Next Article in Journal
Adhesion and Electron Properties of Quasi-2D Mo2C, Ti2C, and V2C MXene Flakes after Van Der Waals Adsorption of Alcohol Molecules: Influence of Humidity
Previous Article in Journal
Theoretical Study and Adsorption Behavior of Urea on Mild Steel in Automotive Gas Oil (AGO) Medium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Macroscale Superlubricity of Black Phosphorus Quantum Dots

1
School of Mechanical and Electrical Engineering, Xi’an University of Architecture and Technology, Xi’an 710055, China
2
School of Economics and Management, Beijing Jiaotong University, Beijing 100044, China
3
School of Metallurgy Engineering, Xi’an University of Architecture and Technology, Xi’an 710055, China
4
School of Mechanical, Electronic and Control Engineering, Beijing Jiaotong University, Beijing 100044, China
*
Author to whom correspondence should be addressed.
Lubricants 2022, 10(7), 158; https://doi.org/10.3390/lubricants10070158
Submission received: 9 June 2022 / Revised: 30 June 2022 / Accepted: 12 July 2022 / Published: 15 July 2022

Abstract

:
In the present work, Black Phosphorus Quantum Dots (BPQDs) were synthesized via sonication-assisted liquid-phase exfoliation. The average size of the BPQDs was 3.3 ± 0.85 nm. The BPQDs exhibited excellent dispersion stability in ultrapure water. Macroscale superlubricity was realized with the unmodified BPQDs on rough Si3N4/SiO2 interfaces. A minimum coefficient of friction (COF) of 0.0022 was achieved at the concentration of 0.015 wt%. In addition, the glycerol was introduced to promote the stability of the superlubricity state. The COF of the BPQDs-Glycerol aqueous solution (BGaq) was 83.75% lower than that of the Glycerol aqueous solution (Gaq). Based on the above analysis, the lubrication model was presented. The hydrogen-bonded network and silica gel layer were formed on the friction interface, which played a major role in the realization of macroscale superlubricity. In addition, the adsorption water layer could also prevent the worn surfaces from making contact with each other. Moreover, the synergistic effect between BPQDs and glycerol could significantly decrease the COF and maintain the superlubricity state. The findings theoretically support the realization of macroscale superlubricity with unmodified BPQDs as a water-based lubrication additive.

1. Introduction

Superlubricity, which is an ideal state (coefficient of friction (COF) < 0.01), was described in 1990 by Hirano and Shinjo [1]. Previous research indicates that superlubricity can be classified into two groups based on the lubrication materials. First, solid superlubricity describes a state in which the COF is lower than 0.01 between two sliding interfaces. Solid lubricants with incommensurate structures or weak interlayer interactions, such as diamond-like carbon film [2], MoS2 [3], carbon nitride film [4,5] and graphite [6], were used. Although superlubricity can be achieved via solid lubricants, the process needs to be carried out under special conditions with relatively perfect materials. In comparison, the liquid superlubricity described by the second group indicates a state of superlubricity that can be obtained via liquid lubricants. This superlubricity can be achieved easily, continuously and steadily on a macroscopic scale under the proper atmospheric conditions.
Until now, superlubricity could be realized by various liquid lubricant systems, such as hydrophilic polymer brushes [7], ionic liquid [8,9], polyhydroxy aqueous solutions [10,11,12,13], acid-based solutions [14,15,16] and two-dimensional (2D) nano-additives, after a period of running-in. Compared with other liquid lubricant systems, the introduction of 2D nano-additives efficiently enhances the contact pressure of the lubricants. Therefore, 2D nano-additives such as graphene oxide [17], MXene [18], double hydroxide nanosheets [19] and black phosphorus [20,21] show great potential for liquid superlubricity. Among these 2D nano-additives, black phosphorus (BP) has inspired a great deal of curiosity from researchers in the lubricant field owing to its excellent tribological performance and extreme pressure and anti-wear property. Wang et al. [22] showed that macroscale superlubricity can be obtained with BP-OH modified by NaOH. It was revealed that the addition of BP-OH can result in macroscale superlubricity under different conditions, even on the roughest friction surface. Ren et al. [23] reported that stable superlubricity could be achieved with oxidized black phosphorus nanosheets. Macroscale superlubricity could also be achieved with aqueous ethylene glycol via the introduction of BPQDs under high contact pressure (336 MPa) [24]. Wu et al. [25,26] reported on the superlubricity of degraded black phosphorus, demonstrating that the degradation of black phosphorus significantly favors lubrication. However, there is scant research on the use of unmodified BPQDs as a water-based lubrication additive to realize macroscale superlubricity.
In the present study, macroscale superlubricity could be realized by unmodified BPQDs on a rough silicon nitride (Si3N4)/silicon dioxide (SiO2) surface. This is the first work to show that unmodified BPQDs can be used as a water-based lubrication additive to achieve macroscale superlubricity. However, macroscale superlubricity with BPQDs could not be maintained for a longer period. To solve the issue of the unstable macroscale superlubricity behavior of BPQD aqueous solution (Baq), glycerol was introduced into ultrapure water. Macroscale superlubricity remained steady for more than 1 h when the glycerol was introduced. In addition, the systematic tribological performance of Baq and BPQDs-glycerol aqueous solution (BGaq) were also tested. The chemical composition of a worn surface was examined by X-ray photoelectron spectroscopy (XPS). The morphology of the worn surface was characterized via scanning electron microscopy (SEM) and via optical microscope. The superlubricity mechanisms were proposed and analyzed in detail.

2. Materials and Methods

2.1. Preparation of BPQDs, Baq and BGaq

The black phosphorus powder was fabricated via the high-energy ball milling method from red phosphorus (RP, > 99.9%) which has already been reported by our group [27,28]. The BPQDs were prepared from bulk BP by sonication-assisted liquid-phase exfoliation. Figure 1a illustrates the synthesis process of the BPQDs. 100 mg of BP powder was added into 80 mL of N-methyl pyrrolidone (≥ 99.5%) for 6 h probe sonication. During the process of probe sonication, an ice bath was used to maintain the temperature below 0 °C. Subsequently, the processed solution was centrifuged, then the supernatant was gathered and centrifuged, and the precipitate was BPQDs. Subsequently, the precipitate was washed with ethanol (AR) and ultrapure water (1.0 mL, 18.2 MΩ) through centrifugation for 2~3 times. After the BPQDs fabrication, four concentrations of the Baq were prepared by complete dispersion of BPQDs (400, 200, 150, 100 μg) in ultrapure water via sonication at 25 °C for 30 min. In addition, the glycerol (AR) was introduced. The glycerol aqueous (Gaq) was obtained by mixing glycerol and ultrapure water with the ratio 1:5. Then, the BPQDs were dispersed in the Gaq via sonication for 30 min, which was named BGaq.

2.2. Characterization of BPQDs

The crystal structures of BPQDs were observed via X-ray diffraction (XRD, Bruker D8) in the 2θ range of 5–90°. The XRD instrument with Cu Kα radiation using 40 kV and 40 mA and scan rate of 5° min−1 was used. The chemical composition of BPQDs was determined by Raman spectroscopy (HORIBA Evolution JYT6400, HORIBA Scientific, France). The morphology of BPQDs was evaluated via transmission electron microscopy (TEM, FEI Talos F200S, Thermo Fisher Scientific Inc., US) operated at 200 kV acceleration voltage. The thickness of the BPQDs was characterized via atomic force microscopy (AFM, Bruker Dimension Icon, Germany). The stability of the Baq was evaluated via Zeta meter (Malvern Zetasizer Nano ZS90, Malvern Panalytical, UK).

2.3. Tribological Tests

The universal microtribotester (UMT-5, Bruker, Germany) was used to evaluate the tribological performance. The Si3N4 ball (Φ = 10 mm or Φ = 4 mm, Ra ≈ 54 nm) and SiO2 plate (15 × 15 × 5 mm, Ra ≈ 0.674 μm) were used as friction pair. During the tribological tests, the different volumes (100 μL, 50 μL and 20 μL) of the lubricant were introduced. A load of 0.5–3.5 N (401–890 MPa) was applied. The sliding speeds in this work were in the range 31.4–157 mm·s −1. Three independent tribological tests were carried out. The tribological tests were done at room temperature.

2.4. Characterization of the Worn Surfaces

After the tribological tests, the morphologies of the worn surfaces were characterized via the 3D white-light interferometer (Bruker, Contour GT-K, Germany) and SEM (Gemini 300, Zeiss, Germany) and optical microscope (GX51, Olympus Corporation, Japan). The wear volume was measured via the 3D white-light interferometer, and the wear rate was calculated as [29].
W B = V P S
where W B is the wear rate, V is the wear volume of ball and plate, P is the normal load and S is the sliding distance. Further, the chemical properties of the worn surface were analyzed via XPS (PHI-5000 VersaProbe III, ThermoFischer, US). The Al-Kα radiation (1486.6 eV) was used as the excitation source. The XPS spectra were calibrated by the C 1s peak. The XPS instrument used 14.6 kV and 13.5 mA.

3. Results and Discussion

The crystal structure, chemical properties and morphologies of BPQDs were characterized via XRD, Raman, TEM and AFM (Figure 1b–i). The TEM image revealed that the BPQDs have a relatively spherical shape (Figure 1b,c). A lattice fringe of 0.21 was observed, which can be attributed to the BPQDs [30]. The average diameter of BPQDs was about 3.3 ± 0.85 nm, as shown in the diameter distribution histogram (Figure 1d). In addition, the XRD spectrum exhibited characteristic peaks at 16.5°, 26.45° and 35.0°, corresponding to the orthorhombic BP (JCPSD No. 21−1272). The characteristic peaks on the Raman spectrum located at 366.2, 431.5 and 459.4 cm−1 were attributed to A g 1 , B 2 g and A g 2 [31]. The AFM images of the BPQDs, as shown in Figure 1g–i, revealed the average thickness of the BPQDs was 3.18 ± 0.144 nm. As mentioned above, the results revealed that the BPQDs were synthesized successfully via sonication-assisted liquid-phase exfoliation.
Dispersity plays a key role in the realization of the excellent lubrication performance. To evaluate the stability of the Baq, the Zeta potential was measured by Zeta meter. A higher absolute value (35 mV) of Zeta potential was obtained, which indicated that the Baq is a stable suspension (Figure 1j) [32,33]. Furthermore, the dispersity of the Baq was also evaluated via optical images (Figure 2). A series of the dispersions with different concentrations were obtained via sonication, the colors of dispersion became darker as the concentration increased, as shown in Figure 2a. When the dispersion settled for 7 days, no evident separation and sediment appeared (Figure 2b) and the color of the dispersion changed only slightly. This indicates that the stability of the Baq was excellent, which is consistent with the value of Zeta potential.
The tribological tests were carried out at the Si3N4/SiO2 interface under different conditions. Firstly, the Baq lubricant and Si3N4 balls with a diameter of 10 mm were used in this tribological test. Before the tribological test, the 100 μL lubricant was introduced into the friction surface. The COF curve under the lubrication condition of ultrapure water and Baq (0.015 wt%) is shown in Figure 3a. Unless otherwise stated, the load of 3 N and the sliding speed of 62.8 mm∙s−1 was used in tribological tests. Under the lubrication condition of ultrapure water, the COF decreased from 0.45 to 0.28 at first, then suddenly increased to about 0.5 and fluctuated in the range of 0.45–0.65. The high COF value for ultrapure water can be attributed to the evaporation of free water molecules and the lack of formation of an effective lubrication film [34]. By contrast, the COF dramatically decreased when the Baq was introduced. After about 3100 s of a running-in period, the COF entered the region of superlubricity (approximately 0.0022). Furthermore, the superlubricity performance of Baq was also evaluated under different concentrations, loads and sliding speeds. Obviously, macroscale superlubricity was achieved for all the tests (Figure 3b–d). The average COF decreased firstly then increased with the increasing of concentration and load (Figure 3b,c), reaching the minimum value of 0.0022 under the concentration of 0.015 wt% and the load of 3 N. It was revealed that the BPQDs could trigger macroscale superlubricity. The trend in COF curve variation under the lubrication of Baq suggested the possible formation of the lubrication films during the running-in period, making it easy to realize macroscale superlubricity. However, the macroscale superlubricity with Baq could not be maintained for a longer period; the macroscale superlubricity period was about 1100 s (Figure 3a). In the end, the COF curve sharply increased to 0.6. This suggests that the lubrication film was possibly broken in this case.
In order to solve the issue of unstable superlubricity under the lubrication condition of Baq, glycerol was introduced into the ultrapure water. As a contrast, the tribological test under the lubrication condition of Gaq was carried out. After the introduction of glycerol to the ultrapure water, the COF gradually reduced from 0.28 to 0.032 and then fluctuated in a range of 0.03–0.045 (Figure 4a). It was revealed that the lubrication films were unable achieve macroscale superlubricity. When the BPQDs were added into the Gaq, the COF gradually decreased to 0.01 and then entered the superlubricity state (COF < 0.01). In addition, the superlubricity performance of BGaq was evaluated under different lubricant volumes (Figure 4b). The running-in time reduced from 1320 s to 380 s when the lubricant volume decreased from 100 μL to 20 μL. The running-in time is dependent on the evaporation of the free water molecules; the larger the initial volume of lubricant, the longer the running-in time. By contrast, the average COF was increased from 0.0026 to 0.007 with the decrease in lubricant volume. In the subsequent section, the 20 μL volume of the lubricant was selected for a more elaborate discussion. Furthermore, macroscale superlubricity remained steady for more than 1 h under the lubrication condition of BGaq. The tribological tests of different conditions were conducted, and the Φ 4 mm ball was used in the following section. Macroscale superlubricity can be observed for all conditions with the BGaq after the running-in period, as shown in Figure 4c–h. With the increase of the concentration from 0.01 to 0.04 wt%, the COF firstly decreased from 0.0076 to 0.0065, and then increased to 0.008, under the lubrication condition of BGaq. The minimum COF (~0.0065) was achieved under the concentration of 0.015 wt% (Figure 4c,d). The average COF was 83.75% less than that of Gaq. Furthermore, the minimum COF value was achieved at a load of 0.5 N (~0.004). Notably, the COF first decreased then increased, and the running-in time reduced from 500 to 300 s with the increase of sliding speed from 62.8 to 157 mm·s−1 (Figure 4g,h). The reduction of the running-in time revealed that the high sliding speed favors the formation of effective lubrication film. Stability is important when evaluating the quality of a lubricant. In order to investigate the stability of BGaq, the optical image and superlubricity performance were evaluated after one week. After one week, no significant difference in color, and no distinct sediment could be observed. Furthermore, the macroscale superlubricity also could be obtained with BGaq after a week, and the average COF value increased slightly, as shown in Figure 4i. The results revealed that the BGaq has better stability and dispersity.
After the tribological tests, the morphology of the worn surface was characterized via optical microscope, SEM and 3D white-light interferometer. The worn surface on the Si3N4 balls and SiO2 plate under the lubrication condition of ultrapure water and Baq are shown in Figure 5, respectively. When the ultrapure water was utilized, the wear scar on the Si3N4 ball lubricated with ultrapure water was extremely rough. The diameter of the wear scar was 803.8 μm, and there were deep furrows on the worn surface, which indicated that the wear mechanism was mainly abrasive wear. In comparison, the furrows were shallow and the worn surface became smooth after the introduction of BPQDs in ultrapure water. In addition, the diameter of the wear scar under the lubrication condition of Baq was 463.9 μm, which was 42.3 % less than that of the ultrapure water. Furthermore, the wear volume was measured via 3D white-light interferometer (Table 1) and the wear rate was calculated (Figure 5g). The wear rates of SiO2 plate and Si3N4 ball lubricated by ultrapure water were 7.9323 × 10−8 and 6.05298 × 10−9 mm3/(N∙m), respectively. By comparison, the wear rates of the SiO2 plate and the Si3N4 ball lubricated by Baq were decreased by 78.3 % and 87.2 %. The results revealed that the water molecules will remain on the worn surface due to the excellent water retention property of BPQDs. Therefore, starved lubrication or boundary lubrication would occur. Even though the wear rate under the lubrication of Baq was less than that of ultrapure water, severe wear had also occurred during the tests. Similarly, when the glycerol was introduced into ultrapure water, shallow furrows and wear debris appeared (Figure 6b). By contrast, the wear debris disappeared and the worn surface became smoothly lubricated by BGaq. The diameter of the wear scar on the Si3N4 ball under the lubrication condition of ultrapure water, Gaq and BGaq were 574.4, 280.9 and 264.4 μm (Figure 6a–c), respectively. The diameter of the wear scar under the lubrication condition of BGaq was 54.3 % less than that of ultrapure water. Similarly, the width of the worn track on the SiO2 plate was obviously decreased (Figure 6d–i) after the introduction of glycerol and BPQDs. The worn track had a depth of 5.85 μm and a width of 0.645 mm under the lubrication condition of ultrapure water. The width of worn track under the lubrication condition of BGaq was 0.395 mm, which was 38.8% less than that of ultrapure water. However, the width of the worn track was increased slightly under the lubrication condition of BGaq, compared with Gaq (0.21 mm). The depth of worn surface on the SiO2 plate under the lubrication conditions of Gaq and BGaq were 0.685 and 0.355 μm, respectively. Furthermore, the wear volume was measured via 3D profilometer (Table 2), and the wear rates were calculated (Figure 6j). The wear rates of the SiO2 plate and Si3N4 ball lubricated by Gaq were 1.41543 × 10−8 and 3.38337 × 10−10 mm3/(N∙m), respectively. By comparison, the wear rates of plate and ball under the lubrication condition of BGaq were 9.46568 × 10−9 and 2.57577 × 10−10 mm3/(N∙m), which were 33.1 % and 23.9 % lower than that of Gaq. It was revealed that serious wear occurred under the lubrication condition of Gaq. In addition, the wear rates of ball and plate were calculated after lubrication by BGaq under different conditions (Figure 7). The wear rate of the SiO2 plate decreased with the increase of the concentration, when the BGaq was used as lubricant. The wear rate of the Si3N4 ball increased with the concentration. The wear rate of the SiO2 plate decreased with the increase in normal load, and the wear rate of the Si3N4 ball first decreased and then increased with the increase of load. In addition, with the increase of sliding speed, the wear rate of the SiO2 ball first decreased then increased.
To further clarify the lubrication mechanism, the chemical composition of the SiO2 plate-worn surface was analyzed via XPS. The full-scan XPS spectra of the worn surfaces lubricated by Baq and BGaq are shown in Figure 8a,f. The C 1s, O 1s, P 2p and Si 2p both could be detected. The chemical composition of the worn surface on the SiO2 plate after the tribological test (unclean) with the Baq and BGaq are shown in Figure 8. The C 1s three peaks under the lubrication condition of Baq (BGaq) are located at 283.4 (283.2), 284.8 (284.6) and 286.9 (286.5) eV, corresponding to C−Si [17], C−H/C=H [27] and C−O bond [17], respectively. The O 1s peak at 531.2 eV is attributed to P-OH [28] for the worn surface lubricated by the Baq. For the worn surface lubricated by BGaq, the positions of O 1s peak is located at 530.9 eV, corresponding to the P=O bond [26]. For the worn surface under the lubrication condition of Baq, the Si 2p shows the peak at 101.9 eV, assigned as the Si−O, which is indicative of polymerized siloxane [22,35]. In comparison, the double peaks on the Si 2p at 101.5 and 100.2 eV are assigned to Si−O and Si−OH [17]. For the spectrum of P 2p on the worn surface lubricated by Baq, the double peaks at 131.6 and 132.5 eV are identified as P−O−P and O−P=O [23]. For the worn surface lubricated by BGaq, the peak located at 131.6 eV is attributed to bridging (P−O−P) phosphorus-oxygen bond.
Some research reported that the bonds of P=O and P−OH are favorable to combining with free water molecules [23,25]. In the present work, the bonds of P−OH and P=O were detected on the worn surface of the SiO2 plate under the lubrication condition of Baq and BGaq, respectively. Therefore, the free water layers could have been formed on the worn surface during the tribological test. Furthermore, the greater number of combined water molecules could have caused the massive P−OH bonds to be released, due to the bond breakage of P−O−P. The water layers also could form via hydrogen bonds of the P=O∙∙∙HOH and P−OH∙∙∙OH2. The adsorption and retention of free water molecules for BPQDs play a significant role in achieving and sustaining a superlubricity state. In addition, the P=O bonds have a greater inclination to immediately form two hydrogen bonds, allowing the hydrogen-bonded network to be formed via the hydrogen-bond interactions [24]. For the system of ceramics materials, the colloidal silicon oxide film was formed by tribochemical reactions during the running-in period as follows [18,36,37].
Si3N4 + 6H2O → 3SiO2 + 4NH3
SiO2 + H2O = Si(OH)4
In addition, the hydrolysis reaction could take place due to the thermal activation during the tribological tests. The hydrolysis reaction was conducted as follows [22,38].
≡ Si−O−Si ≡ + H2O → ≡ Si−OH ∙∙∙ OH−Si
The Si-O bond on the XPS spectrum revealed the formation of a silica gel layer on the surface of the SiO2 plate. In addition, the existence of Si-OH/Si-O bonds on the worn surface lubricated by BGaq demonstrated the formation of silica layers on the worn surface [26]. Beyond that, the existence of the Si-OH bonds allows for the absorption of free water molecules, which is favored for the formation of the water layer on the worn surface. The minimum lubrication film thickness (hmin) under the lubrication condition of BGaq was calculated by the Hamrock–Dowson theory [39,40].
h min = 2.56 α 0.53 R 0.464 u 0.67 η 0 0.67 E ' 0.073 W 0.067 ( 1 0.61 e 0.73 k )
where R is the equivalent radius of the ball, α is the pressure-viscosity coefficient of the BGaq (about 4 × 10−9 Pa−1) [41] and u is the sliding speed of the Si3N4 ball and SiO2 plate (62.8 mm/s). The applied load W is 3 N. η 0 is the dynamic lubricant viscosity. E is the effective modulus of elasticity of the friction pair, which can calculated as follows:
1 E ' = 1 2 [ 1 v 1 2 E 1 + 1 v 2 2 E 2 ]
where v1 and v2 are the Poisson’s ratio of the Si3N4 ball (v1 = 0.26) and SiO2 plate (v2 = 0.17), respectively. E1 and E2 are the elasticity modulus of the ball (E1 = 310 GPa) and plate (E2 = 77.8 GPa). The R is the equivalent radius of the ball, which can be calculated as follows:
R = E d 3 6 W
where d is the diameter of the worn scar of the ball, and the E is the effective elastic modulus of the friction pair, which can be calculated by the following:
1 E = 1 v 1 2 E 1 + 1 v 2 2 E 2
In the end, the calculation of hc is about 34.4 nm with a load of 3 N and a sliding speed of 62.8 mm/s. Then the λ can be calculated by the ratio of the theoretical minimum film thickness (hmin) to the roughness as follows:
λ = h min σ 1 2 + σ 2 2
where σ1 and σ2 are the roughness of the Si3N4 ball and SiO2 plate. Thus, the λ < 1 in this present study, which indicated that the superlubricity state corresponds to boundary lubrication.
Nevertheless, a superlubricity state is difficult to sustain under the lubrication condition of Baq (Figure 3a). To solve this issue, glycerol was selected to promote the stability of the superlubricity state. The reason that glycerol can promote the stability of the friction is that the water molecules can be aggregated on the worn surface due to the degradation of glycerol. In addition, it can adsorb molecular water from humid air [42]. To exclude the influence of the glycerol on superlubricity, the tribological performance of Gaq was evaluated, and the result indicated that the macroscale superlubricity could not be achieved on the Si3N4/SiO2 interface. However, the synergistic effect between BPQDs and glycerol could effectively decrease the COF and maintain the superlubricity state.
As mentioned above, the lubrication mechanisms were proposed, and the schematic diagram is shown in Figure 9. During the tribological tests, the boundary lubrication occurred in the contact zone. The colloidal silicon oxide film was formed due to the tribochemical reaction, and the silica gel layer was formed due to the hydrolysis reaction. The colloidal silicon oxide film can promote boundary lubrication [37]. Additionally, the hydrogen-bonded network was formed during the tribological reaction via the hydrogen-bond interactions (Figure 9b,d). On the one hand, the hydroxyl group can be attached to the surface of the silica gel layer via the hydrogen bond interactions, which can form the hydrogen-bonded network. On the other hand, the existence of the P=O bonds are favor the formation of two hydrogen bonds, which can be attached to the surface of BPQDs, thereby forming the hydrogen-bonded network. Moreover, the adsorption water layers were formed via the combination of P=O/P−OH and free water molecules. Meanwhile, the Si−OH group on the worn surface facilitated the formation of the free water layer. The existence of the hydrogen-bonded network and adsorption water layer played a major role in the reduction of the average COF. Furthermore, the hydrogen-bonded network and adsorption water layer could prevent the worn surfaces from contacting one another. In addition, the synergistic effect between BPQDs and glycerol could maintain a low and stable COF.

4. Conclusions

BPQDs were successfully prepared via sonication-assisted liquid-phase exfoliation. The average particle size of the BPQDs was 3.3 ± 0.85 nm. The BPQDs exhibited excellent dispersion stability in ultrapure water. Macroscale superlubricity could be realized with the unmodified BPQDs on the roughest interfaces. The minimum COF of 0.0022 was achieved at the concentration of 0.015 wt% (3 N, 62.8 mm∙s−1). The wear rates of the SiO2 plate and Si3N4 ball were 78.3 % and 87.2 % lower than that of ultrapure water. Furthermore, glycerol was introduced into the ultrapure water to promote the stability of the superlubricity state. Similarly, the macroscale superlubricity could be obtained with the BGaq, and the superlubricity remained steady for more than 1 h. The wear rates of plate and ball under the lubrication condition of BGaq were 9.46568 × 10−9 and 2.57577 × 10−10 mm3/(N·m), which were 33.1% and 23.9% lower than that of Gaq. In addition, the macroscale superlubricity state could still be achieved after one week, and the average COF value increased slightly. Ultimately, the macroscale superlubricity mechanisms were proposed. During the tribological tests, the colloidal silicon oxide film and the silica gel layer were formed due to the tribochemical reaction and hydrolysis reaction. The hydrogen-bonded network was formed via hydrogen-bond interactions. The silica gel layer and hydrogen-bonded network play major roles in realizing macroscale superlubricity. Additionally, the adsorption water layers could also reduce the COF during the tribological tests. The synergistic effect between BPQDs and glycerol can effectively decrease the COF and maintain the superlubricity state.

Author Contributions

P.G. and W.W. designed the investigation. Y.Q. and F.L. prepared all sample and materials for analysis. P.G. and Y.Q. collected and analysed the date. Y.Q. and J.J. performed statistical analysis and collected field data. P.G., F.L. and W.W. interpreted the results and wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support from the National Natural Science Foundation of China (Grant No. 51975450), Youth Science and Technology New Star Project of Shaanxi Province Innovation Ability Support Plan (2021KJXX-32) and Advanced technology research program of Xi’an (21XJZZ0031).

Institutional Review Board Statement

No applicable.

Informed Consent Statement

No applicable.

Data Availability Statement

The data used to support the findings of this study are available from the corresponding author upon request.

Acknowledgments

We thank instructional support specialists of Xi’an university of Architecture and Technology laboratory.

Conflicts of Interest

The authors declare that there are no conflict of interests.

Nomenclature

ΦThe diameter of the Si3N4 ball
RaThe roughness of the friction pairs
WBWear rate
VWear volume
PNormal load
SSliding distance of the Si3N4 ball
hminThe minimum lubrication film thickness
αPressure-viscosity coefficient
REquivalent radius of the Si3N4 ball
uSliding speed of the Si3N4 ball and SiO2 plate
η 0 Dynamic lubricant viscosity
EEffective modulus of elasticity of the friction pair
WApplied load
viPoisson’s ratio of the friction pair
EiElasticity modulus of the friction pair
EEffective elastic modulus of the friction pair
dDiameter of the worn scar of the ball
λThe ratio of the theoretical minimum film thickness to the roughness
σThe roughness of the friction pair after the tests

References

  1. Hirano, M.; Shinjo, K. Atomistic locking and friction. Phys. Rev. B 1990, 41, 11837–11851. [Google Scholar] [CrossRef] [PubMed]
  2. Erdemir, A.; Eryilmaz, O. Achieving superlubricity in dlc films by controlling bulk, surface, and tribochemistry. Friction 2014, 2, 140–155. [Google Scholar] [CrossRef] [Green Version]
  3. Zhao, Y.; Mei, H.; Chang, P.; Yang, Y.; Huang, W.; Liu, Y.; Cheng, L.; Zhang, L. 3D-Printed Topological Mos2/Mose2heterostructures for Macroscale Superlubricity. Acs Appl. Mater. Interfaces 2021, 13, 34984–34995. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, J.; Wang, Y. Superlubricity in Carbon Nanostructural Films: From Mechanisms to Modulating Strategies. In Superlubricity; Elsevier Bv: Amsterdam, The Netherlands, 2021. [Google Scholar]
  5. Bakoglidis, K.D.; Palisaitis, J.; Dos Santos, R.B.; Rivelino, R.; Persson, P.O.; Gueorguiev, G.K.; Hultman, L. Self-Healing in carbon nitride evidenced as material inflation and superlubric behavior. Acs Appl. Mater. Interfaces 2018, 10, 16238–16243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Liu, Y.; Wang, K.; Xu, Q.; Zhang, J.; Hu, Y.; Ma, T.; Zheng, Q.; Luo, J. Superlubricity between graphite layers in ultrahigh vacuum. Acs Appl. Mater. Interfaces 2020, 12, 43167–43172. [Google Scholar] [CrossRef] [PubMed]
  7. Røn, T.; Javakhishvili, I.; Hvilsted, S.; Jankova, K.; Lee, S. Ultralow friction with hydrophilic polymer brushes in water as segregated from silicone matrix. Adv. Mater. Interfaces 2016, 3, 1500472. [Google Scholar] [CrossRef] [Green Version]
  8. Fajardo, O.Y.; Bresme, F.; Kornyshev, A.A.; Urbakh, M. Electrotunable lubricity with ionic liquid nanoscale films. Sci. Rep. 2015, 5, 7698. [Google Scholar] [CrossRef] [Green Version]
  9. Zhou, Y.; Qu, J. Ionic Liquids as lubricant additives: A review. Acs Appl. Mater. Interfaces 2017, 9, 3209–3222. [Google Scholar] [CrossRef]
  10. Ge, X.; Li, J.; Zhang, C.; Luo, J. Liquid superlubricity of polyethylene glycol aqueous solution achieved with boric acid additive. Langmuir 2018, 34, 3578–3587. [Google Scholar] [CrossRef]
  11. Wang, H.; Liu, Y.; Li, J.; Luo, J. Investigation of superlubricity achieved by polyalkylene glycol aqueous solutions. Adv. Mater. Interfaces 2016, 3, 1600531. [Google Scholar] [CrossRef]
  12. Ma, Q.; He, T.; Khan, A.M.; Wang, Q.; Chung, Y.W. Achieving macroscale liquid superlubricity using glycerol aqueous solutions. Tribol. Int. 2021, 160, 107006. [Google Scholar] [CrossRef]
  13. Liu, W.; Wang, H.; Liu, Y.; Li, J.; Erdemir, A.; Luo, J. Mechanism of superlubricity conversion with polyalkylene glycol aqueous solutions. Langmuir 2019, 35, 11784–11790. [Google Scholar] [CrossRef] [PubMed]
  14. Deng, M.; Zhang, C.; Li, J.; Ma, L.; Luo, J. Hydrodynamic Effect on the superlubricity of phosphoric acid between ceramic and sapphire. Friction 2014, 2, 173–181. [Google Scholar] [CrossRef] [Green Version]
  15. Li, J.; Zhang, C.; Deng, M.; Luo, J. Investigations of the superlubricity of sapphire against ruby under phosphoric acid lubrication. Friction 2014, 2, 164–172. [Google Scholar] [CrossRef] [Green Version]
  16. Xiao, C.; Li, J.; Chen, L.; Zhang, C.; Zhou, N.; Qian, L.; Luo, J. Speed Dependence of liquid superlubricity stability with H3PO4 solution. R. Soc. Chem. 2017, 7, 49337–49343. [Google Scholar] [CrossRef] [Green Version]
  17. Yi, S.; Chen, X.; Li, J.; Liu, Y.; Ding, S.; Luo, J. Macroscale Superlubricity of Si-Doped Diamond-Like Carbon Film Enabled by Graphene Oxide As Additives. Carbon 2021, 176, 358–366. [Google Scholar] [CrossRef]
  18. Yi, S.; Li, J.; Liu, Y.; Ge, X.; Zhang, J.; Luo, J. In-Situ Formation of Tribofilm with Ti3c2tx Mxene Nanoflakes Triggers Macroscale Superlubricity. Tribol. Int. 2021, 154, 106695. [Google Scholar] [CrossRef]
  19. Wang, H.; Liu, Y.; Liu, W.; Liu, Y.; Wang, K.; Li, J.; Ma, T.; Eryilmaz, O.L.; Shi, Y.; Erdemir, A.; et al. Superlubricity of Polyalkylene Glycol Aqueous Solutions Enabled by Ultrathin Layered Double Hydroxide Nanosheets. Acs Appl. Mater. Interfaces 2019, 11, 20249–20256. [Google Scholar] [CrossRef]
  20. Liu, Y.; Li, J.; Li, J.; Yi, S.; Ge, X.; Zhang, X.; Luo, J. Shear-Induced interfacial structural conversion triggers macroscale superlubricity: From black phosphorus nanoflakes to phosphorus oxide. Acs Appl. Mater. Interfaces 2021, 13, 31947–31956. [Google Scholar] [CrossRef]
  21. Tang, G.; Wu, Z.; Su, F.; Wang, H.; Xu, X.; Li, Q.; Ma, G.; Chu, P.K. Macroscale Superlubricity on engineering steel in the presence of black phosphorus. Nano Lett. 2021, 21, 5308–5315. [Google Scholar] [CrossRef]
  22. Wang, W.; Xie, G.; Luo, J. Superlubricity of black phosphorus as lubricant additive. Acs Appl. Mater. Interfaces 2018, 10, 43203–43210. [Google Scholar] [CrossRef] [PubMed]
  23. Ren, X.; Yang, X. Superlubricity under ultrahigh contact pressure enabled by partially oxidized black phosphorus nanosheets. Npj 2D Mater. Appl. 2021, 5, 44. [Google Scholar] [CrossRef]
  24. Ren, X.; Yang, X.; Xie, G.; Luo, J. Black phosphorus quantum dots in aqueous ethylene glycol for macroscale superlubricity. Acs Appl. Nano Mater. 2020, 3, 4799–4809. [Google Scholar] [CrossRef]
  25. Wu, S.; He, F.; Xie, G.; Bian, Z.; Luo, J.; Wen, S. Black Phosphorus: Degradation favors lubrication. Nano Lett. 2018, 18, 5618–5627. [Google Scholar] [CrossRef] [PubMed]
  26. Wu, S.; He, F.; Xie, G.; Bian, Z.; Ren, Y.; Liu, X.; Yang, H.; Guo, D.; Zhang, L.; Wen, S.; et al. Super-Slippery degraded black phosphorus/silicon dioxide interface. Acs Appl. Mater. Interfaces 2020, 12, 7717–7726. [Google Scholar] [CrossRef]
  27. Wang, Q.; Hou, T.; Wang, W.; Zhang, G.; Gao, Y.; Wang, K. Tribological Behavior of black phosphorus nanosheets as water-based lubrication additives. Friction 2022, 10, 374–387. [Google Scholar] [CrossRef]
  28. Wang, Q.; Hou, T.; Wang, W.; Zhang, G.; Gao, Y.; Wang, K. Tribological properties of black phosphorus nanosheets as oil-based lubricant additives for titanium alloy-steel contacts: Lubricaiton of black phosphene. R. Soc. Open Sci. 2020, 7, 200530. [Google Scholar] [CrossRef]
  29. Tang, G.; Su, F.; Xu, X.; Chu, P.K. 2d Black Phosphorus dotted with silver nanoparticles: An excellent lubricant additive for tribological applications. Chem. Eng. J. 2020, 392, 123631. [Google Scholar] [CrossRef]
  30. Tang, W.; Jiang, Z.; Wang, B.; Li, Y. Black phosphorus quantum dots: A new-type of water-based high-efficiency lubricant additive. Friction 2021, 9, 1528–1542. [Google Scholar] [CrossRef]
  31. Liu, H.; Lian, P.; Tang, Y.; Zhao, Z.; Mei, Y. The preparation of black phosphorus quantum dots by gas exfoliation with the assistance of liquid N 2. J. Nanosci. Nanotechnol. 2020, 20, 6458–6462. [Google Scholar] [CrossRef]
  32. Sun, D.; Kang, S.; Liu, C.; Lu, Q.; Cui, L.; Hu, B. Effect of Zeta Potential and Particle Size on the Stability of SiO2 Nanospheres as Carrier For Ultrasound Imaging Contrast Agents. Int. J. Electrochem. Sci. 2016, 11, 8520–8529. [Google Scholar] [CrossRef]
  33. Wang, P.; Keller, A.A. Natural and engineered nano and colloidal transport: Role of zeta potential in prediction of particle deposition. Langmuir 2009, 25, 6856–6862. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, H.; Liu, Y. Superlubricity Achieved with two-dimensional nano-additives to liquid lubricants. Friction 2020, 8, 1007–1024. [Google Scholar] [CrossRef]
  35. Dietrich, P.M.; Streeck, C.; Glamsch, S.; Ehlert, C.; Lippitz, A.; Nutsch, A.; Kulak, N.; Beckhoff, B.; Unger, W.E.S. Quantification of silane molecules on oxidized silicon: Are There options for a traceable and absolute determination? Anal. Chem. 2015, 87, 10117–10124. [Google Scholar] [CrossRef] [PubMed]
  36. Ge, X.; Li, J.; Zhang, C.; Liu, Y.; Luo, J. Superlubricity And antiwear properties of in situ-formed ionic liquids at ceramic Interfaces induced by tribochemical reactions. Acs Appl. Mater. Interfaces 2019, 11, 6568–6574. [Google Scholar] [CrossRef]
  37. Xu, J.; Kato, K. Formation of tribochemical layer of ceramics sliding in water and its role for low friction. Wear 2000, 245, 61–75. [Google Scholar] [CrossRef]
  38. Nakamura, Y.; Muto, J.; Nagahama, H.; Shimizu, I.; Miura, T.; Arakawa, I. Amorphization of quartz by friction: Implication to silica-gel lubrication of fault surfaces. Geophys. Res. Lett. 2012, 39, 2–7. [Google Scholar] [CrossRef]
  39. Dowson, D.; Higginson, G.R. Elasto-Hydrodynamic Lubrication: International Series On Materials Science and Technology; Elsevier: Amsterdam, The Netherlands, 2014. [Google Scholar]
  40. Li, J.; Zhang, C.; Deng, M.; Luo, J. Investigation of the difference in liquid superlubricity between water- and oil-based lubricants. Rsc Adv. 2015, 5, 63827–63833. [Google Scholar] [CrossRef]
  41. Shi, Y.; Minami, I.; Grahn, M.; Björling, M.; Larsson, R. Boundary and elastohydrodynamic lubrication studies of glycerol aqueous solutions as green lubricants. Tribol. Int. 2014, 69, 39–45. [Google Scholar] [CrossRef]
  42. Long, Y.; Bouchet, M.I.D.B.; Lubrecht, T.; Onodera, T.; Martin, J.M. Superlubricity of glycerol by self-sustained chemical polishing. Sci. Rep. 2019, 9, 6286. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic diagram of the synthesis process of the Baq (b) TEM of BPQDs and (c) HRTEM image of the BPQDs, (d) Statistical analysis of the size of 120 BPQDs based on the TME image, (e) XRD spectrum of BP power, (f) Raman spectrum of BPQDs, (g) AFM image of BPQDs, (h) Height profiles along the white line in (g), (i) Thickness histogram of BPQDs based on the AFM image, (j) Zeta potential of Baq.
Figure 1. (a) Schematic diagram of the synthesis process of the Baq (b) TEM of BPQDs and (c) HRTEM image of the BPQDs, (d) Statistical analysis of the size of 120 BPQDs based on the TME image, (e) XRD spectrum of BP power, (f) Raman spectrum of BPQDs, (g) AFM image of BPQDs, (h) Height profiles along the white line in (g), (i) Thickness histogram of BPQDs based on the AFM image, (j) Zeta potential of Baq.
Lubricants 10 00158 g001
Figure 2. The optical image of the Baq with varied concentration (a) 0 h and (b) after 7 days.
Figure 2. The optical image of the Baq with varied concentration (a) 0 h and (b) after 7 days.
Lubricants 10 00158 g002
Figure 3. (a) The COF curves under the lubrication condition of ultrapure water (UW) and Baq (3 N, 62.8 mm∙s−1); the average COF value of the Baq under different concentrations (b), loads (c) and sliding speeds (d).
Figure 3. (a) The COF curves under the lubrication condition of ultrapure water (UW) and Baq (3 N, 62.8 mm∙s−1); the average COF value of the Baq under different concentrations (b), loads (c) and sliding speeds (d).
Lubricants 10 00158 g003
Figure 4. (a) The COF curves under the lubrication conditions of Gaq and BGaq as lubricant (3 N, 62.8 mm·s−1); (b) The COF curves under the different volumes of BGaq; the COF curve and the average COF value under the lubrication condition of the BGaq with different concentrations (c,d), loads (e,f) and sliding speeds (g,h); (i) Variations of COF of BGaq after one week.
Figure 4. (a) The COF curves under the lubrication conditions of Gaq and BGaq as lubricant (3 N, 62.8 mm·s−1); (b) The COF curves under the different volumes of BGaq; the COF curve and the average COF value under the lubrication condition of the BGaq with different concentrations (c,d), loads (e,f) and sliding speeds (g,h); (i) Variations of COF of BGaq after one week.
Lubricants 10 00158 g004
Figure 5. The optical image of the wear scar on the Si3N4 ball lubricated with (a) ultrapure water and (b) Baq; the morphology of the worn surface on the SiO2 plate under the lubrication condition of ultrapure water (c,e) and BGaq (d,f); the wear rate of SiO2 plate and Si3N4 ball (Φ 10 mm) (g).
Figure 5. The optical image of the wear scar on the Si3N4 ball lubricated with (a) ultrapure water and (b) Baq; the morphology of the worn surface on the SiO2 plate under the lubrication condition of ultrapure water (c,e) and BGaq (d,f); the wear rate of SiO2 plate and Si3N4 ball (Φ 10 mm) (g).
Lubricants 10 00158 g005
Figure 6. The optical microscope image under the lubrication condition of ultrapure water (Note: UW is ultrapure water) (a), Gaq (b) and BGaq (c); the 3D micrographs and two-dimensional profiles of worn surface of SiO2 plate under the lubrication condition of ultrapure water (d), Gaq (e) and BGaq (f); the optical microscope image of the worn track lubricated by ultrapure water (g), Gaq (h) and BGaq (i) using the Φ 4 mm ball; (j) the wear rate of SiO2 plate and Si3N4 ball (Φ 4 mm) after the tribological tests.
Figure 6. The optical microscope image under the lubrication condition of ultrapure water (Note: UW is ultrapure water) (a), Gaq (b) and BGaq (c); the 3D micrographs and two-dimensional profiles of worn surface of SiO2 plate under the lubrication condition of ultrapure water (d), Gaq (e) and BGaq (f); the optical microscope image of the worn track lubricated by ultrapure water (g), Gaq (h) and BGaq (i) using the Φ 4 mm ball; (j) the wear rate of SiO2 plate and Si3N4 ball (Φ 4 mm) after the tribological tests.
Lubricants 10 00158 g006
Figure 7. The wear rates of BGaq under different conditions; (a) concentration, (b) loads, (c) sliding speeds. (Note: UW is ultrapure water).
Figure 7. The wear rates of BGaq under different conditions; (a) concentration, (b) loads, (c) sliding speeds. (Note: UW is ultrapure water).
Lubricants 10 00158 g007
Figure 8. XPS spectra of the worn surface on the SiO2 plate under the lubrication condition of Baq (ae) and BGaq (fj).
Figure 8. XPS spectra of the worn surface on the SiO2 plate under the lubrication condition of Baq (ae) and BGaq (fj).
Lubricants 10 00158 g008
Figure 9. Schematic illustration of the lubrication model with BPQD additives in glycerol aqueous solution. (a,b) the Macroscopic schematic diagram, (c) formation of silica gel layer and colloidal silicon oxide film, (d) the formation of hydrogen-bonded network and the synergistic effect between BPQDs and glycerol.
Figure 9. Schematic illustration of the lubrication model with BPQD additives in glycerol aqueous solution. (a,b) the Macroscopic schematic diagram, (c) formation of silica gel layer and colloidal silicon oxide film, (d) the formation of hydrogen-bonded network and the synergistic effect between BPQDs and glycerol.
Lubricants 10 00158 g009
Table 1. The average COFs and wear volumes under the lubrication condition of ultrapure water and Baq.
Table 1. The average COFs and wear volumes under the lubrication condition of ultrapure water and Baq.
ConditionAverage COFsWear Volume /um3
Si3N4 BallSiO2 Plate
Ultrapure Water3 N, 62.8 mm·s−1/4.1054 × 1065.3800 × 107
Baq with different concentration0.01 wt%0.00711.0514 × 1062.2980 × 107
0.015 wt%0.00245.2355 × 1051.1660 × 107
0.002 wt%0.00347.7463 × 1051.6740 × 107
0.004 wt%0.00625.8154 × 1051.4960 × 107
Baq with different load2.0 N0.0073.8851 × 1051.0490 × 107
2.5 N0.00384.5477 × 1051.3730 × 107
3.5 N0.00491.2518 × 1061.5720 × 107
Baq with different sliding speed31.4 mm∙s10.00727.5252 × 1053.0560 × 107
94.2 mm∙s−10.00381.4740 × 1061.3180 × 107
125.6 mm∙s−10.00271.1428 × 1061.4230 × 107
Table 2. The average COFs and wear volumes under the lubrication condition of Gaq and BGaq.
Table 2. The average COFs and wear volumes under the lubrication condition of Gaq and BGaq.
ConditionAverage COFsWear Volume /um3
Si3N4 BallSiO2 Plate
Gaq3N, 62.8 mm·s−1/1.5298 × 1056.40 × 106
BGaq with different concentration0.01 wt%0.00761.2580 × 1054.96 × 106
0.015 wt%0.00651.1647 × 1054.28 × 106
0.02 wt%0.00771.5081 × 1053.82 × 106
0.04 wt%0.0082.0608 × 1052.14 × 106
BGaq with different load0.5 N0.0041.0700 × 1054.00 × 106
1.0 N0.00653.7734 × 1045.80 × 106
1.5 N0.00744.6759 × 1045.37 × 106
2.0 N0.00837.1887 × 1044.31 × 106
2.5 N0.00631.0485 × 1054.24 × 106
3.5 N0.00881.6510 × 1054.52 × 106
BGaq with different sliding speed94.2 mm∙s−10.00551.4154 × 1055.99 × 106
125.6 mm∙s−10.00831.5761 × 1051.52 × 107
157.0 mm∙s−10.00952.3144 × 1053.93 × 106
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gong, P.; Qu, Y.; Wang, W.; Lv, F.; Jin, J. Macroscale Superlubricity of Black Phosphorus Quantum Dots. Lubricants 2022, 10, 158. https://doi.org/10.3390/lubricants10070158

AMA Style

Gong P, Qu Y, Wang W, Lv F, Jin J. Macroscale Superlubricity of Black Phosphorus Quantum Dots. Lubricants. 2022; 10(7):158. https://doi.org/10.3390/lubricants10070158

Chicago/Turabian Style

Gong, Penghui, Yishen Qu, Wei Wang, Fanfan Lv, and Jie Jin. 2022. "Macroscale Superlubricity of Black Phosphorus Quantum Dots" Lubricants 10, no. 7: 158. https://doi.org/10.3390/lubricants10070158

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop