoff, D.: A multi-technique approach of tribofilm characterisation. Thin Solid Films 447–448, 272–277 (2004)
92.
go back to reference Ito, K., Martin, J.M., Minfray, C., Kato, K.: Low-friction tribofilm formed by the reaction of ZDDP on iron oxide. Tribol. Int. 39, 538–1544 (2006) Ito, K., Martin, J.M., Minfray, C., Kato, K.: Low-friction tribofilm formed by the reaction of ZDDP on iron oxide. Tribol. Int. 39, 538–1544 (2006)
93.
go back to reference Dawczyk, J., Ware, E., Ardakani, M., Russo, J., Spikes, H.A.: Use of FIB to study ZDDP tribofilms. Tribol. Lett. 66, 155 (2018) Dawczyk, J., Ware, E., Ardakani, M., Russo, J., Spikes, H.A.: Use of FIB to study ZDDP tribofilms. Tribol. Lett. 66, 155 (2018)
94.
go back to reference Notthoff, C., Winterer, M., Beckel, A., Geller, M.: Spatial high resolution energy dispersive X-ray spectroscopy on thin lamellas. Ultramicroscopy 129, 30–35 (2013)PubMed Notthoff, C., Winterer, M., Beckel, A., Geller, M.: Spatial high resolution energy dispersive X-ray spectroscopy on thin lamellas. Ultramicroscopy 129, 30–35 (2013)PubMed
95.
go back to reference Guo, W., Zhou, Y., Sang, X., Leonard, D.N., Qu, J., Poplawsky, J.D.: Atom probe tomography unveils formation mechanisms of wear-protective tribofilms by ZDDP, ionic liquid, and their combination. ACS Appl. Mater. Interfaces 9, 3152–23163 (2017) Guo, W., Zhou, Y., Sang, X., Leonard, D.N., Qu, J., Poplawsky, J.D.: Atom probe tomography unveils formation mechanisms of wear-protective tribofilms by ZDDP, ionic liquid, and their combination. ACS Appl. Mater. Interfaces 9, 3152–23163 (2017)
96.
go back to reference Hsu, C.J., Barrirero, J., Merz, R., Stratmann, A., Aboulfadl, H., Jacobs, G., Kopnarski, M., Mücklich, F., Gachot, C.: Revealing the interface nature of ZDDP tribofilm by X-ray photoelectron spectroscopy and atom probe tomography. Ind. Lubr. Tribol. 72, 923–930 (2020) Hsu, C.J., Barrirero, J., Merz, R., Stratmann, A., Aboulfadl, H., Jacobs, G., Kopnarski, M., Mücklich, F., Gachot, C.: Revealing the interface nature of ZDDP tribofilm by X-ray photoelectron spectroscopy and atom probe tomography. Ind. Lubr. Tribol. 72, 923–930 (2020)
97.
go back to reference Thompson, K., Lawrence, D., Larson, D.J., Olson, J.D., Kelly, T.F., Gorman, B.: In situ site-specific specimen preparation for atom probe tomography. Ultramicroscopy 107, 131–139 (2007)PubMed Thompson, K., Lawrence, D., Larson, D.J., Olson, J.D., Kelly, T.F., Gorman, B.: In situ site-specific specimen preparation for atom probe tomography. Ultramicroscopy 107, 131–139 (2007)PubMed
98.
go back to reference Dorgham, A., Parsaeian, P., Neville, A., Ignatyev, K., Mosselmans, F., Masuko, M., Morina, A.: In situ synchrotron XAS study of the decomposition kinetics of ZDDP triboreactive interfaces. RSC Adv. 8, 34168–34181 (2018)PubMedPubMedCentral Dorgham, A., Parsaeian, P., Neville, A., Ignatyev, K., Mosselmans, F., Masuko, M., Morina, A.: In situ synchrotron XAS study of the decomposition kinetics of ZDDP triboreactive interfaces. RSC Adv. 8, 34168–34181 (2018)PubMedPubMedCentral
99.
go back to reference Morina, A., Zhao, H., Mosselmans, J.F.W.: In-situ reflection-XANES study of ZDDP and MoDTC lubricant films formed on steel and diamond like carbon (DLC) surfaces. Appl. Surf. Sci. 297, 167–175 (2014) Morina, A., Zhao, H., Mosselmans, J.F.W.: In-situ reflection-XANES study of ZDDP and MoDTC lubricant films formed on steel and diamond like carbon (DLC) surfaces. Appl. Surf. Sci. 297, 167–175 (2014)
100.
go back to reference Loeser, E.H., Wiquist, R.C., Twiss, S.B.: Cam and tappet lubrication III — radioactive study of phosphorus in the EP film. ASLE Trans. 1, 329–335 (1958) Loeser, E.H., Wiquist, R.C., Twiss, S.B.: Cam and tappet lubrication III — radioactive study of phosphorus in the EP film. ASLE Trans. 1, 329–335 (1958)
101.
go back to reference Loeser, E.H., Wiquist, R.C., Twiss, S.B.: Cam and tappet lubrication. IV–radioactive study of sulfur in the EP film. ASLE Trans. 2, 199–207 (1959) Loeser, E.H., Wiquist, R.C., Twiss, S.B.: Cam and tappet lubrication. IV–radioactive study of sulfur in the EP film. ASLE Trans. 2, 199–207 (1959)
102.
go back to reference Bird, R.J., Galvin, G.D.: The application of photoelectron spectroscopy to the study of E.P. films on lubricated surfaces. Wear 37, 143–167 (1976) Bird, R.J., Galvin, G.D.: The application of photoelectron spectroscopy to the study of E.P. films on lubricated surfaces. Wear 37, 143–167 (1976)
103.
go back to reference Fujita, H., Spikes, H.A.: The formation of zinc dithiophosphate antiwear films. Proc. Inst. Mech. Eng. Part J 218, 265–277 (2004) Fujita, H., Spikes, H.A.: The formation of zinc dithiophosphate antiwear films. Proc. Inst. Mech. Eng. Part J 218, 265–277 (2004)
104.
go back to reference Hershberger, J., Ajayi, O.O., Fenske, G.R.: Zinc content of ZDDP films formed thermally and mechanically. Tribol. Int. 38, 299–303 (2005) Hershberger, J., Ajayi, O.O., Fenske, G.R.: Zinc content of ZDDP films formed thermally and mechanically. Tribol. Int. 38, 299–303 (2005)
105.
go back to reference Pauli, M.D., Rufael, T.S., Mowlem, J.K., Weinert, M., Saldin, D.K., Tysoe, W.T.: X-ray absorption near-edge structure analysis of the chemical environment of zinc in the tribological film formed by zinc dialkyl dithiophosphate decomposition on steel. Tribol. Int. 38, 195–204 (2005) Pauli, M.D., Rufael, T.S., Mowlem, J.K., Weinert, M., Saldin, D.K., Tysoe, W.T.: X-ray absorption near-edge structure analysis of the chemical environment of zinc in the tribological film formed by zinc dialkyl dithiophosphate decomposition on steel. Tribol. Int. 38, 195–204 (2005)
106.
go back to reference Bancroft, G.M., Kasrai, M., Fuller, M., Yin, Z., Fyfe, K., Tan, K.H.: Mechanisms of tribochemical film formation: Stability of tribo- and thermally-generated ZDDP films. Tribol. Lett. 3, 47–51 (1997) Bancroft, G.M., Kasrai, M., Fuller, M., Yin, Z., Fyfe, K., Tan, K.H.: Mechanisms of tribochemical film formation: Stability of tribo- and thermally-generated ZDDP films. Tribol. Lett. 3, 47–51 (1997)
107.
go back to reference So, H., Lin, Y.C.: The theory of antiwear for ZDDP at elevated temperature in boundary lubrication condition. Wear 177, 105–115 (1994) So, H., Lin, Y.C.: The theory of antiwear for ZDDP at elevated temperature in boundary lubrication condition. Wear 177, 105–115 (1994)
108.
go back to reference Beyer, M.K., Clausen-Schaumann, H.: Mechanochemistry : the mechanical activation of covalent bonds. Chem. Rev. 105, 2921–2948 (2005)PubMed Beyer, M.K., Clausen-Schaumann, H.: Mechanochemistry : the mechanical activation of covalent bonds. Chem. Rev. 105, 2921–2948 (2005)PubMed
109.
go back to reference Tan, D., Garcia, F.: Main group mechanochemistry: from curiosity to established protocols. Chem. Soc. Rev. 48, 2267–2496 (2019) Tan, D., Garcia, F.: Main group mechanochemistry: from curiosity to established protocols. Chem. Soc. Rev. 48, 2267–2496 (2019)
110.
go back to reference Friscic, T., Mottillo, C., Titi, H.M.: Mechanochemistry for synthesis. Angewante Chemie Int. Ed. 59, 1018–1029 (2020) Friscic, T., Mottillo, C., Titi, H.M.: Mechanochemistry for synthesis. Angewante Chemie Int. Ed. 59, 1018–1029 (2020)
111.
go back to reference Zhang, J., Spikes, H.A.: On the mechanism of ZDDP antiwear film formation. Tribol. Lett. 63, 24 (2016) Zhang, J., Spikes, H.A.: On the mechanism of ZDDP antiwear film formation. Tribol. Lett. 63, 24 (2016)
112.
go back to reference Zhang, J., Ewen, J.P., Ueda, M., Wong, J.S.S., Spikes, H.A.: Mechanochemistry of zinc dialkyldithiophosphate on steel surfaces under elastohydrodynamic lubrication conditions. ACS Appl. Mater. Interfaces 12, 6662–6676 (2020)PubMed Zhang, J., Ewen, J.P., Ueda, M., Wong, J.S.S., Spikes, H.A.: Mechanochemistry of zinc dialkyldithiophosphate on steel surfaces under elastohydrodynamic lubrication conditions. ACS Appl. Mater. Interfaces 12, 6662–6676 (2020)PubMed
113.
go back to reference Fang, L., Korres, S., Lamberti, W.A., Webster, M.N., Carpick, R.W.: What stress components drive mechanochemistry? A study of ZDDP tribofilm formation. Faraday Discuss. 241, 394–412 (2022) Fang, L., Korres, S., Lamberti, W.A., Webster, M.N., Carpick, R.W.: What stress components drive mechanochemistry? A study of ZDDP tribofilm formation. Faraday Discuss. 241, 394–412 (2022)
114.
go back to reference Yeon, J., He, X., Martini, A., Kim, S.H.: Mechanochemistry at solid surfaces: polymerization of adsorbed molecules by mechanical shear at tribological interfaces. ACS Appl. Mater. Interfaces 9, 3142–3148 (2017)PubMed Yeon, J., He, X., Martini, A., Kim, S.H.: Mechanochemistry at solid surfaces: polymerization of adsorbed molecules by mechanical shear at tribological interfaces. ACS Appl. Mater. Interfaces 9, 3142–3148 (2017)PubMed
115.
go back to reference Casale, A., Porter, R.S., Johnson, J.F.: Mechanochemistry of high polymers. Rubber Chem. Technol. 44, 534–777 (1971) Casale, A., Porter, R.S., Johnson, J.F.: Mechanochemistry of high polymers. Rubber Chem. Technol. 44, 534–777 (1971)
116.
go back to reference Akbulatov, S., Tian, Z.Y., Huang, Z., Kucharski, T.J., Yang, Q.Z., Boulatov, R.: Experimentally realized mechanochemistry distinct from force-accelerated scission of loaded bonds. Science 357, 299–303 (2017)PubMed Akbulatov, S., Tian, Z.Y., Huang, Z., Kucharski, T.J., Yang, Q.Z., Boulatov, R.: Experimentally realized mechanochemistry distinct from force-accelerated scission of loaded bonds. Science 357, 299–303 (2017)PubMed
117.
go back to reference Baumgarten, E.: Radiochemische Untersuchung uber die Sorption von Metall-diisopropyl-dithiophosphat an aufgedamften Metallschichten. Erdol und Kohle, Erdgas, Petrochemie 25, 577–582 (1972) Baumgarten, E.: Radiochemische Untersuchung uber die Sorption von Metall-diisopropyl-dithiophosphat an aufgedamften Metallschichten. Erdol und Kohle, Erdgas, Petrochemie 25, 577–582 (1972)
118.
go back to reference Baumgarten, E.: Radiochemische Untersuchung uber die Sorption von Metall-dialkyl-dithiophosphat an Metallpulvern. Erdol u Kohle 25, 28–32 (1972) Baumgarten, E.: Radiochemische Untersuchung uber die Sorption von Metall-dialkyl-dithiophosphat an Metallpulvern. Erdol u Kohle 25, 28–32 (1972)
119.
go back to reference Baumgarten, E.: Radiochemische Untersuchungen über die Sorption von Dialkyl-dithiophosphat an Bleiglanz und Zinkblende. Chemie Ing. Tech. 44, 613–617 (1972) Baumgarten, E.: Radiochemische Untersuchungen über die Sorption von Dialkyl-dithiophosphat an Bleiglanz und Zinkblende. Chemie Ing. Tech. 44, 613–617 (1972)
120.
go back to reference Dacre, B., Bovington, C.H.: The effect of metal composition on the adsorption of zinc di-isopropyldithiophosphate. ASLE Trans. 26, 333–343 (1983) Dacre, B., Bovington, C.H.: The effect of metal composition on the adsorption of zinc di-isopropyldithiophosphate. ASLE Trans. 26, 333–343 (1983)
121.
go back to reference Dacre, B., Bovington, C.H.: The adsorption and desorption of zinc di-isopropyldithiophosphate on steel. ASLE Trans. 25, 546–554 (1982) Dacre, B., Bovington, C.H.: The adsorption and desorption of zinc di-isopropyldithiophosphate on steel. ASLE Trans. 25, 546–554 (1982)
122.
go back to reference Otto, K., Sorek, H.: Adsorption studies with radiolabeled zinc dialkyldithiophosphate. Trans ASME J. Tribol. 108, 340–344 (1986) Otto, K., Sorek, H.: Adsorption studies with radiolabeled zinc dialkyldithiophosphate. Trans ASME J. Tribol. 108, 340–344 (1986)
123.
go back to reference Bovington, C.H., Dacre, B.: The adsorption and reaction of decomposition products of zinc di-isopropyldiophosphate on steel. ASLE Trans. 27, 252–258 (1984) Bovington, C.H., Dacre, B.: The adsorption and reaction of decomposition products of zinc di-isopropyldiophosphate on steel. ASLE Trans. 27, 252–258 (1984)
124.
go back to reference Tabibi, M.: Dispersant effects on zinc dialkyldithiophosphate (ZDDP) tribofilm structure and composition. MSc Thesis, Virginia Commonweath University (VCU) Graduate School, (2015) Tabibi, M.: Dispersant effects on zinc dialkyldithiophosphate (ZDDP) tribofilm structure and composition. MSc Thesis, Virginia Commonweath University (VCU) Graduate School, (2015)
125.
go back to reference Ueda, M., Kadiric, A., Spikes, H.A.: ZDDP tribofilm formation on non-ferrous surfaces. Tribol. Online 15, 318–331 (2020) Ueda, M., Kadiric, A., Spikes, H.A.: ZDDP tribofilm formation on non-ferrous surfaces. Tribol. Online 15, 318–331 (2020)
126.
go back to reference Ueda, M., Spikes, H., Kadiric, A.: Influence of black oxide coating on micropitting and ZDDP tribofilm formation. Tribol. Trans. 65, 242–259 (2021) Ueda, M., Spikes, H., Kadiric, A.: Influence of black oxide coating on micropitting and ZDDP tribofilm formation. Tribol. Trans. 65, 242–259 (2021)
127.
go back to reference Hastie, G.P., Roberts, K.J., Adams, D., Fischer, D.: Meitzner, G: Investigating the structural chemistry at the interface formed between zinc dialkyldithiophosphate (ZDDP) and mild steel using ultra-soft x-ray absorption spectroscopy. Jpn. J. Appl. Phys. 32, 407–409 (1993) Hastie, G.P., Roberts, K.J., Adams, D., Fischer, D.: Meitzner, G: Investigating the structural chemistry at the interface formed between zinc dialkyldithiophosphate (ZDDP) and mild steel using ultra-soft x-ray absorption spectroscopy. Jpn. J. Appl. Phys. 32, 407–409 (1993)
128.
go back to reference Yamaguchi, E.S., Ryason, P.R.: Inelastic electron tunneling spectra of lubricant oil additives on native aluminum oxide surfaces. Tribol. Trans. 36, 367–374 (1993) Yamaguchi, E.S., Ryason, P.R.: Inelastic electron tunneling spectra of lubricant oil additives on native aluminum oxide surfaces. Tribol. Trans. 36, 367–374 (1993)
129.
go back to reference Larsson, A.C., Ivanov, A.: A 31P CP/MAS NMR study of PbS surface O, O′-dialkyldithiophosphate lead(II) complexes. J. Colloid Interface Sci. 327, 370–376 (2008)PubMed Larsson, A.C., Ivanov, A.: A 31P CP/MAS NMR study of PbS surface O, O′-dialkyldithiophosphate lead(II) complexes. J. Colloid Interface Sci. 327, 370–376 (2008)PubMed
130.
go back to reference Chen, J., Lan, L., Chen, Y.: Computational simulation of adsorption and thermodynamic study of xanthate, dithiophosphate and dithiocarbamate on galena and pyrite surfaces. Miner. Eng. 46–47, 136–143 (2013) Chen, J., Lan, L., Chen, Y.: Computational simulation of adsorption and thermodynamic study of xanthate, dithiophosphate and dithiocarbamate on galena and pyrite surfaces. Miner. Eng. 46–47, 136–143 (2013)
131.
go back to reference Luther, D., Baumgarten, V.H., Staek, E.: Investigation of decomposition of dialkyldithiophosphates in hydrocarbons. Erdol und Kohle, Erdgas, Petrochemie 22, 530–536 (1969) Luther, D., Baumgarten, V.H., Staek, E.: Investigation of decomposition of dialkyldithiophosphates in hydrocarbons. Erdol und Kohle, Erdgas, Petrochemie 22, 530–536 (1969)
132.
go back to reference Jensen, R.K., Korcek, S., Johnson, M.D.: Friction-reducing and antioxidant capabilities of engine oil additive systems under oxidative conditions. II. Understanding ligand exchange in a molybdenum dialkyldithiocarbamate/zinc dialkyldithiophosphate additive system in various base oils. Lubr. Sci. 14, 25–42 (2001) Jensen, R.K., Korcek, S., Johnson, M.D.: Friction-reducing and antioxidant capabilities of engine oil additive systems under oxidative conditions. II. Understanding ligand exchange in a molybdenum dialkyldithiocarbamate/zinc dialkyldithiophosphate additive system in various base oils. Lubr. Sci. 14, 25–42 (2001)
133.
go back to reference Kiw, Y.M., Schaeffer, P., Adam, P., Thiébaut, B., Boyer, C.P.: Ligand exchange processes between molybdenum and zinc additives in lubricants: evidence from NMR (1H, 13C, 31P) and HPLC-MS analysis. RSC Adv. 10, 37962–37973 (2020)PubMedPubMedCentral Kiw, Y.M., Schaeffer, P., Adam, P., Thiébaut, B., Boyer, C.P.: Ligand exchange processes between molybdenum and zinc additives in lubricants: evidence from NMR (1H, 13C, 31P) and HPLC-MS analysis. RSC Adv. 10, 37962–37973 (2020)PubMedPubMedCentral
134.
go back to reference Smith, G.C., Bell, J.C.: Multi-technique surface analytical studies of automotive anti-wear films. Appl. Surf. Sci. 144–145, 222–227 (1999) Smith, G.C., Bell, J.C.: Multi-technique surface analytical studies of automotive anti-wear films. Appl. Surf. Sci. 144–145, 222–227 (1999)
135.
go back to reference Soltanahmadi, S., Morina, A., van Eijk, M.C.P., Nedelcu, I., Neville, A.: Experimental observation of zinc dialkyl dithiophosphate (ZDDP)-induced iron sulphide formation. Appl. Surf. Sci. 414, 41–51 (2017) Soltanahmadi, S., Morina, A., van Eijk, M.C.P., Nedelcu, I., Neville, A.: Experimental observation of zinc dialkyl dithiophosphate (ZDDP)-induced iron sulphide formation. Appl. Surf. Sci. 414, 41–51 (2017)
136.
go back to reference Kontou, A., Southby, M., Morgan, N., Spikes, H.A.: Influence of dispersant and ZDDP on soot wear. Tribol. Lett. 66, 157 (2018) Kontou, A., Southby, M., Morgan, N., Spikes, H.A.: Influence of dispersant and ZDDP on soot wear. Tribol. Lett. 66, 157 (2018)
137.
go back to reference Dorgham, A., Azam, A., Parsaeian, P., Wang, C., Morina, A., Neville, A.: Tribochemistry evolution of DDP tribofilms over time using in-situ synchrotron XAS. Tribol. Int. 160, 107026 (2021) Dorgham, A., Azam, A., Parsaeian, P., Wang, C., Morina, A., Neville, A.: Tribochemistry evolution of DDP tribofilms over time using in-situ synchrotron XAS. Tribol. Int. 160, 107026 (2021)
138.
go back to reference Hilgetag, G., Teichmann, H.: The alkylating properties of alkyl thiophosphates. Angew. Chemie Int. Ed. English 4, 914–922 (1965) Hilgetag, G., Teichmann, H.: The alkylating properties of alkyl thiophosphates. Angew. Chemie Int. Ed. English 4, 914–922 (1965)
139.
go back to reference Coy, R.C., Jones, R.B.: The thermal degradation and ep performance of zinc dialkyldithiophosphate additives in white oil. ASLE Trans. 24, 77–90 (1981) Coy, R.C., Jones, R.B.: The thermal degradation and ep performance of zinc dialkyldithiophosphate additives in white oil. ASLE Trans. 24, 77–90 (1981)
140.
go back to reference Jones, R.B., Coy, R.C.: The chemistry of the thermal degradation of zinc dialkyldithiophosphate additives. ASLE Trans. 24, 91–97 (1981) Jones, R.B., Coy, R.C.: The chemistry of the thermal degradation of zinc dialkyldithiophosphate additives. ASLE Trans. 24, 91–97 (1981)
141.
go back to reference Fuller, M.L.S., Kasrai, M., Bancroft, G.M., Fyfe, K., Tan, K.H.: Solution decomposition of zinc dialkyl dithiophosphate and its effect on antiwear and thermal film formation studied by X-ray absorption spectroscopy. Tribol. Int. 31, 627–644 (1998) Fuller, M.L.S., Kasrai, M., Bancroft, G.M., Fyfe, K., Tan, K.H.: Solution decomposition of zinc dialkyl dithiophosphate and its effect on antiwear and thermal film formation studied by X-ray absorption spectroscopy. Tribol. Int. 31, 627–644 (1998)
142.
go back to reference Hoshino, K., Yagishita, K., Tagawa, K., Spikes, H.: Tribological properties of sulphur-free antiwear additives zinc dialkylphosphates (ZDPs). SAE Int. J. Fuels Lubr. 5, 504–510 (2012) Hoshino, K., Yagishita, K., Tagawa, K., Spikes, H.: Tribological properties of sulphur-free antiwear additives zinc dialkylphosphates (ZDPs). SAE Int. J. Fuels Lubr. 5, 504–510 (2012)
143.
go back to reference Njiwa, P., Minfray, C., Le Mogne, T., Vacher, B., Martin, J.M., Matsui, S., Mishina, M.: Zinc dialkyl phosphate (ZP) as an anti-wear additive : comparison with ZDDP. Tribol. Lett. 44, 19–30 (2011) Njiwa, P., Minfray, C., Le Mogne, T., Vacher, B., Martin, J.M., Matsui, S., Mishina, M.: Zinc dialkyl phosphate (ZP) as an anti-wear additive : comparison with ZDDP. Tribol. Lett. 44, 19–30 (2011)
144.
go back to reference Mosey, N.J., Woo, T.K.: A quantum chemical study of the unimolecular decomposition mechanisms of zinc dialkyldithiophosphate antiwear additives. J. Phys. Chem. A 108, 6001–6016 (2004) Mosey, N.J., Woo, T.K.: A quantum chemical study of the unimolecular decomposition mechanisms of zinc dialkyldithiophosphate antiwear additives. J. Phys. Chem. A 108, 6001–6016 (2004)
145.
go back to reference Hudgens, J.J.: The structure and properties of anhydrous, alkali ultra-phosphate glasses. Iowa State University, Ames (1994) Hudgens, J.J.: The structure and properties of anhydrous, alkali ultra-phosphate glasses. Iowa State University, Ames (1994)
146.
go back to reference Ray, N.H.: Structure and properties of inorganic polymeric phosphates. Br. Polym. J. 11, 163–177 (1979) Ray, N.H.: Structure and properties of inorganic polymeric phosphates. Br. Polym. J. 11, 163–177 (1979)
147.
go back to reference Yin, Z., Kasrai, M., Fuller, M., Bancroft, G.M., Fyfe, K., Tan, K.H.: Application of soft X-ray absorption spectroscopy in chemical characterization of antiwear films generated by ZDDP Part I: the effects of physical parameters. Wear 202, 172–191 (1997) Yin, Z., Kasrai, M., Fuller, M., Bancroft, G.M., Fyfe, K., Tan, K.H.: Application of soft X-ray absorption spectroscopy in chemical characterization of antiwear films generated by ZDDP Part I: the effects of physical parameters. Wear 202, 172–191 (1997)
148.
go back to reference Willermet, P.A., Carter, R.O., III., Boulos, E.N.: Lubricant-derived tribochemical films - an infra-red spectroscopic study. Tribol. Int. 5, 371–380 (1992) Willermet, P.A., Carter, R.O., III., Boulos, E.N.: Lubricant-derived tribochemical films - an infra-red spectroscopic study. Tribol. Int. 5, 371–380 (1992)
149.
go back to reference Berkani, S., Dassenoy, F., Minfray, C., Belin, M., Vacher, B., Martin, J.M., Cardon, H., Montagnac, G., Reynard, B.: Structural changes in tribo-stressed zinc polyphosphates. Tribol. Lett. 51, 489–498 (2013) Berkani, S., Dassenoy, F., Minfray, C., Belin, M., Vacher, B., Martin, J.M., Cardon, H., Montagnac, G., Reynard, B.: Structural changes in tribo-stressed zinc polyphosphates. Tribol. Lett. 51, 489–498 (2013)
150.
go back to reference Ueda, M., Spikes, H.: ZDDP tribofilm formation and removal. Tribol. Lett. 72, 109 (2024) Ueda, M., Spikes, H.: ZDDP tribofilm formation and removal. Tribol. Lett. 72, 109 (2024)
151.
go back to reference Hanneman, W.W., Porter, R.S.: The thermal decomposition of dialkyl phosphates and O, O-dialkyl dithiophosphates. J. Org. Chem. 29, 2996–2998 (1964) Hanneman, W.W., Porter, R.S.: The thermal decomposition of dialkyl phosphates and O, O-dialkyl dithiophosphates. J. Org. Chem. 29, 2996–2998 (1964)
152.
go back to reference Berkani, S., Dassenoy, F., Minfray, C., Belin, M., Vacher, B., Martin, J.M., Cardon, H., Montagnac, G., Reynard, B.: Model formation of ZDDP tribofilm from a mixture of zinc metaphosphate and goethite. Tribol. Int. 79, 197–203 (2014) Berkani, S., Dassenoy, F., Minfray, C., Belin, M., Vacher, B., Martin, J.M., Cardon, H., Montagnac, G., Reynard, B.: Model formation of ZDDP tribofilm from a mixture of zinc metaphosphate and goethite. Tribol. Int. 79, 197–203 (2014)
153.
go back to reference Jahanmir, S.: Wear reduction and surface layer formation by a ZDDP additive. Trans. ASME J. Tribol. 109, 577–586 (1987) Jahanmir, S.: Wear reduction and surface layer formation by a ZDDP additive. Trans. ASME J. Tribol. 109, 577–586 (1987)
154.
go back to reference Mourhatch, R., Aswath, P.B.: Nanoscale properties of tribofilms formed With zinc dialkyl dithiophosphate (ZDDP) under extreme pressure condition. J. Nanosci. Nanotechnol. 9, 2682–2691 (2009)PubMed Mourhatch, R., Aswath, P.B.: Nanoscale properties of tribofilms formed With zinc dialkyl dithiophosphate (ZDDP) under extreme pressure condition. J. Nanosci. Nanotechnol. 9, 2682–2691 (2009)PubMed
155.
go back to reference Dorinson, A.: The antiwear action of zinc di–n–butyl phosphate. ASLE Trans. 22, 190–192 (1979) Dorinson, A.: The antiwear action of zinc di–n–butyl phosphate. ASLE Trans. 22, 190–192 (1979)
156.
go back to reference Yagishita, K., Igarashi, J.: Long drain/fuel efficient engine oils based on ZDTP substitute additive technology. SAE Tech. Pap. no. 2003–03–20, (2003). Yagishita, K., Igarashi, J.: Long drain/fuel efficient engine oils based on ZDTP substitute additive technology. SAE Tech. Pap. no. 2003–03–20, (2003).
157.
go back to reference Tsujimoto, T., Yaguchi, A., Yagishita, K.: Operational performance of eco-friendly engine oils formulated with the sulfur-free additive ZP. SAE Trans. J. Fuels Lubr. 116, 596–602 (2007) Tsujimoto, T., Yaguchi, A., Yagishita, K.: Operational performance of eco-friendly engine oils formulated with the sulfur-free additive ZP. SAE Trans. J. Fuels Lubr. 116, 596–602 (2007)
158.
go back to reference Hoshino, K., Yagishita, K,. Tagawa, K., Spikes, H.: Sulphur-free antiwear additives zinc dialkylphosphates (ZPs) - Film forming and friction properties. In: 17th International Colloquium Tribology Esslingen 2010 - Solving Friction and Wear Problems, 467–477, (2010) Hoshino, K., Yagishita, K,. Tagawa, K., Spikes, H.: Sulphur-free antiwear additives zinc dialkylphosphates (ZPs) - Film forming and friction properties. In: 17th International Colloquium Tribology Esslingen 2010 - Solving Friction and Wear Problems, 467–477, (2010)
159.
go back to reference Asseff, P.A.: Lubricant. US Patent 2261047, (1941) Asseff, P.A.: Lubricant. US Patent 2261047, (1941)
160.
go back to reference Thomas, J.R., Harle, O.L., Richardson, W.L., Bowman, L.: Copper-lead bearing corrosion inhibition in diesel service. Ind. Eng. Chem. 49, 703–708 (1957) Thomas, J.R., Harle, O.L., Richardson, W.L., Bowman, L.: Copper-lead bearing corrosion inhibition in diesel service. Ind. Eng. Chem. 49, 703–708 (1957)
161.
go back to reference Martin, J.M., Belin, M., Mansot, J.L.: Friction-induced amorphization with zddp—an exafs study. ASLE Trans. 29, 523–531 (1986) Martin, J.M., Belin, M., Mansot, J.L.: Friction-induced amorphization with zddp—an exafs study. ASLE Trans. 29, 523–531 (1986)
162.
go back to reference Martin, J.M.: Antiwear mechanisms of zinc dithiophosphate: a chemical hardness approach. Tribol. Lett. 6, 1–8 (1999) Martin, J.M.: Antiwear mechanisms of zinc dithiophosphate: a chemical hardness approach. Tribol. Lett. 6, 1–8 (1999)
163.
go back to reference Rounds, F.: Effects of hydroperoxides on wear as measured in four-ball wear tests. Tribol. Trans. 36, 297–303 (1993) Rounds, F.: Effects of hydroperoxides on wear as measured in four-ball wear tests. Tribol. Trans. 36, 297–303 (1993)
164.
go back to reference Habeeb, J.J., Stover, W.H.: The role of hydroperoxides in engine wear and the effect of zinc dialkyldithiophosphates. ASLE Trans. 30, 419–426 (1986) Habeeb, J.J., Stover, W.H.: The role of hydroperoxides in engine wear and the effect of zinc dialkyldithiophosphates. ASLE Trans. 30, 419–426 (1986)
165.
go back to reference Benedet, J.F.L.: Low and zero SAPS antiwear additives for engine oils. PhD Thesis, Imperial College London, London (2012) Benedet, J.F.L.: Low and zero SAPS antiwear additives for engine oils. PhD Thesis, Imperial College London, London (2012)
166.
go back to reference Konicek, A.R., Jacobs, P.W., Webster, M.N., Schilowitz, A.M.: Role of tribofilms in wear protection. Tribol. Int. 94, 4–19 (2016) Konicek, A.R., Jacobs, P.W., Webster, M.N., Schilowitz, A.M.: Role of tribofilms in wear protection. Tribol. Int. 94, 4–19 (2016)
167.
go back to reference Bennett, P.A.: A surface effect associated with the use of oils containing zinc dialkyl dithiophosphate. ASLE Trans. 2, 78–90 (1959) Bennett, P.A.: A surface effect associated with the use of oils containing zinc dialkyl dithiophosphate. ASLE Trans. 2, 78–90 (1959)
168.
go back to reference Chern, S., Ta, T., Horng, J., Wu, Y.: Wear and vibration behavior of ZDDP-containing oil considering scuffing failure. Wear 478–479, 203923 (2021) Chern, S., Ta, T., Horng, J., Wu, Y.: Wear and vibration behavior of ZDDP-containing oil considering scuffing failure. Wear 478–479, 203923 (2021)
169.
go back to reference Ueda, M., Spikes, H.A., Kadiric, A.: In-situ observation of the effect of the tribofilm growth on scuffing in rolling-sliding contact. Tribol. Lett. 70, 76 (2022) Ueda, M., Spikes, H.A., Kadiric, A.: In-situ observation of the effect of the tribofilm growth on scuffing in rolling-sliding contact. Tribol. Lett. 70, 76 (2022)
170.
go back to reference Kennedey, S., Moore. L.D.: Additive effects on lubricant fuel economy. SAE Tech. Pap. no. 872121, (1987) Kennedey, S., Moore. L.D.: Additive effects on lubricant fuel economy. SAE Tech. Pap. no. 872121, (1987)
171.
go back to reference Tripaldi, G., Vettor, A., Spikes, H.: Friction behaviour of ZDDP films in the mixed, boundary/EHD regime. SAE Tech. Pap. no. 962036, (1996) Tripaldi, G., Vettor, A., Spikes, H.: Friction behaviour of ZDDP films in the mixed, boundary/EHD regime. SAE Tech. Pap. no. 962036, (1996)
172.
go back to reference Taylor, L., Spikes, H., Camenzind, H.: Film-forming properties of zinc-based and ashless antiwear additives. SAE Tech. Pap. no. 2000–01–2030, (2000) Taylor, L., Spikes, H., Camenzind, H.: Film-forming properties of zinc-based and ashless antiwear additives. SAE Tech. Pap. no. 2000–01–2030, (2000)
173.
go back to reference Aoki, S., Suzuki, A., Masuko, M.: Comparison of sliding speed dependency of friction between steel surfaces lubricated with several ZnDTPs with different hydrocarbon moieties. Proc. Inst. Mech. Eng. Part J 220, 343–351 (2006) Aoki, S., Suzuki, A., Masuko, M.: Comparison of sliding speed dependency of friction between steel surfaces lubricated with several ZnDTPs with different hydrocarbon moieties. Proc. Inst. Mech. Eng. Part J 220, 343–351 (2006)
174.
go back to reference Zhang, J., Ueda, M., Campen, S., Spikes, H.A.: Boundary friction of ZDDP tribofilms. Tribol. Lett. 69, 8 (2021) Zhang, J., Ueda, M., Campen, S., Spikes, H.A.: Boundary friction of ZDDP tribofilms. Tribol. Lett. 69, 8 (2021)
175.
go back to reference Benyajati, C., Olver, A.V., Hamer, C.: An experimental study of micropitting, using a new miniature test-rig. Proc Leeds Lyon symposium on tribology. In: Tribology Series Transient Processes in Tribology, vol. 43, pp. 601–609. Elsevier, Amsterdam (2003) Benyajati, C., Olver, A.V., Hamer, C.: An experimental study of micropitting, using a new miniature test-rig. Proc Leeds Lyon symposium on tribology. In: Tribology Series Transient Processes in Tribology, vol. 43, pp. 601–609. Elsevier, Amsterdam (2003)
176.
go back to reference Olver, A.V., Beveridge, T.A., Laine, E.: Effect of lubricants on micropitting and wear. Tribol. Int. 41, 1049–1055 (2008) Olver, A.V., Beveridge, T.A., Laine, E.: Effect of lubricants on micropitting and wear. Tribol. Int. 41, 1049–1055 (2008)
177.
go back to reference Ueda, M., Spikes, H., Kadiric, A.: In-situ observations of the effect of the ZDDP tribofilm growth on micropitting. Tribol. Int. 138, 42–352 (2019) Ueda, M., Spikes, H., Kadiric, A.: In-situ observations of the effect of the ZDDP tribofilm growth on micropitting. Tribol. Int. 138, 42–352 (2019)
178.
go back to reference Ratoi, M., Castle, R.C., Bovington, C.H., Spikes, H.A.: The influence of soot and dispersant on ZDDP film thickness and friction. Lubr. Sci. 17, 25–43 (2004) Ratoi, M., Castle, R.C., Bovington, C.H., Spikes, H.A.: The influence of soot and dispersant on ZDDP film thickness and friction. Lubr. Sci. 17, 25–43 (2004)
179.
go back to reference Booth, J.E., Nelson, K.D., Harvey, T.J., Wood, R.J.K., Wang, L., Powrie, H.E.G., Martinez, J.G.: The feasibility of using electrostatic monitoring to identify diesel lubricant additives and soot contamination interactions by factorial analysis. Tribol. Int. 39, 1564–1575 (2006) Booth, J.E., Nelson, K.D., Harvey, T.J., Wood, R.J.K., Wang, L., Powrie, H.E.G., Martinez, J.G.: The feasibility of using electrostatic monitoring to identify diesel lubricant additives and soot contamination interactions by factorial analysis. Tribol. Int. 39, 1564–1575 (2006)
180.
go back to reference Motamen Salehi, F., Morina, A., Neville, A.: Zinc dialkyldithiophosphate additive adsorption on carbon black particles. Tribol. Lett. 66, 118 (2018) Motamen Salehi, F., Morina, A., Neville, A.: Zinc dialkyldithiophosphate additive adsorption on carbon black particles. Tribol. Lett. 66, 118 (2018)
181.
go back to reference Schilowitz, A.M., Konicek, A.R.: Role of zinc dialkyl dithiophosphate in carbon black induced abrasive wear. Wear 376–377, 771–776 (2017) Schilowitz, A.M., Konicek, A.R.: Role of zinc dialkyl dithiophosphate in carbon black induced abrasive wear. Wear 376–377, 771–776 (2017)
182.
go back to reference Olomolehin, Y.: The influence of zinc dialkyldithiophosphate and other lubricant additives on soot-induced wear. PhD Thesis, Imperial College London, London (2010) Olomolehin, Y.: The influence of zinc dialkyldithiophosphate and other lubricant additives on soot-induced wear. PhD Thesis, Imperial College London, London (2010)
183.
go back to reference Kumara, C., Meyer, H.M., Qu, J.: Material-dependent antagonistic effects between soot and ZDDP. Adv. Mater. Interfaces 7, 1901956 (2020) Kumara, C., Meyer, H.M., Qu, J.: Material-dependent antagonistic effects between soot and ZDDP. Adv. Mater. Interfaces 7, 1901956 (2020)
184.
go back to reference Kontou, A., Southby, M., Spikes, H.A.: Effect of steel hardness on soot wear. Wear 390–391, 236–245 (2017) Kontou, A., Southby, M., Spikes, H.A.: Effect of steel hardness on soot wear. Wear 390–391, 236–245 (2017)
185.
go back to reference Kirkby, T., Smith, J.J., Berryman, J., Fowell, M., Reddyhoff, T.: Soot wear mechanisms in heavy-duty diesel engine contacts. Wear 524–525, 04733 (2023) Kirkby, T., Smith, J.J., Berryman, J., Fowell, M., Reddyhoff, T.: Soot wear mechanisms in heavy-duty diesel engine contacts. Wear 524–525, 04733 (2023)
186.
go back to reference Rounds, R.: Additive interactions and their effect on the performance of a zinc dialkyl dithiophosphate. ASLE Trans. 21, 91–101 (1978) Rounds, R.: Additive interactions and their effect on the performance of a zinc dialkyl dithiophosphate. ASLE Trans. 21, 91–101 (1978)
187.
go back to reference Nicholls, M.A., Do, T., Norton, P.R., Kasrai, M., Bancroft, G.M.: Review of the lubrication of metallic surfaces by zinc dialkyl-dithiophosphates. Tribol. Int. 38, 15–39 (2005) Nicholls, M.A., Do, T., Norton, P.R., Kasrai, M., Bancroft, G.M.: Review of the lubrication of metallic surfaces by zinc dialkyl-dithiophosphates. Tribol. Int. 38, 15–39 (2005)
188.
go back to reference Mansuy, H., Beccat, P., Huiban, Y., Palermo, T., Cesbat, B.: Investigation of interaction between antiwear and dispersant additives and their effect on surface activity of ZDDP. In: Tribology series lubricants and lubrication, vol. 30, pp. 423–432. Elsevier, Amsterdam (1995) Mansuy, H., Beccat, P., Huiban, Y., Palermo, T., Cesbat, B.: Investigation of interaction between antiwear and dispersant additives and their effect on surface activity of ZDDP. In: Tribology series lubricants and lubrication, vol. 30, pp. 423–432. Elsevier, Amsterdam (1995)
189.
go back to reference Plaza, S.: The effect of other lubricating oil additives on the adsorption of zinc di-isobutyldithiophosphate on fe and γ-Fe2O3 powders. ASLE Trans. 30, 241–247 (1987) Plaza, S.: The effect of other lubricating oil additives on the adsorption of zinc di-isobutyldithiophosphate on fe and γ-Fe2O3 powders. ASLE Trans. 30, 241–247 (1987)
190.
go back to reference Kasrai, M., Fuller, M.S., Bancroft, G.M., Yamaguchi, E.S., Ryason, P.R.: X-ray absorption study of the effect of calcium sulfonate on antiwear film formation generated from neutral and basic zddps: Part 1—phosphorus species. Tribol. Trans. 46, 534–542 (2003) Kasrai, M., Fuller, M.S., Bancroft, G.M., Yamaguchi, E.S., Ryason, P.R.: X-ray absorption study of the effect of calcium sulfonate on antiwear film formation generated from neutral and basic zddps: Part 1—phosphorus species. Tribol. Trans. 46, 534–542 (2003)
191.
go back to reference Chinas-Castillo, F., Spikes, H.A.: The behavior of colloidal solid particles in elastohydrodynamic contacts. Tribol. Trans. 43, 387–394 (2000) Chinas-Castillo, F., Spikes, H.A.: The behavior of colloidal solid particles in elastohydrodynamic contacts. Tribol. Trans. 43, 387–394 (2000)
192.
go back to reference Topolovec-Miklozic, K., Lockwood, F., Spikes, H.: Behaviour of boundary lubricating additives on DLC coatings. Wear 265, 1893–1901 (2008) Topolovec-Miklozic, K., Lockwood, F., Spikes, H.: Behaviour of boundary lubricating additives on DLC coatings. Wear 265, 1893–1901 (2008)
193.
go back to reference Wan, Y., Kasrai, M., Bancroft, G.M.: X-ray absorption study of tribofilms from ZDDP and overbased salicylate detergents. Chinese Chem. Lett. 20, 119–122 (2009) Wan, Y., Kasrai, M., Bancroft, G.M.: X-ray absorption study of tribofilms from ZDDP and overbased salicylate detergents. Chinese Chem. Lett. 20, 119–122 (2009)
194.
go back to reference Yin, Z., Kasrai, M., Bancroft, G.M., Fyfe, K., Colaianni, M.L., Tan, K.H.: Application of soft X-ray absorption spectroscopy in chemical characterization of antiwear films generated by ZDDP Part II: The effect of detergents and dispersants. Wear 202, 192–201 (1997) Yin, Z., Kasrai, M., Bancroft, G.M., Fyfe, K., Colaianni, M.L., Tan, K.H.: Application of soft X-ray absorption spectroscopy in chemical characterization of antiwear films generated by ZDDP Part II: The effect of detergents and dispersants. Wear 202, 192–201 (1997)
195.
go back to reference Costello, M.T., Urrego, R.A.: Study of surface films of the ZDDP and the MoDTC with crystalline and amorphous overbased calcium sulfonates by XPS. Tribol. Trans. 50, 217–226 (2007) Costello, M.T., Urrego, R.A.: Study of surface films of the ZDDP and the MoDTC with crystalline and amorphous overbased calcium sulfonates by XPS. Tribol. Trans. 50, 217–226 (2007)
196.
go back to reference Wan, Y., Fuller, M.L.S., Kasrai, M., Bancroft, G.M., Fyfe, K., Torkelson, J.R., Hu, Y.F., Tan, K.H.: Effects of detergent on the chemistry of tribofilms from ZDDP: studied by X-ray absorption spectroscopy and XPS. In: Tribology series boundary and mixed lubrication: science and application, vol. 40, pp. 155–166. Elsevier, Amsterdam (2002) Wan, Y., Fuller, M.L.S., Kasrai, M., Bancroft, G.M., Fyfe, K., Torkelson, J.R., Hu, Y.F., Tan, K.H.: Effects of detergent on the chemistry of tribofilms from ZDDP: studied by X-ray absorption spectroscopy and XPS. In: Tribology series boundary and mixed lubrication: science and application, vol. 40, pp. 155–166. Elsevier, Amsterdam (2002)
197.
go back to reference Huq, M.Z., Chen, X., Aswath, P.B., Elsenbaumer, R.L.: Thermal degradation behavior of zinc dialkyldithiophosphate in presence of catalyst and detergents in neutral oil. Tribol. Lett. 19, 127–134 (2005) Huq, M.Z., Chen, X., Aswath, P.B., Elsenbaumer, R.L.: Thermal degradation behavior of zinc dialkyldithiophosphate in presence of catalyst and detergents in neutral oil. Tribol. Lett. 19, 127–134 (2005)
198.
go back to reference Yu, L.G., Yamaguchi, E.S., Kasrai, M., Bancroft, G.M.: The chemical characterization of tribofilms using XANES - interaction of nanosize calciumcontaining detergents with zinc dialkyldithiophosphate. Can. J. Chem. 85, 675–684 (2007) Yu, L.G., Yamaguchi, E.S., Kasrai, M., Bancroft, G.M.: The chemical characterization of tribofilms using XANES - interaction of nanosize calciumcontaining detergents with zinc dialkyldithiophosphate. Can. J. Chem. 85, 675–684 (2007)
199.
go back to reference Kaneko, T., Yamamori, K., Suzuki, H., Onodera, K., Ogano, S.: Friction reduction technology for low viscosity engine oil compatible with LSPI prevention performance. SAE Tech. Pap. no. 2016–01–22, (2016) Kaneko, T., Yamamori, K., Suzuki, H., Onodera, K., Ogano, S.: Friction reduction technology for low viscosity engine oil compatible with LSPI prevention performance. SAE Tech. Pap. no. 2016–01–22, (2016)
200.
go back to reference Taylor, L.J., Spikes, H.A.: Friction-enhancing properties of zddp antiwear additive: Part 1-friction and morphology of zddp reaction films. Tribol. Trans. 46, 303–309 (2003) Taylor, L.J., Spikes, H.A.: Friction-enhancing properties of zddp antiwear additive: Part 1-friction and morphology of zddp reaction films. Tribol. Trans. 46, 303–309 (2003)
201.
go back to reference Topolovec-Miklozic, K., Forbus, T.R., Spikes, H.A.: Film forming and friction properties of overbased calcium sulphonate detergents. Tribol. Lett. 29, 33–44 (2008) Topolovec-Miklozic, K., Forbus, T.R., Spikes, H.A.: Film forming and friction properties of overbased calcium sulphonate detergents. Tribol. Lett. 29, 33–44 (2008)
202.
go back to reference Gallopoulos, N.E., Murphy, C.K.: Interactions between a zinc dialkylphosphorodithioate and lubricating oil dispersants. ASLE Trans. 14, 1–7 (1971) Gallopoulos, N.E., Murphy, C.K.: Interactions between a zinc dialkylphosphorodithioate and lubricating oil dispersants. ASLE Trans. 14, 1–7 (1971)
203.
go back to reference Shiomi, M., Tokashiki, M., Tomizawa, H., Kuribayashi, T.: Interaction between zinc dialkyldithiophosphate and amine. Lubr. Sci. 1, 131–147 (1989) Shiomi, M., Tokashiki, M., Tomizawa, H., Kuribayashi, T.: Interaction between zinc dialkyldithiophosphate and amine. Lubr. Sci. 1, 131–147 (1989)
204.
go back to reference Ramakumar, S.S.V., Aggarwal, N., Rao, M., Sarpal, A.S., Srivastava, S.P., Bhatnagar, A.K.: Studies on additive-additive interactions: effect of dispersant and antioxidant additives on the synergistic combination of overbased sulphonate and ZDDP. Lubr. Sci. 7, 25–38 (1994) Ramakumar, S.S.V., Aggarwal, N., Rao, M., Sarpal, A.S., Srivastava, S.P., Bhatnagar, A.K.: Studies on additive-additive interactions: effect of dispersant and antioxidant additives on the synergistic combination of overbased sulphonate and ZDDP. Lubr. Sci. 7, 25–38 (1994)
205.
go back to reference Kapur, G.S., Chopra, A., Ramakumar, S.S.V., Sarpal, A.S.: Molecular spectroscopic studies of ZDDP-PIBS interactions. Lubr. Sci. 10, 309–321 (1998) Kapur, G.S., Chopra, A., Ramakumar, S.S.V., Sarpal, A.S.: Molecular spectroscopic studies of ZDDP-PIBS interactions. Lubr. Sci. 10, 309–321 (1998)
206.
go back to reference Yamaguchi, E.S., Roby, S.H., Francisco, M.M., Ruelas, S.G., Godfrey, D.: Antiwear film formation by zndtp, detergent, and dispersant components of passenger car motor oils. Tribol. Trans. 45, 425–429 (2002) Yamaguchi, E.S., Roby, S.H., Francisco, M.M., Ruelas, S.G., Godfrey, D.: Antiwear film formation by zndtp, detergent, and dispersant components of passenger car motor oils. Tribol. Trans. 45, 425–429 (2002)
207.
go back to reference Matsui, Y., Aoki, S., Masuko, M.: Elucidation of the action of functional groups in the coexisting ashless compounds on the tribofilm formation and friction characteristic of zinc dialkyldithiophosphate-formulated lubricating oils. Tribol. Trans. 61, 220–228 (2018) Matsui, Y., Aoki, S., Masuko, M.: Elucidation of the action of functional groups in the coexisting ashless compounds on the tribofilm formation and friction characteristic of zinc dialkyldithiophosphate-formulated lubricating oils. Tribol. Trans. 61, 220–228 (2018)
208.
go back to reference Massoud, T., De Matos, R.P., Le Mogne, T., Belin, M., Cobian, M., Thiebaut, B., Loehle, S., Dahlem, F., Minfray, C.: Effect of ZDDP on lubrication mechanisms of linear fatty amines under boundary lubrication conditions. Tribol. Int. 141, 105954 (2020) Massoud, T., De Matos, R.P., Le Mogne, T., Belin, M., Cobian, M., Thiebaut, B., Loehle, S., Dahlem, F., Minfray, C.: Effect of ZDDP on lubrication mechanisms of linear fatty amines under boundary lubrication conditions. Tribol. Int. 141, 105954 (2020)
209.
go back to reference Soltanahmadi, S., Esfahani, E.A., Nedelcu, I., Morina, A., van Eijk, M.C.P., Neville, A.: Surface reaction films from amine-based organic friction modifiers and their influence on surface fatigue and friction. Tribol. Lett. 67, 80 (2019) Soltanahmadi, S., Esfahani, E.A., Nedelcu, I., Morina, A., van Eijk, M.C.P., Neville, A.: Surface reaction films from amine-based organic friction modifiers and their influence on surface fatigue and friction. Tribol. Lett. 67, 80 (2019)
210.
go back to reference Cyriac, F., Yi, T.X., Poornachary, S.K., Chow, P.S.: Boundary lubrication performance of polymeric and organic friction modifiers in the presence of an anti-wear additive. Tribol. Int. 165, 107256 (2022) Cyriac, F., Yi, T.X., Poornachary, S.K., Chow, P.S.: Boundary lubrication performance of polymeric and organic friction modifiers in the presence of an anti-wear additive. Tribol. Int. 165, 107256 (2022)
211.
go back to reference Shen, W., Hirayama, T., Yamashita, N., Adachi, M., Oshio, T., Tsuneoka, H., Tagawa, K., Yagishita, K., Yamada, N.L.: Relationship between interfacial adsorption of additive molecules and reduction of friction coefficient in the organic friction modifiers-ZDDP combinations. Tribol. Int. 167, 107365 (2022) Shen, W., Hirayama, T., Yamashita, N., Adachi, M., Oshio, T., Tsuneoka, H., Tagawa, K., Yagishita, K., Yamada, N.L.: Relationship between interfacial adsorption of additive molecules and reduction of friction coefficient in the organic friction modifiers-ZDDP combinations. Tribol. Int. 167, 107365 (2022)
212.
go back to reference Spikes, H.: Friction modifier additives. Tribol. Lett. 60, 5 (2015) Spikes, H.: Friction modifier additives. Tribol. Lett. 60, 5 (2015)
213.
go back to reference De Barros Bouchet, M.I., Martin, J.M., Le Mogne, T., Bilas, P., Vacher, B., Yamada, Y.: Mechanisms of MoS2 formation by MoDTC in presence of ZnDTP: effect of oxidative degradation. Wear 258, 1643–1650 (2005) De Barros Bouchet, M.I., Martin, J.M., Le Mogne, T., Bilas, P., Vacher, B., Yamada, Y.: Mechanisms of MoS2 formation by MoDTC in presence of ZnDTP: effect of oxidative degradation. Wear 258, 1643–1650 (2005)
214.
go back to reference Gorbatchev, O., Bouchet, M.D.B., Martin, J.M., Léonard, D., Le-Mogne, T., Iovine, R., Thiebaut, B., Héau, C.: Friction reduction efficiency of organic Mo-containing FM additives associated to ZDDP for steel and carbon-based contacts. Tribol. Int. 99, 278–288 (2016) Gorbatchev, O., Bouchet, M.D.B., Martin, J.M., Léonard, D., Le-Mogne, T., Iovine, R., Thiebaut, B., Héau, C.: Friction reduction efficiency of organic Mo-containing FM additives associated to ZDDP for steel and carbon-based contacts. Tribol. Int. 99, 278–288 (2016)
215.
go back to reference Morina, A., Neville, A., Priest, M., Green, J.H.: ZDDP and MoDTC interactions in boundary lubrication - the effect of temperature and ZDDP/MoDTC ratio. Tribol. Int. 39, 1545–1557 (2006) Morina, A., Neville, A., Priest, M., Green, J.H.: ZDDP and MoDTC interactions in boundary lubrication - the effect of temperature and ZDDP/MoDTC ratio. Tribol. Int. 39, 1545–1557 (2006)
216.
go back to reference Morina, A., Neville, A., Priest, M., Green, J.H.: ZDDP and MoDTC interactions and their effect on tribological performance - tribofilm characteristics and its evolution. Tribol. Lett. 24, 243–256 (2006) Morina, A., Neville, A., Priest, M., Green, J.H.: ZDDP and MoDTC interactions and their effect on tribological performance - tribofilm characteristics and its evolution. Tribol. Lett. 24, 243–256 (2006)
217.
go back to reference Yamamoto, Y., Gondo, S., Kamakura, T., Konishi, M.: Organoamine and organophosphate molybdenum complexes as lubricant additives. Wear 120, 51–60 (1987) Yamamoto, Y., Gondo, S., Kamakura, T., Konishi, M.: Organoamine and organophosphate molybdenum complexes as lubricant additives. Wear 120, 51–60 (1987)
218.
go back to reference Huai, W., Chen, X., Lu, F., Zhang, C., Ma, L., Wen, S.: Tribological properties of sulfur- and phosphorus-free organic molybdenum compound as additive in oil. Tribol. Int. 141, 105944 (2020) Huai, W., Chen, X., Lu, F., Zhang, C., Ma, L., Wen, S.: Tribological properties of sulfur- and phosphorus-free organic molybdenum compound as additive in oil. Tribol. Int. 141, 105944 (2020)
219.
go back to reference Nedelcu, I., Piras, F., Rossi, A., Pasaribu, H.R.: XPS analysis on the influence of water on the evolution of zinc dialkyldithiophosphate-derived reaction layer in lubricated rolling contacts. Surf. Interface Anal. 44, 1219–1224 (2012) Nedelcu, I., Piras, F., Rossi, A., Pasaribu, H.R.: XPS analysis on the influence of water on the evolution of zinc dialkyldithiophosphate-derived reaction layer in lubricated rolling contacts. Surf. Interface Anal. 44, 1219–1224 (2012)
220.
go back to reference Cen, H., Morina, A., Neville, A., Pasaribu, R., Nedelcu, I.: Effect of water on ZDDP anti-wear performance and related tribochemistry in lubricated steel/steel pure sliding contacts. Tribol. Int. 56, 47–57 (2012) Cen, H., Morina, A., Neville, A., Pasaribu, R., Nedelcu, I.: Effect of water on ZDDP anti-wear performance and related tribochemistry in lubricated steel/steel pure sliding contacts. Tribol. Int. 56, 47–57 (2012)
221.
go back to reference Parsaeian, P., Ghanbarzadeh, A., Wilson, M., Van Eijk, M.C.P., Nedelcu, I., Dowson, D., Neville, A., Morina, A.: An experimental and analytical study of the effect of water and its tribochemistry on the tribocorrosive wear of boundary lubricated systems with ZDDP-containing oil. Wear 358–359, 23–31 (2016) Parsaeian, P., Ghanbarzadeh, A., Wilson, M., Van Eijk, M.C.P., Nedelcu, I., Dowson, D., Neville, A., Morina, A.: An experimental and analytical study of the effect of water and its tribochemistry on the tribocorrosive wear of boundary lubricated systems with ZDDP-containing oil. Wear 358–359, 23–31 (2016)
222.
go back to reference Dorgham, A., Azam, A., Parsaeian, P., Khan, T., Sleiman, M., Wang, C., Morina, A., Neville, A.: Understanding the effect of water on the transient decomposition of zinc dialkyldithiophosphate (ZDDP). Tribol. Int. 157, 106855 (2021) Dorgham, A., Azam, A., Parsaeian, P., Khan, T., Sleiman, M., Wang, C., Morina, A., Neville, A.: Understanding the effect of water on the transient decomposition of zinc dialkyldithiophosphate (ZDDP). Tribol. Int. 157, 106855 (2021)
223.
go back to reference Dorgham, A., Azam, A., Parsaeian, P., Wang, C., Morina, A., Neville, A.: An assessment of the effect of relative humidity on the decomposition of the ZDDP antiwear additive. Tribol. Lett. 69, 1–12 (2021) Dorgham, A., Azam, A., Parsaeian, P., Wang, C., Morina, A., Neville, A.: An assessment of the effect of relative humidity on the decomposition of the ZDDP antiwear additive. Tribol. Lett. 69, 1–12 (2021)
224.
go back to reference Costa, H.L., Spikes, H.A.: Impact of ethanol on the formation of antiwear tribofilms from engine lubricants. Tribol. Int. 93, 364–376 (2016) Costa, H.L., Spikes, H.A.: Impact of ethanol on the formation of antiwear tribofilms from engine lubricants. Tribol. Int. 93, 364–376 (2016)
225.
go back to reference Pagkalis, K., Spikes, H., Jelita, J., Marc, R., Kadiric, A.: The influence of steel composition on the formation and effectiveness of anti-wearfilms in tribological contacts. Tribol. Lett. 69, 75 (2021) Pagkalis, K., Spikes, H., Jelita, J., Marc, R., Kadiric, A.: The influence of steel composition on the formation and effectiveness of anti-wearfilms in tribological contacts. Tribol. Lett. 69, 75 (2021)
226.
go back to reference Zhu, F., Xu, J., Han, X., Shen, Y., Jin, M.: Deposit formation on chromium-plated cylinder liner in a fully formulated oil. Proc IMechE Part J 230, 1415–1422 (2016) Zhu, F., Xu, J., Han, X., Shen, Y., Jin, M.: Deposit formation on chromium-plated cylinder liner in a fully formulated oil. Proc IMechE Part J 230, 1415–1422 (2016)
227.
go back to reference Sheasby, T.S., Caughlin, T.A., Mackwood, W.A.: A comparison of the boundary lubrication of 52100 steel, zirconia and silicon nitride by S, P, S/P and zinc dialkyl dithiophosphate additives. Wear 196, 100–109 (1996) Sheasby, T.S., Caughlin, T.A., Mackwood, W.A.: A comparison of the boundary lubrication of 52100 steel, zirconia and silicon nitride by S, P, S/P and zinc dialkyl dithiophosphate additives. Wear 196, 100–109 (1996)
228.
go back to reference Nicholls, M.A., Norton, P.R., Bancroft, G.M., Kasrai, M., De Stasio, G.: Spatially resolved nanoscale chemical and mechanical characterization of ZDDP antiwear films on aluminum – silicon alloys under cylinder / bore wear conditions. Tribol. Lett. 18, 261–278 (2005) Nicholls, M.A., Norton, P.R., Bancroft, G.M., Kasrai, M., De Stasio, G.: Spatially resolved nanoscale chemical and mechanical characterization of ZDDP antiwear films on aluminum – silicon alloys under cylinder / bore wear conditions. Tribol. Lett. 18, 261–278 (2005)
229.
go back to reference Xia, X., Morina, A., Neville, A., Priest, M., Roshan, R., Warrens, C.P., Payne, M.J.: Tribological performance of an Al – Si alloy lubricated in the boundary regime with zinc dialkyldithiophosphate and molybdenum dithiocarbamate additives. Proc I. Mech. E Part J 222, 305–314 (2008) Xia, X., Morina, A., Neville, A., Priest, M., Roshan, R., Warrens, C.P., Payne, M.J.: Tribological performance of an Al – Si alloy lubricated in the boundary regime with zinc dialkyldithiophosphate and molybdenum dithiocarbamate additives. Proc I. Mech. E Part J 222, 305–314 (2008)
230.
go back to reference Dienwiebel, M., Pohlmann, K., Scherge, M.: Origins of the wear resistance of AlSi cylinder bore surfaces studies by surface analytical tools. Tribol. Int. 40, 597–1602 (2007) Dienwiebel, M., Pohlmann, K., Scherge, M.: Origins of the wear resistance of AlSi cylinder bore surfaces studies by surface analytical tools. Tribol. Int. 40, 597–1602 (2007)
231.
go back to reference Shimizu, Y., Spikes, H.A.: The influence of aluminium–silicon alloy on ZDDP tribofilm formation on the counter-surface. Tribol. Lett. 65, 137 (2017) Shimizu, Y., Spikes, H.A.: The influence of aluminium–silicon alloy on ZDDP tribofilm formation on the counter-surface. Tribol. Lett. 65, 137 (2017)
232.
go back to reference Equey, S., Roos, S., Mueller, U., Hauert, R., Spencer, N.D., Crockett, R.: Tribofilm formation from ZnDTP on diamond-like carbon. Wear 264, 316–321 (2008) Equey, S., Roos, S., Mueller, U., Hauert, R., Spencer, N.D., Crockett, R.: Tribofilm formation from ZnDTP on diamond-like carbon. Wear 264, 316–321 (2008)
233.
go back to reference Vengudusamy, B., Green, G.J.H., Lamb, G.D., Spikes, H.A.: Influence of hydrogen and tungsten concentration on the tribological properties of DLC/DLC contacts with ZDDP. Wear 298–299, 109–119 (2013) Vengudusamy, B., Green, G.J.H., Lamb, G.D., Spikes, H.A.: Influence of hydrogen and tungsten concentration on the tribological properties of DLC/DLC contacts with ZDDP. Wear 298–299, 109–119 (2013)
234.
go back to reference Vengudusamy, B.: Behaviour of lubricant additives on DLC coatings. PhD Thesis, Imperial College London, London (2011) Vengudusamy, B.: Behaviour of lubricant additives on DLC coatings. PhD Thesis, Imperial College London, London (2011)
235.
go back to reference Vengudusamy, B., Green, J.H., Lamb, G.D., Spikes, H.A.: Durability of ZDDP tribofilms formed in DLC/DLC contacts. Tribol. Lett. 51, 469–478 (2013) Vengudusamy, B., Green, J.H., Lamb, G.D., Spikes, H.A.: Durability of ZDDP tribofilms formed in DLC/DLC contacts. Tribol. Lett. 51, 469–478 (2013)
236.
go back to reference Bartolomé, L., Oblak, E., Kalin, M.: Mechanical behaviour and constitutive models of ZDDP tribofilms on DLC coatings using nano-indentation data and finite element modelling. Tribol. Int. 95, 19–26 (2016) Bartolomé, L., Oblak, E., Kalin, M.: Mechanical behaviour and constitutive models of ZDDP tribofilms on DLC coatings using nano-indentation data and finite element modelling. Tribol. Int. 95, 19–26 (2016)
237.
go back to reference Okubo, H., Watanabe, S., Tadokoro, C., Sasaki, S.: Effects of structure of zinc dialkyldithiophosphates on tribological properties of tetrahedral amorphous carbon film under boundary lubrication. Tribol. Int. 98, 26–40 (2016) Okubo, H., Watanabe, S., Tadokoro, C., Sasaki, S.: Effects of structure of zinc dialkyldithiophosphates on tribological properties of tetrahedral amorphous carbon film under boundary lubrication. Tribol. Int. 98, 26–40 (2016)
238.
go back to reference Okubo, H., Watanabe, S., Tadokoro, C., Hirata, Y., Sasaki, S.: Effects of molecular structure of zinc dialkyldithiophosphates on tribological properties of a hydrogenated amorphous carbon film under boundary lubrication. Tribol. Online 12, 221–228 (2017) Okubo, H., Watanabe, S., Tadokoro, C., Hirata, Y., Sasaki, S.: Effects of molecular structure of zinc dialkyldithiophosphates on tribological properties of a hydrogenated amorphous carbon film under boundary lubrication. Tribol. Online 12, 221–228 (2017)
239.
go back to reference Andersson, J., Larsson, R., Almqvist, A., Grahn, M., Minami, I.: Semi-deterministic chemo-mechanical model of boundary lubrication. RSC Faraday Discuss. 156, 343–360 (2012) Andersson, J., Larsson, R., Almqvist, A., Grahn, M., Minami, I.: Semi-deterministic chemo-mechanical model of boundary lubrication. RSC Faraday Discuss. 156, 343–360 (2012)
240.
go back to reference Ghanbarzadeh, A., Wilson, M., Morina, A., Dowson, D., Neville, A.: Development of a new mechano-chemical model in boundary lubrication. Tribol. Int. 93, 573–582 (2016) Ghanbarzadeh, A., Wilson, M., Morina, A., Dowson, D., Neville, A.: Development of a new mechano-chemical model in boundary lubrication. Tribol. Int. 93, 573–582 (2016)
241.
go back to reference Ghanbarzadeh, A., Piras, E., Nedelcu, I., Brizmer, V., Wilson, M.C., Morina, A., Dowson, D., Neville, A.: Zinc dialkyl dithiophosphate antiwear tribo film and its effect on the topography evolution of surfaces - a numerical and experimental study. Wear 362–363, 186–198 (2016) Ghanbarzadeh, A., Piras, E., Nedelcu, I., Brizmer, V., Wilson, M.C., Morina, A., Dowson, D., Neville, A.: Zinc dialkyl dithiophosphate antiwear tribo film and its effect on the topography evolution of surfaces - a numerical and experimental study. Wear 362–363, 186–198 (2016)
242.
go back to reference Akchurin, A., Bosman, R.: A deterministic stress-activated model for tribo-film growth and wear simulation. Tribol. Lett. 65, 59 (2017) Akchurin, A., Bosman, R.: A deterministic stress-activated model for tribo-film growth and wear simulation. Tribol. Lett. 65, 59 (2017)
243.
go back to reference Azam, A., Ghanbarzadeh, A., Neville, A., Morina, A., Wilson, M.C.T.: Modelling tribochemistry in the mixed lubrication regime. Tribol. Int. 132, 265–274 (2019) Azam, A., Ghanbarzadeh, A., Neville, A., Morina, A., Wilson, M.C.T.: Modelling tribochemistry in the mixed lubrication regime. Tribol. Int. 132, 265–274 (2019)
244.
go back to reference Azam, A., Dorgham, A., Parsaeian, P., Morina, A., Neville, A., Wilson, M.C.T.: The mutual interaction between tribochemistry and lubrication - Interfacial mechanics of tribo film. Tribol. Int. 135, 161–169 (2019) Azam, A., Dorgham, A., Parsaeian, P., Morina, A., Neville, A., Wilson, M.C.T.: The mutual interaction between tribochemistry and lubrication - Interfacial mechanics of tribo film. Tribol. Int. 135, 161–169 (2019)
245.
go back to reference Wang, Y., Azam, A.: Towards optimum additive performance : A numerical study to understand the influence of roughness parameters on the zinc dialkyldithiophosphates tribofilm growth. Lubr. Sci. 33, 1–14 (2021) Wang, Y., Azam, A.: Towards optimum additive performance : A numerical study to understand the influence of roughness parameters on the zinc dialkyldithiophosphates tribofilm growth. Lubr. Sci. 33, 1–14 (2021)
246.
go back to reference Chen, Z., Gu, C., Tian, T.: Modeling of formation and removal of ZDDP tribofilm on rough surfaces. Tribol. Lett. 69, 13 (2021) Chen, Z., Gu, C., Tian, T.: Modeling of formation and removal of ZDDP tribofilm on rough surfaces. Tribol. Lett. 69, 13 (2021)
247.
go back to reference Yang, J., Pu, A.W., Pei, X.: A deterministic tribofilm growth model considering flash temperature and surface roughness in mixed lubrication. Surf. Topogr. Metrol. Prop. 9, 025017 (2021) Yang, J., Pu, A.W., Pei, X.: A deterministic tribofilm growth model considering flash temperature and surface roughness in mixed lubrication. Surf. Topogr. Metrol. Prop. 9, 025017 (2021)
248.
go back to reference Lyu, B., Zhang, L., Meng, X., Wang, C.: A boundary lubrication model and experimental study considering ZDDP tribofilms on reciprocating friction pairs. Tribol. Lett. 70, 65 (2022) Lyu, B., Zhang, L., Meng, X., Wang, C.: A boundary lubrication model and experimental study considering ZDDP tribofilms on reciprocating friction pairs. Tribol. Lett. 70, 65 (2022)
249.
go back to reference Bulgarevich, S.B., Boiko, M.V., Olesnikov, V.I., Korets, K.E.: Population of transition states of triboactivated chemical processes. J. Friction Wear 31, 288–293 (2010) Bulgarevich, S.B., Boiko, M.V., Olesnikov, V.I., Korets, K.E.: Population of transition states of triboactivated chemical processes. J. Friction Wear 31, 288–293 (2010)
250.
go back to reference Martini, A., Eder, S.J., Dörr, N.: Tribochemistry: a review of reactive molecular dynamics simulations. Lubricants 8, 1–20 (2020) Martini, A., Eder, S.J., Dörr, N.: Tribochemistry: a review of reactive molecular dynamics simulations. Lubricants 8, 1–20 (2020)
251.
go back to reference Ta, H.T.T., Tran, N.V., Tieu, A.K., Zhu, H., Yu, H., Ta, T.D.: Computational tribochemistry: a review from classical and quantum mechanics studies. J. Phys. Chem. C 125, 16875–16891 (2021) Ta, H.T.T., Tran, N.V., Tieu, A.K., Zhu, H., Yu, H., Ta, T.D.: Computational tribochemistry: a review from classical and quantum mechanics studies. J. Phys. Chem. C 125, 16875–16891 (2021)
252.
go back to reference Ayestarán Latorre, C., Remias, J.E., Moore, J.D., Spikes, H.A., Dini, D., Ewen, J.P.: Mechanochemistry of phosphate esters confined between sliding iron surfaces. Commun. Chem. 4, 1–11 (2021) Ayestarán Latorre, C., Remias, J.E., Moore, J.D., Spikes, H.A., Dini, D., Ewen, J.P.: Mechanochemistry of phosphate esters confined between sliding iron surfaces. Commun. Chem. 4, 1–11 (2021)
253.
go back to reference Jiang, S., Dasgupta, S., Blanco, M., Frazier, R., Yamaguchi, E.S., Tang, Y., Goddard, W.A.: Structures, vibrations, and force fields of dithiophosphate wear inhibitors from ab initio quantum chemistry. J. Phys. Chem. 100, 15760–15769 (1996) Jiang, S., Dasgupta, S., Blanco, M., Frazier, R., Yamaguchi, E.S., Tang, Y., Goddard, W.A.: Structures, vibrations, and force fields of dithiophosphate wear inhibitors from ab initio quantum chemistry. J. Phys. Chem. 100, 15760–15769 (1996)
254.
go back to reference Armstrong, D.R., Ferrari, E.S., Roberts, K.J., Adams, D.: An investigation into the molecular stability of zinc di-alkyl-di-thiophosphates (ZDDPs) in relation to their use as anti-wear and anti-corrosion additives in lubricating oils. Wear 208, 138–146 (1997) Armstrong, D.R., Ferrari, E.S., Roberts, K.J., Adams, D.: An investigation into the molecular stability of zinc di-alkyl-di-thiophosphates (ZDDPs) in relation to their use as anti-wear and anti-corrosion additives in lubricating oils. Wear 208, 138–146 (1997)
255.
go back to reference Jiang, S., Frazier, R., Yamaguchi, E.S., Blanco, M., Dasgupta, S., Zhou, Y., Cagin, T., Tang, Y., Goddard, W.A.: The SAM model for wear inhibitor performance of dithiophosphates on iron oxide. J. Phys. Chem. B 101, 7702–7709 (1997) Jiang, S., Frazier, R., Yamaguchi, E.S., Blanco, M., Dasgupta, S., Zhou, Y., Cagin, T., Tang, Y., Goddard, W.A.: The SAM model for wear inhibitor performance of dithiophosphates on iron oxide. J. Phys. Chem. B 101, 7702–7709 (1997)
256.
go back to reference Mosey, N.J., Woo, T.K.: Finite temperature structure and dynamics of zinc dialkyldithiophosphate wear inhibitors: a density functional theory and ab initio molecular dynamics study. J. Phys. Chem. A 107, 5058–5070 (2003) Mosey, N.J., Woo, T.K.: Finite temperature structure and dynamics of zinc dialkyldithiophosphate wear inhibitors: a density functional theory and ab initio molecular dynamics study. J. Phys. Chem. A 107, 5058–5070 (2003)
257.
go back to reference Mosey, N.J., Woo, T.K.: Insights into the chemical behavior of zinc dialkyldithiophosphate anti-wear additives in their isomeric and decomposed forms through molecular simulation. Tribol. Int. 39, 979–993 (2006) Mosey, N.J., Woo, T.K.: Insights into the chemical behavior of zinc dialkyldithiophosphate anti-wear additives in their isomeric and decomposed forms through molecular simulation. Tribol. Int. 39, 979–993 (2006)
258.
go back to reference Mosey, N.J., Müser, M.H., Woo, T.K.: Molecular mechanisms for the functionality of lubricant additives. Science 307, 1612–1615 (2005)PubMed Mosey, N.J., Müser, M.H., Woo, T.K.: Molecular mechanisms for the functionality of lubricant additives. Science 307, 1612–1615 (2005)PubMed
259.
go back to reference Gauvin, M., Dassenoy, F., Minfray, C., Martin, J.M., Montagnac, G., Reynard, B.: Zinc phosphate chain length study under high hydrostatic pressure by Raman spectroscopy. J. Appl. Phys. 101, 063505 (2007) Gauvin, M., Dassenoy, F., Minfray, C., Martin, J.M., Montagnac, G., Reynard, B.: Zinc phosphate chain length study under high hydrostatic pressure by Raman spectroscopy. J. Appl. Phys. 101, 063505 (2007)
260.
go back to reference Tse, J.S., Song, Y., Liu, Z.: Effects of temperature and pressure on ZDDP. Tribol. Lett. 28, 45–49 (2007) Tse, J.S., Song, Y., Liu, Z.: Effects of temperature and pressure on ZDDP. Tribol. Lett. 28, 45–49 (2007)
261.
go back to reference Gauvin, M., Minfray, C., Belin, M., Aquilanti, G., Martin, J.M., Dassenoy, F.: Pressure-induced amorphization of zinc orthophosphate-Insight in the zinc coordination by XAS. Tribol. Int. 67, 222–228 (2013) Gauvin, M., Minfray, C., Belin, M., Aquilanti, G., Martin, J.M., Dassenoy, F.: Pressure-induced amorphization of zinc orthophosphate-Insight in the zinc coordination by XAS. Tribol. Int. 67, 222–228 (2013)
262.
go back to reference Mosey, N.J., Woo, T.K.: An ab initio molecular dynamics and density functional theory study of the formation of phosphate chains from metathiophosphates. Inorg. Chem. 45, 7464–7479 (2006)PubMed Mosey, N.J., Woo, T.K.: An ab initio molecular dynamics and density functional theory study of the formation of phosphate chains from metathiophosphates. Inorg. Chem. 45, 7464–7479 (2006)PubMed
263.
go back to reference Onodera, T., Morita, Y., Suzuki, A., Sahnoun, R., Koyama, M., Tsuboi, H., Hatakeyama, N., Endou, A., Takaba, H., Kubo, M., Del Carpio, C.A.: A theoretical investigation on the abrasive wear prevention mechanism of ZDDP and ZP tribofilms. Appl. Surf. Sci. 254, 7976–7979 (2008) Onodera, T., Morita, Y., Suzuki, A., Sahnoun, R., Koyama, M., Tsuboi, H., Hatakeyama, N., Endou, A., Takaba, H., Kubo, M., Del Carpio, C.A.: A theoretical investigation on the abrasive wear prevention mechanism of ZDDP and ZP tribofilms. Appl. Surf. Sci. 254, 7976–7979 (2008)
264.
go back to reference Minfray, C., Le Mogne, T., Martin, J.M., Onodera, T., Nara, S., Takahashi, S., Tsuboi, H., Koyama, M., Endou, A., Takaba, H., Kubo, M.: Experimental and molecular dynamics simulations of tribochemical reactions with ZDDP : zinc phosphate – iron oxide reaction. Tribol. Trans. 51, 589–601 (2008) Minfray, C., Le Mogne, T., Martin, J.M., Onodera, T., Nara, S., Takahashi, S., Tsuboi, H., Koyama, M., Endou, A., Takaba, H., Kubo, M.: Experimental and molecular dynamics simulations of tribochemical reactions with ZDDP : zinc phosphate – iron oxide reaction. Tribol. Trans. 51, 589–601 (2008)
265.
go back to reference Onodera, T., Martin, J.M., Minfray, C., Dassenoy, F., Miyamoto, A.: Antiwear chemistry of ZDDP: coupling classical MD and tight-binding quantum chemical MD methods (TB-QCMD). Tribol. Lett. 50, 31–39 (2013) Onodera, T., Martin, J.M., Minfray, C., Dassenoy, F., Miyamoto, A.: Antiwear chemistry of ZDDP: coupling classical MD and tight-binding quantum chemical MD methods (TB-QCMD). Tribol. Lett. 50, 31–39 (2013)
266.
go back to reference Onodera, T., et al.: Influence of nanometer scale film structure of ZDDP tribofilm on its mechanical properties: a computational chemistry study. Appl. Surf. Sci. 256, 976–979 (2009) Onodera, T., et al.: Influence of nanometer scale film structure of ZDDP tribofilm on its mechanical properties: a computational chemistry study. Appl. Surf. Sci. 256, 976–979 (2009)
267.
go back to reference Salinas Ruiz, V.R., Kuwahara, T., Galipaud, J., Masenelli-Varlot, K., Hassine, M.B., Héau, C., Stoll, M., Mayrhofer, L., Moras, G., Martin, J.M., Moseler, M.: Interplay of mechanics and chemistry governs wear of diamond-like carbon coatings interacting with ZDDP-additivated lubricants. Nat. Commun. 12, 1–15 (2021) Salinas Ruiz, V.R., Kuwahara, T., Galipaud, J., Masenelli-Varlot, K., Hassine, M.B., Héau, C., Stoll, M., Mayrhofer, L., Moras, G., Martin, J.M., Moseler, M.: Interplay of mechanics and chemistry governs wear of diamond-like carbon coatings interacting with ZDDP-additivated lubricants. Nat. Commun. 12, 1–15 (2021)
268.
go back to reference Peeters, S., Barlini, A., Jain, J., Nand, N., Righi, M.C.: Applied Surface Science Adsorption and decomposition of ZDDP on lightweight metallic substrates : Ab initio and experimental insights. Appl. Surf. Sci. 600, 153947 (2022) Peeters, S., Barlini, A., Jain, J., Nand, N., Righi, M.C.: Applied Surface Science Adsorption and decomposition of ZDDP on lightweight metallic substrates : Ab initio and experimental insights. Appl. Surf. Sci. 600, 153947 (2022)
269.
go back to reference Benini, F., Restuccia, P., Righi, M.C.: Zinc dialkyldithiophosphates adsorption and dissociation on ferrous substrates: an ab initio study. Appl. Surf. Sci. 642, 158419 (2024) Benini, F., Restuccia, P., Righi, M.C.: Zinc dialkyldithiophosphates adsorption and dissociation on ferrous substrates: an ab initio study. Appl. Surf. Sci. 642, 158419 (2024)
Metadata
Title
Mechanisms of ZDDP—An Update
Author
Hugh Spikes
Publication date
01-03-2025
Publisher
Springer US
Published in
Tribology Letters / Issue 1/2025
Print ISSN: 1023-8883
Electronic ISSN: 1573-2711
DOI
https://doi.org/10.1007/s11249-025-01968-3

Premium Partners