Tayssir Hamieh*abc
aFaculty of Science and Engineering, Maastricht University, P. O. Box 616, 6200 MD Maastricht, The Netherlands. E-mail: t.hamieh@maastrichtuniversity.nl
bInstitut de Science des Matériaux de Mulhouse, Université de Haute-Alsace, CNRS, IS2M UMR 7361, F-68100 Mulhouse, France
cLaboratory of Materials, Catalysis, Environment and Analytical Methods (MCEMA), Faculty of Sciences, Lebanese University, P. O. Box 6573/14, Beirut, Lebanon
First published on 6th August 2025
Understanding the role of temperature in modulating surface adhesion properties of graphene and its derivatives is essential for their effective integration in nano- and optoelectronic devices. In this study, the temperature-dependent dispersive and polar components of work of adhesion was systematically investigated across graphene (G), reduced graphene oxide (rGO), and graphene oxide (GO), using inverse gas chromatography (IGC) and selected polar/nonpolar solvent interactions. Our results reveal a consistent hierarchy in adhesion energies (G > rGO > GO) and show that elevated temperatures significantly influence interfacial interactions by modifying surface energy components. Furthermore, the solvent-specific trends suggest a strong interplay between molecular polarity and surface functionalization. This study not only provides thermodynamic insights into graphene-based adhesion but also contributes to rational interface engineering in 2D materials under thermal fluctuation.
However, despite their intrinsic chemical versatility, the temperature-dependent nature of their surface adhesion work remains insufficiently explored. Indeed, the work of adhesion (Wa), an important thermodynamic variable governing the interfacial interactions of solvents on solid surfaces, quantifies the energy required to separate two phases in contact.23,24 The molecular mechanisms and thermodynamics for determining solvent adhesion to solid material surfaces are essential for controlling wettability, solvent adsorption, and interfacial interactions.25,26 The concept of thermodynamic “work of adhesion”, Wa, was first introduced by Harkins27 who used the various surface tensions of liquids adhering to a solid. The combination of Harkins's concept,27 Young–Dupré equation,28,29 Fowkes,30 Owens–Wendt,31 and van Oss et al.,32,33 led to the determination of the work of adhesion taking in consideration the dispersive, polar, and acid–base interactions.30–33 Several experimental techniques, such as dynamic contact angle technique, isothermal titration calorimetry (ITC), and inverse gas chromatography (IGC), were used for evaluating the adhesion works and contributing to the quantification of intermolecular interactions.24,34
Adhesion plays a critical role in applications such as flexible electronics, membrane-based sensing, nanofluidics, and heterogeneous integration. Graphene's adhesion to various substrates has been attributed primarily to van der Waals forces, whereas GO and rGO exhibit complex interactions involving hydrogen bonding and dipolar forces due to surface oxygen functionalities.35 The interplay of these forces evolves with temperature, affecting not only contact energy but also interfacial mobility, roughness adaptation, and phonon coupling.36
Recent studies demonstrate that the adhesion of 2D materials can be “gas-like,” characterized by a temperature-sensitive entropic contribution that decreases linearly with increased thermal vibration.37 Moreover, functional group density and orientation on GO/rGO surfaces are thermally active and can reconfigure interfacial potential landscapes.38
Temperature-controlled surface energy variation is not merely of academic interest—it directly affects material transfer yield, contact resistance, and durability in heterostructures.39 While contact mechanics and adhesion of pristine graphene have been modeled thermodynamically, empirical and solvent-specific trends for functionalized graphene variants are scarce.40
Some surface properties of graphene oxide (GO) and graphene (rGO) were determined by Dai et al.41 using the inverse gas chromatography (IGC) technique at infinite dilution. However, these results cannot be considered as accurate due to the wrong hypothesis admitted by Dai et al.41 supposing the surface area and the London dispersive surface energy of n-alkanes as constant independent from the temperature. Indeed, our previous works42–46 proved a strong effect of the temperature on the surface area of organic molecules and gave the variations of the surface area and the London dispersive surface energy of the different solvents as a function of temperature. The same previous errors were committed by Lee et al.47 when determining the London dispersive and polar surface of graphene materials using the classic chromatographic methods by neglecting the temperature effect on the surface area and surface tension of solvents. The IGC technique at infinite dilution was used during the last fifty years for the determination of surface properties of solid materials such as oxides, polymers, metals, or fibers.48–72
In a previous study,73 the London dispersive and polar surface properties of different graphene and carbon materials were determined as a function of temperature by using our new chromatographic approach based on the Hamieh thermal model.42–46 On the other hand, some specific surface chemistries and morphologies of graphenes that directly influence their interfacial adhesion properties were studied in the literature. Ferrari et al.74 and Bunch et al.75 have confirmed the monolayer and monocrystalline structure of graphene (G) with minimal defects and high in-plane ordering. Whereas Dreyer et al.76 and Lerf et al.77 revealed an oxygen-to-carbon (O/C) atomic ratio of 0.45, indicating a moderate-to-high oxidation level and showing the presence of hydroxyl, epoxy, and carboxyl functional groups. The reduced graphene oxide (rGO) was investigated by Eda et al.78 and Stankovich et al.,79 showing that the O/C ratio decreased to approximately 0.12, indicating partial removal of oxygenated groups and a sharp reduction in OH and COOH features, though some residual carbonyl groups remained.
The temperature-dependent surface adhesion behavior of graphene materials plays a pivotal role in the design and optimization of microfluidic and nanofluidic systems, where precise control of fluid–surface interactions is critical. In particular, tunable adhesion at the solid–liquid or solid–vapor interface influences droplet mobility, flow resistance, and interfacial slip, enabling applications in thermal regulation, self-cleaning surfaces, and active fluid transport. Graphene coatings on substrates have been shown to enhance dropwise condensation and heat transfer by up to fourfold relative to conventional coatings, while offering ultrathin, thermally stable surfaces with minimal added resistance.79 Furthermore, bioinspired graphene–PDMS patterned surfaces have demonstrated reversible wettability switching between hydrophobic and hydrophilic states between 0 °C and 200 °C, enabling directed droplet transport and fog harvesting in engineered fluidic platforms.80 At the nanoscale, graphene's adhesion mechanics—dominated by van der Waals interactions and membrane flexibility—are critical to interfacial slip and droplet behavior in confined fluidic systems.81 Recent studies further show that graphene's wettability can be dynamically altered by external stimuli such as temperature or plasma exposure, resulting in reversible modulation of surface energy and adhesion properties.82 Finally, graphene-based open microfluidic channels engineered via laser patterning exploit these tunable surface forces to passively transport fluids without pumps, illustrating direct applications in lab-on-chip device technology.83
This study aims to perform a comprehensive, solvent-resolved thermodynamic analysis of adhesion behaviors of graphene, rGO, and GO by varying the temperature. New theoretical models that have been validated by several recent works based on the Hamieh thermal model41–45 and the new separation method using the London dispersive and polar free energy of adsorption70,71 were applied. By combining the new chromatographic models with quantitative work of adhesion calculations, a clearer picture of how thermal effects alter interfacial physics in 2D materials is given.
Vn = jDc(tR − t0) | (1) |
![]() | (2) |
![]() | (3) |
The determination of the retention volume leads to the variations of the standard free energy ΔG0a of adsorption of organic solvents on the different solid surfaces as a function of temperature using the fundamental equation of inverse gas chromatography:
−ΔG0a(T) = RT![]() ![]() | (4) |
ΔG0a is equal to the London dispersive free energy ΔGda and the polar interaction energy ΔGpa of adsorption of solvents on solid surfaces:
ΔG0a = ΔGda + ΔGpa | (5) |
The two previous contributions of the free energy of adsorption were separately determined in previous works using the London dispersion interaction equation.
![]() | (6) |
In the case of non-polar solvents such as n-alkanes adsorbed on solid surfaces, eqn (6) can be written as follows:
![]() | (7) |
The variations of a(T) and γdl(T) of n-alkanes adsorbed on graphenes as a function of temperature led to accurate values of the London dispersive surface energy of different graphenes. Whereas the dispersive free energy ΔG0a(T) and the polar free energy ΔGpa(T) of different organic solvents adsorbed on graphene, GO, and rGO, were separated using our new methodology based on the London interaction energy equation. By applying the Van Oss et al.‘s method,33 the Lewis acid γs+, and base γs− surface energies of graphenes were obtained by choosing two polar solvents such as ethyl acetate and dichloromethane and using eqn (8) given −ΔGpa(T):
![]() | (8) |
The polar surface energy of the solvent is obtained from eqn (8), while the polar surface energy γps(T) of graphenes was determined by the following equation:
![]() | (9) |
γs(T) = γds(T) + γps(T) | (10) |
![]() | (11) |
However, Dupré17 using Young's equation28 gave the work of adhesion by eqn (12):
Wa = γl(1 + cos![]() | (12) |
While Fowkes30 gave a new expression relative to the dispersive work of adhesion Wda by taking the geometric mean of the dispersive components of the solid and liquid:
![]() | (13) |
By considering the polar interaction, Fowkes30 gave the new expression of the work of adhesion:
![]() | (14) |
Finally, the work of adhesion can be written as follows:
![]() | (15) |
Knowing the different surface energy components of graphene materials previously determined,73 the dispersive Wda(T), polar Wpa(T)and total Wa(T) works of adhesion were determined using eqn (8) and (13)–(15).
T (K) | Graphene | Graphene oxide | Reduced graphene oxide |
---|---|---|---|
313.15 | 279.19 | 118.24 | 150.97 |
323.15 | 261.11 | 109.29 | 147.26 |
333.15 | 243.30 | 100.60 | 143.33 |
343.15 | 225.78 | 92.15 | 139.24 |
353.15 | 208.54 | 83.95 | 134.93 |
363.15 | 191.63 | 76.03 | 130.33 |
373.15 | 175.03 | 68.36 | 125.48 |
Table 1 showed that γds of graphenes linearly decreases as the temperature increased. The lowest γds was obtained with the graphene oxide, whereas the highest value was shown with the graphene. It seems that the oxidation of graphene decreases the values of γds to about 40% of its initial value, while the γds variations of reduced graphene oxide decreased until 54% at 313.15 K to 71.7% at 373.15 K. The various graphenes can be classified in increasing order of the London dispersive surface energy as follows:
Graphene oxide < Reduced graphene oxide < Graphene |
The increasing order of London dispersive surface energy can be explained based on their chemical structure, electronic properties, and surface morphology, which directly affect the van der Waals (London dispersion) interactions. London dispersion forces are a type of van der Waals force, arising from instantaneous dipole-induced dipole interactions. These forces increase with surface electron density and polarizability. The dispersive component of surface energy is closely linked to π–π stacking ability, sp2 character, and electronic delocalization. The Graphene oxide (GO) structure is heavily oxidized with abundant oxygen-containing hydroxyl, epoxide, carbonyl, and carboxyl groups. This 2D material exhibits high disruption of sp2 domains and many sp3 hybridized carbons due to its functionalization with low surface π-electron delocalization. This gives weak London dispersion forces due to reduced polarizability and disrupted π-system leading to the lowest London dispersive surface energy. Whereas the reduced graphene oxide (rGO) is partially restored sp2 network, but with defects and residual oxygen groups with moderate π-delocalization and less polarity than GO. However, the graphene (G) structure presents 2D sheet of sp2-bonded carbon atoms with pure sp2 hybridization, highly delocalized π-electron cloud, non-polar hydrophobic material, and strongest London dispersion forces due to high polarizability and dense, continuous π-electron system, and consequently, the graphene exhibits the highest London dispersive surface energy γds.
Temperature T (K) | 313.15 | 323.15 | 333.15 | 343.15 | 353.15 | 363.15 | 373.15 |
---|---|---|---|---|---|---|---|
Base surface energy γs−(T) of graphenes (mJ m−2) | |||||||
Graphene | 73.55 | 68.61 | 63.95 | 59.53 | 55.36 | 51.42 | 47.70 |
Reduced graphene oxide | 70.30 | 63.58 | 57.32 | 51.50 | 46.10 | 41.09 | 36.46 |
Graphene oxide | 20.29 | 19.69 | 19.12 | 18.58 | 18.05 | 17.53 | 17.03 |
![]() |
|||||||
Acid surface energy γs+(T) of graphenes (mJ m−2) | |||||||
Graphene | 167.08 | 162.07 | 157.32 | 152.79 | 148.51 | 144.41 | 140.52 |
Reduced graphene oxide | 103.34 | 98.87 | 94.69 | 90.70 | 86.96 | 83.44 | 80.07 |
Graphene oxide | 5.32 | 7.25 | 9.41 | 11.79 | 14.29 | 16.93 | 19.68 |
![]() |
|||||||
Polar surface energy γps(T) of graphenes (mJ m−2) | |||||||
Graphene | 221.71 | 210.91 | 200.61 | 190.75 | 181.34 | 172.35 | 163.73 |
Reduced graphene oxide | 170.47 | 158.57 | 147.35 | 136.69 | 126.62 | 117.11 | 108.06 |
Graphene oxide | 20.78 | 23.91 | 26.84 | 29.60 | 32.12 | 34.45 | 36.61 |
![]() |
|||||||
Total surface energy γs(T) of graphenes (mJ m−2) | |||||||
Graphene | 500.90 | 472.02 | 443.90 | 416.53 | 389.88 | 363.98 | 338.77 |
Reduced graphene oxide | 321.4 | 305.8 | 290.7 | 275.9 | 261.6 | 247.4 | 233.5 |
Graphene oxide | 139.0 | 133.2 | 127.4 | 121.7 | 116.1 | 110.5 | 105.0 |
Table 2 showed a highest acid–base surface energy character of graphene, followed by rGO, and GO. This is due to the polarizable electron cloud of graphene, even though it's chemically inert—it behaves like a strong polarizable surface, not because of functional groups, but due to delocalized π-electrons. Whereas rGO bridges the gap, exhibiting intermediate polarity and tunable surface chemistry depending on reduction extent. While the GO's polar behavior is governed by surface functionalities, which become more active with temperature, especially in acid–base interactions.
The various solvents adsorbed on graphene were classified in increasing order of polar surface energy as follows:
Dichloromethane < THF < Ethyl acetate < Acetone < Diethyl ether < Acetonitrile |
The above order proved that the solvents with increasing dipole moment and hydrogen-bonding capacity showed stronger polar interaction. The graphene interacts weakly via polar forces, so more strongly polar solvents (e.g., acetonitrile) stand out in polar interactions.
In the case of reduced graphene oxide (rGO), the following increasing order of γpl(T) was obtained:
Dichloromethane < Acetonitrile < THF < Acetone < Ethyl acetate < Diethyl ether |
It was observed that acetonitrile moves lower in polarity rank compared to graphene. Indeed, rGO has a complex surface with both polar and nonpolar domains. Solvents with dual character (e.g., ethers) can form more stable interactions due to both dipolar alignment and weak H-bonding with residual oxygen groups. However, one obtained the following order of solvents adsorbed on graphene oxide (GO):
Dichloromethane < THF < Diethyl ether < Acetone < Ethyl acetate < Acetonitrile |
This order showed a strongest polar interaction with acetonitrile, a highly polar aprotic solvent. In fact, GO has abundant polar groups (–OH, –COOH, epoxides) that interact via dipole–dipole and hydrogen bonding, favoring polar solvents.
![]() | ||
Fig. 1 Variations of the dispersive work of adhesion Wda(T) of solvents on graphenes versus the temperature. Graphene (a), reduced graphene oxide (b), and graphene oxide (c). |
Fig. 1 and Table S2 clearly showed the same following increasing order of the work of adhesion of solvents on the different graphenes:
Diethyl ether < n-Hexane < n-Heptane < Acetone < Acetonitrile < Ethyl acetate < n-octane < n-nonane < THF < CH2Cl2 |
Indeed, diethyl ether is characterized by its low polarity and lowest dispersive surface energy, and then very poor adhesion for all graphene materials. Now, the n-alkanes from n-hexane to n-nonane are purely dispersive, increasing the dispersive work of adhesion Wda(T) with chain length due to greater van der Waals contact area. Acetone and acetonitrile are polar aprotic solvents with relatively low dispersive surface energy. Whereas, ethyl acetate has both dispersive and moderately polar surface energy, and provides better adhesion. The higher dispersive surface energy of tetrahydrofuran (THF) and better matching to carbon surfaces result in higher Wda. Though dichloromethane is moderately polar, it has a relatively high dispersive surface energy, explaining its high adhesion with graphene surfaces.
Even though values of Wda(T) differ for graphene surfaces with the highest value for graphene followed by reduced graphene oxide, and graphene oxide, the solvent ranking remains constant because:
1. The dispersive component Wda of adhesion work which mainly depends on γds of graphene which is constant for each graphene at a fixed temperature, and on γdl of solvent which varies by solvent. The solvent then dominates the trend.
2. The ranking reflects increasing γdl and/or increasing molecular size/interaction area of the solvents.
3. Polar solvents (e.g. acetone, acetonitrile) appear earlier in the ranking because their γdl is relatively low, despite having dipoles.
The determination of polar work of adhesion Wpa(T) of solvents on graphenes versus the temperature was given in Table S3. The variations of Wpa(T) were drawn in Fig. 2. The results in Table S3 and Fig. 2 allowed giving the different solvents in increasing order of Wpa(T) on graphenes:
![]() | ||
Fig. 2 Variations of the polar work of adhesion Wpa(T) of solvents on graphenes versus the temperature. Graphene (a), reduced graphene oxide (b), and graphene oxide (c). |
For graphene:
CH2Cl2 < THF < Ethyl acetate < Acetone < Diethyl ether < Acetonitrile |
For reduced graphene oxide:
CH2Cl2 < Acetonitrile < THF < Acetone < Ethyl acetate < Diethyl ether |
For graphene oxide:
CH2Cl2 < THF < Diethyl ether < Acetone < Ethyl acetate < Acetonitrile |
The above rankings reflect how well the polarity of graphenes matches the polarity of the solvent. The graphene is considered as non-polar surface. This justifies the very low its polar surface energy γps. Even highly polar solvents (like acetonitrile) don't interact much the polar work of adhesion Wpa(T) remains low, but small differences emerge from solvent polarity. Acetonitrile has the highest polar surface energy γpl and therefore the highest polar adhesion with G. Whereas the reduced graphene oxide (rGO) is characterized by intermediate γps and partial removal of oxygen groups. The polar work of adhesion Wpa(T) now depends more subtly on both γps and γpl. Even if acetonitrile is a polar solvent, however, it has low hydrogen-bonding ability leading to low polar adhesion with rGO. while Diethyl ether, though weakly polar, can interact through lone pairs and π-donation on oxygen and give relatively high polar work of adhesion Wpa. The graphene oxide exhibiting the highest γps due to hydroxyl, epoxide, carboxyl groups, interacts strongly with polar solvents, especially hydrogen-bond donors and acceptors. Thus, Wpa increases sharply with solvents like acetonitrile and ethyl acetate, while diethyl ether which is weakly polar (only lone pairs) leads to lower interaction than acetone.
In conclusion, the different molecular interactions (dipole–dipole, H-bonding, lone pair interaction) govern how polar work of adhesion varies across graphene surfaces. The trend inversion between rGO and GO shows the sensitivity of Wpa to subtle changes in surface chemistry. Finally, acetonitrile interacts strongly with GO but poorly with rGO, showing how hydrophilicity and donor–acceptor capacity influence polar interactions.
The total work of adhesion Wa(T) of solvents on graphenes was obtained by summing the London dispersive and polar works of adhesion as a function of temperature. The linear variations of Wa(T) were plotted versus the temperature in Fig. S1. The results clearly showed the highest values of the work of adhesion of polar solvents compared to those of n-alkanes due the highest polar interactions of polar molecules with graphenes. The linear relations of the work of adhesion Wa(T) of different solvents on graphenes as a function of temperature were given in Table 3. It was deduced that Wa(T) can be thermodynamically written as:
Wa(T) = ΔHS − TΔSS | (16) |
Solvents | Wa(T) | ΔSS | ΔHS | TMax. (K) | R2 |
---|---|---|---|---|---|
Graphene | |||||
n-Hexane | Wa(T) = −0.843T + 398.73 | 0.843 | 398.73 | 473.05 | 0.9999 |
n-Heptane | Wa(T) = −0.828T + 401.62 | 0.828 | 401.62 | 484.93 | 0.9999 |
n-Octane | Wa(T) = −0.823T + 405.74 | 0.823 | 405.74 | 493.30 | 0.9999 |
n-Nonane | Wa(T) = −0.822T + 410.53 | 0.822 | 410.53 | 499.49 | 1.0000 |
CH2Cl2 | Wa(T) = −1.016T + 545.64 | 1.016 | 545.64 | 536.84 | 0.9999 |
Diethyl ether | Wa(T) = −1.407T + 668.69 | 1.407 | 668.69 | 475.23 | 0.9997 |
THF | Wa(T) = −1.399T + 683.1 | 1.399 | 683.1 | 488.24 | 0.9998 |
Ethyl acetate | Wa(T) = −1.017T + 552.89 | 1.017 | 552.89 | 543.49 | 1.0000 |
Acetone | Wa(T) = −1.129T + 587.95 | 1.129 | 587.95 | 520.86 | 0.9999 |
Acetonitrile | Wa(T) = −0.984T + 567.1 | 0.984 | 567.1 | 576.15 | 1.0000 |
![]() |
|||||
Reduced graphene oxide | |||||
n-Hexane | Wa(T) = −0.463T + 244.46 | 0.463 | 244.46 | 528.11 | 0.9995 |
n-Heptane | Wa(T) = −0.437T + 241.69 | 0.437 | 241.69 | 553.57 | 0.9996 |
n-Octane | Wa(T) = −0.421T + 241.07 | 0.421 | 241.07 | 572.88 | 0.9996 |
n-Nonane | Wa(T) = −0.411T + 241.68 | 0.411 | 241.68 | 587.89 | 0.9996 |
CH2Cl2 | Wa(T) = −0.578T + 351.51 | 0.578 | 351.51 | 608.68 | 0.9996 |
Diethyl ether | Wa(T) = −0.999T + 504.22 | 0.999 | 504.22 | 504.57 | 0.9999 |
THF | Wa(T) = −0.936T + 483.61 | 0.936 | 483.61 | 516.73 | 1.0000 |
Ethyl acetate | Wa(T) = −0.743T + 426.15 | 0.743 | 426.15 | 573.71 | 1.0000 |
Acetone | Wa(T) = −0.405T + 307.74 | 0.405 | 307.74 | 760.79 | 0.9996 |
Acetonitrile | Wa(T) = −0.269T + 260.83 | 0.269 | 260.83 | 968.55 | 0.9954 |
![]() |
|||||
Graphene oxide | |||||
n-Hexane | Wa(T) = −0.585T + 270.74 | 0.585 | 270.74 | 462.88 | 0.9998 |
n-Heptane | Wa(T) = −0.579T + 273.76 | 0.579 | 273.76 | 472.90 | 0.9998 |
n-Octane | Wa(T) = −0.578T + 277.28 | 0.578 | 277.28 | 479.81 | 0.9998 |
n-Nonane | Wa(T) = −0.580T + 281.07 | 0.580 | 281.07 | 484.94 | 0.9998 |
CH2Cl2 | Wa(T) = −0.513T + 280.66 | 0.513 | 280.66 | 547.10 | 0.9998 |
Diethyl ether | Wa(T) = −0.748T + 363.28 | 0.748 | 363.28 | 485.54 | 0.9997 |
THF | Wa(T) = −0.803T + 393.34 | 0.803 | 393.34 | 489.90 | 0.9997 |
Ethyl acetate | Wa(T) = −0.588T + 325.73 | 0.588 | 325.73 | 553.87 | 0.9999 |
Acetone | Wa(T) = −0.526T + 302.00 | 0.526 | 302.00 | 574.47 | 0.9998 |
Acetonitrile | Wa(T) = −0.558T + 333.07 | 0.558 | 333.07 | 597.33 | 0.9998 |
The deduced values of ΔHS and ΔSS of the different organic solvents on graphenes were given in Table 3, including the corresponding values of the maximum temperature TMax defined as follows:
TMax = ΔHS/ΔSS | (17) |
Table 3 showed that the values of ΔHS of the solvents adsorbed on graphenes linearly depends on the surface entropy of adhesion work ΔSS. Equations ΔHS = f(ΔSS) of the different graphenes were given in Table 4. A new surface temperature TS of material was defined.
2D materials | Equations ΔHS = f(ΔSS) | TS |
---|---|---|
Graphene | ΔHS = 471.03 ΔSS + 38.517 | 471.03 |
Reduced graphene oxide | ΔHS = 406.46 ΔSS + 100.21 | 406.46 |
Graphene oxide | ΔHS = 370.87 ΔSS + 85.464 | 370.87 |
The general equations ΔHS = f(ΔSS) of graphenes deduced from Table 4 can be written as follows:
ΔHS = TSΔSS + Q | (18) |
To compare between the work of adhesion Wa(T) of a solvent on graphene materials, the variations of Wa(T) of each solvent on the different graphenes as a function of temperature were plotted in Fig. S2. It was showed that the adhesion of all solvents on graphenes satisfied the same classification. The increasing order of the work of adhesion on graphenes is given as follows:
Graphene oxide < Reduced graphene oxide < Graphene |
The observed trend in the work of adhesion of various solvents on graphene-based materials reflects the fundamental changes in surface chemistry and electronic structure across these materials, directly influencing intermolecular interactions with adsorbing solvent molecules. Indeed, graphene oxide (GO) contains a high density of oxygenated functional groups (hydroxyl, epoxy, carbonyl, carboxyl), which increase polarity and surface heterogeneity but decrease π-electron density and planar conjugation. These polar sites may disrupt van der Waals (particularly London dispersion) interactions, which are dominant in nonpolar solvent adhesion. Whereas reduced graphene oxide (rGO) is partially restored to a graphitic structure with fewer oxygen groups and a partially recovered π-conjugated system, increasing π–π interactions and improving London dispersion forces, especially with nonpolar or slightly polar solvents. While graphene (G) has a fully conjugated, delocalized π-electron system, enabling maximum van der Waals interactions, especially London dispersion with alkane solvents, and π–π or dipole–π interactions with polar solvents such as THF, acetone, and acetonitrile. These results can be interpreted in term of the work of adhesion. Knowing that the work of adhesion Wa(T) given by eqn (6) between a solvent and a solid surface is generally a function of the interfacial free energy. Graphene has the highest surface energy due to strong cohesive π–π forces, while GO has the lowest because of disrupted π-systems and hydrophilic heterogeneity, reducing solvent–surface affinity. Thus, higher surface energy in graphene enhances interaction strength and thus adhesion. This is consistent with the classification of solvents. Despite the chemical diversity of the solvents (aliphatic, chlorinated, etheric, carbonyl-containing, nitriles), the same adhesion trend across all of them suggests that the surface nature of the graphene material dominates over solvent-specific properties and the main interaction mechanism is non-specific, i.e., driven largely by van der Waals (especially London dispersion) and π–π stacking, rather than hydrogen bonding or strong dipole–dipole interactions. This explains why nonpolar solvents (e.g., n-alkanes) and polar aprotic solvents (e.g., acetone, acetonitrile) follow the same trend. Therefore, the increasing adhesion from GO to graphene is governed by increasing surface energy, enhanced π–electron delocalization, and stronger London dispersion and π–π interactions, which collectively improve the solvent–surface affinity regardless of the solvent's polarity.
Graphene > rGO > GO |
• The high dispersive component of pristine graphene enables stronger interaction with solvents dominated by van der Waals forces.
• rGO, being partially restored in its conjugated π-system, shows moderate affinity due to a balance between polar and dispersive interactions.
• GO, rich in oxygen-containing groups, is predominantly polar but exhibits limited dispersive interaction.
The results also highlight that solvent-specific interaction rankings vary depending on the surface polarity and π-electron density of the carbon material. This work underlines the necessity of tailoring solvent-material pairs for applications such as film transfer, nanofluidics, and composite fabrication.
This supplementary information gives the values of polar surface energy, dispersive work of adhesion, and polar work of adhesion of various solvents adsorbed on graphene G, graphene oxide GO, and reduced graphene oxide rGO as a function of temperature. It includes the comparison between graphenes giving the variations of the total work of adhesion of solvents on graphenes versus the temperature. n-Hexane, n-heptane, n-octane, n-nonane, dichloromethane, diethyl ether, THF, ethyl acetate, acetone, and acetonitrile. See DOI: https://doi.org/10.1039/d5ra03892h.
This journal is © The Royal Society of Chemistry 2025 |