Nanotribology

Last updated

Nanotribology is the branch of tribology that studies friction, wear, adhesion and lubrication phenomena at the nanoscale, where atomic interactions and quantum effects are not negligible. The aim of this discipline is characterizing and modifying surfaces for both scientific and technological purposes.

Contents

Nanotribological research has historically involved both direct and indirect methodologies. [1] [2] [3] Microscopy techniques, including Scanning Tunneling Microscope (STM), Atomic-Force Microscope (AFM) and Surface Forces Apparatus, (SFA) have been used to analyze surfaces with extremely high resolution, while indirect methods such as computational methods [4] and Quartz crystal microbalance (QCM) have also been extensively employed. [5] [6]

Changing the topology of surfaces at the nanoscale, friction can be either reduced or enhanced more intensively than macroscopic lubrication and adhesion; in this way, superlubrication and superadhesion can be achieved. In micro- and nano-mechanical devices problems of friction and wear, that are critical due to the extremely high surface volume ratio, can be solved covering moving parts with super lubricant coatings. On the other hand, where adhesion is an issue, nanotribological techniques offer a possibility to overcome such difficulties.

History

Friction and wear have been technological issues since ancient periods. On the one hand, the scientific approach of the last centuries towards the comprehension of the underlying mechanisms was focused on macroscopic aspects of tribology. On the other hand, in nanotribology, the systems studied are composed of nanometric structures, where volume forces (such as those related to mass and gravity) can often be considered negligible compared to surface forces. Scientific equipment to study such systems have been developed only in the second half of the 20th century. In 1969 the very first method to study the behavior of a molecularly thin liquid film sandwiched between two smooth surfaces through the SFA was developed. [7] From this starting point, in 1980s researchers would employ other techniques to investigate solid state surfaces at the atomic scale.

Direct observation of friction and wear at the nanoscale started with the first Scanning Tunneling Microscope (STM), which can obtain three-dimensional images of surfaces with atomic resolution; this instrument was developed by Gerd Binnig and Henrich Rohrer in 1981. [8] STM can study only conductive materials, but in 1985 with the invention of the Atomic Force Microscope (AFM) by Binning and his colleagues, also non conductive surfaces can be observed. [9] Afterwards, AFMs were modified to obtain data on normal and frictional forces: these modified microscopes are called Friction Force Microscopes (FFM) or Lateral Force Microscopes (LFM). The term "Nanotribology" was first used in the title of a 1990 publication [10] and in a 1991 publication . [11] in a title of a major review paper published in Nature in 1995 [6] and in a title of a major Nanotrobology Handbook in 1995. [1]

From the beginning of the 21st century, computer-based atomic simulation methods have been employed to study the behaviour of single asperities, even those composed by few atoms. Thanks to these techniques, the nature of bonds and interactions in materials can be understood with a high spatial and time resolution.

Surface analysis

Surface forces apparatus

The SFA (Surface Forces Apparatus) is an instrument used for measuring physical forces between surfaces, such as adhesion and capillary forces in liquids and vapors, and van der Waals interactions. [12] Since 1969, the year in which the first apparatus of this kind was described, numerous versions of this tool have been developed.

SFA 2000, which has fewer components and is easier to use and clean than previous versions of the apparatus, is one of the currently most advanced equipment utilized for nanotribological purposes on thin films, polymers, nanoparticles and polysaccharides. SFA 2000 has one single cantilever which is able to generate mechanically coarse and electrically fine movements in seven orders of magnitude, respectively with coils and with piezoelectric materials. The extra-fine control enables the user to have a positional accuracy lesser than 1 Å. The sample is trapped by two molecularly smooth surfaces of mica in which it perfectly adheres epitaxially. [12]

Normal forces can be measured by a simple relation:

where is the applied displacement by using one of the control methods mentioned before, is the spring constant and is the actual deformation of the sample measured by MBI. Moreover, if then there is a mechanical instability and therefore the lower surface will jump to a more stable region of the upper surface. And so, the adhesion force is measured with the following formula:

.

Using the DMT model, the interaction energy per unit area can be calculated:

where is the curvature radius and is the force between cylyndrically curved surfaces. [12] [13]

Scanning probe microscopy

SPM techniques such as AFM and STM are widely used in nanotribology studies. [14] [15] [2] The Scanning Tunneling Microscope is used mostly for morphological topological investigation of a clean conductive sample, because it is able to give an image of its surface with atomic resolution.

The Atomic Force Microscope is a powerful tool in order to study tribology at a fundamental level. It provides an ultra-fine surface-tip contact with a high refined control over motion and atomic-level precision of measure. The microscope consists, basically, in a high flexible cantilever with a sharp tip, which is the part in contact with the sample and therefore the crossing section must be ideally atomic-size, but actually nanometric (radius of the section varies from 10 to 100 nm). In nanotribology AFM is commonly used for measuring normal and friction forces with a resolution of pico-Newtons. [16]

The tip is brought close to the sample's surface, consequently forces between the last atoms of the tip and the sample's deflect the cantilever proportionally to the intensity of this interactions. Normal forces bend the cantilever vertically up or down of the equilibrium position, depending on the sign of the force. The normal force can be calculated by means of the following equation:

where is the spring constant of the cantilever, is the output of the photodetector, which is an electric signal, directly with the displacement of the cantilever and is the optical-lever sensitivity of the AFM. [17] [18]

On the other hand, lateral forces can be measured with the FFM, which is fundamentally very similar to the AFM. The main difference resides in the tip motion, that slides perpendicularly to its axis. These lateral forces, i.e. friction forces in this case, result in twisting the cantilever, which is controlled to ensure that only the tip touches the surface and not other parts of the probe. At every step the twist is measured and related with the frictional force with this formula:

where is the output voltage, is the torsional constant of the cantilever, is the height of the tip plus the cantilever thickness and is the lateral deflection sensitivity. [17]

Since the tip is part of a compliant apparatus, the cantilever, the load can be specified and so the measurement is made in load-control mode; but in this way the cantilever has snap-in and snap-out instabilities and so in some regions measurements cannot be completed stably. These instabilities can be avoided with displacement-controlled techniques, one of this is the interfacial force microscopy. [13] [19] [20]

The tap can be at contact with the sample in the whole measurement process, and this is called contact mode (or static mode), otherwise it can be oscillated and this is called tapping mode (or dynamic mode). Contact mode is commonly applied on hard sample, on which the tip cannot leave any sign of wear, such as scars and debris. For softer materials tapping mode is used to minimize the effects of friction. In this case the tip is vibrated by a piezo and taps the surface at the resonant frequency of the cantilever, i.e. 70-400 kHz, and with an amplitude of 20-100 nm, high enough to allow the tip to not get stuck to the sample because of the adhesion force. [21]

The atomic force microscope can be used as a nanoindenter in order to measure hardness and Young's modulus of the sample. For this application, the tip is made of diamond and it is pressed against the surface for about two seconds, then the procedure is repeated with different loads. The hardness is obtained dividing the maximum load by the residual imprint of the indenter, which can be different from the indenter section because of sink-in or pile-up phenomena. [22] The Young's modulus can be calculated using the Oliver and Pharr method, which allows to obtain a relation between the stiffness of the sample, function of the indentation area, and its Young's and Poisson's moduli. [23]

Atomistic simulations

Computational methods are particularly useful in nanotribology for studying various phenomena, such as nanoindentation, friction, wear or lubrication. [13] In an atomistic simulation, every single atom's motion and trajectory can be tracked with a very high precision and so this information can be related to experimental results, in order to interpret them, to confirm a theory or to have access to phenomena, that are invisible to a direct study. Moreover, many experimental difficulties do not exist in an atomistic simulation, such as sample preparation and instrument calibration. Theoretically every surface can be created from a flawless one to the most disordered. As well as in the other fields where atomistic simulations are used, the main limitations of these techniques relies on the lack of accurate interatomic potentials and the limited computing power. For this reason, simulation time is very often small (femtoseconds) and the time step is limited to 1 fs for fundamental simulations up to 5 fs for coarse-grained models. [13]

It has been demonstrated with an atomistic simulation that the attraction force between the tip and sample's surface in a SPM measurement produces a jump-to-contact effect. [24] This phenomenon has a completely different origin from the snap-in that occurs in load-controlled AFM, because this latter is originated from the finite compliance of the cantilever. [13] The origin of the atomic resolution of an AFM was discovered and it has been shown that covalent bonds form between the tip and the sample which dominate van der Waals interactions and they are responsible for a such high resolution. [25] Simulating an AFM scansion in contact mode, It has been found that a vacancy or an adatom can be detected only by an atomically sharp tip. Whether in non-contact mode vacancies and adatoms can be distinguished with the so-called frequency modulation technique with a non-atomically sharp tip. In conclusion only in non-contact mode can be achieved atomic resolution with an AFM. [26]

Properties

Friction

Friction, the force opposing to the relative motion, is usually idealized by means of some empirical laws such as Amonton’s First and Second laws and Coulomb's law. At the nanoscale, however, such laws may lose their validity. For instance, Amonton's second law states that friction coefficient is independent from the area of contact. Surfaces, in general, have asperities, that reduce the real area of contact and therefore, minimizing such area can minimize friction. [21] [27] [28]

During the scanning process with an AFM or FFM, the tip, sliding on the sample's surface, passes through both low (stable) and high potential energy points, determined, for instance, by atomic positions or, on a larger scale, by surface roughness. [21] Without considering thermal effects, the only force that makes the tip overcome these potential barriers is the spring force given by the support: this causes the stick-slip motion.

At the nanoscale, friction coefficient depends on several conditions. For example, with light loading conditions, tend to be lower than those at the macroscale. With higher loading conditions, such coefficient tends to be similar to the macroscopic one. Temperature and relative motion speed can also affect friction.

Lubricity and superlubricity at the atomic scale

Lubrication is the technique used to reduce friction between two surfaces in mutual contact. Generally, lubricants are fluids introduced between these surfaces in order to reduce friction. [21] [27]

However, in micro- or nano-devices, lubrication is often required and traditional lubricants become too viscous when confined in layers of molecular thickness. A more effective technique is based on thin films, commonly produced by Langmuir–Blodgett deposition, or self-assembled monolayers [29]

Thin films and self-assembled monolayers are also used to increase adhesion phenomena.

Two thin films made of perfluorinated lubricants (PFPE) with different chemical composition were found to have opposite behaviors in humid environment: hydrophobicity increases the adhesive force and decreases lubrication of films with nonpolar end groups; instead, hydrophilicity has the opposite effects with polar end groups.

Superlubricity

Superlubricity is a frictionless tribological state sometimes occurring in nanoscale material junctions”. [30]

At the nanoscale, friction tends to be non isotropic: if two surfaces sliding against each other have incommensurate surface lattice structures, each atom is subject to different amount of force from different directions . Forces, in this situation, can offset each other, resulting in almost zero friction.

The very first proof of this was obtained using a UHV-STM to measure. If lattices are incommensurable, friction was not observed, however, if the surfaces are commensurable, friction force is present. [31] At the atomic level, these tribological properties are directly connected with superlubricity. [32]

An example of this is given by solid lubricants, such as graphite, MoS2 and Ti3SiC2: this can be explained with the low resistance to shear between layers due to the stratified structure of these solids. [33]

Even if at the macroscopic scale friction involves multiple microcontacts with different size and orientation, basing on these experiments one can speculate that a large fraction of contacts will be in superlubric regime. This leads to a great reduction in average friction force, explaining why such solids have a lubricant effect.

Other experiments carried out with the LFM shows that the stick-slip regime is not visible if the applied normal load is negative: the sliding of the tip is smooth and the average friction force seems to be zero. [34]

Other mechanisms of superlubricity may include: [35] (a) Thermodynamic repulsion due to a layer of free or grafted macromolecules between the bodies so that the entropy of the intermediate layer decreases at small distances due to stronger confinement; (b) Electrical repulsion due to external electrical voltage; (c) Repulsion due to electrical double layer; (d) Repulsion due to thermal fluctuations. [36]

Thermolubricity at the atomic scale

With the introduction of AFM and FFM, thermal effects on lubricity at the atomic scale could not be considered negligible any more. [37] Thermal excitation can result in multiple jumps of the tip in the direction of the slide and backward. When the sliding velocity is low, the tip takes a long time to move between low potential energy points and thermal motion can cause it to make a lot of spontaneous forward and reverse jumps: therefore, the required lateral force to make the tip follow the slow support motion is small, so the friction force becomes very low.

For this situation was introduced the term thermolubricity.

Adhesion

Adhesion is the tendency of two surfaces to stay attached together. [21] [27]

The attention in studying adhesion at the micro- and nanoscale increased with the development of AFM: it can be used in nanoindentation experiments, in order to quantify adhesion forces [2] [38] [39]

According to these studies, hardness was found to be constant with film thickness, and it's given by: [40]

where is the indentation's area and is the load applied to the indenter.

Stiffness, defined as , where is the indentation's depth, can be obtained from , the radius of the indenter-contact line.

is the reduced Young's modulus, and are the indenter's Young's modulus and Poisson's ratio and , are the same parameters for the sample.

However, can't always be determined from direct observation; it could be deduced from the value of (depth of indentation), but it's possible only if there is no sink-in or pile-up (perfect Sneddon's surface conditions). [41]

If there is sink in, for example, and the indenter is conical the situation is described below.

Displacement of the tip ( h), elastic displacement of sample surface at the contact line with the indenter ( he), contact depth (hc), contact radius (rc) and cone angle (a) of the indenter are shown. Nanoindentation.jpg
Displacement of the tip ( h), elasticdisplacement of sample surface at the contact line with the indenter ( he), contact depth (hc), contactradius (rc) and cone angle (α) of the indenter are shown.

From the image, we can see that:

and

From Oliver and Pharr's study [38]

where ε depends on the geometry of the indenter; if it's conical, if it's spherical and if it's a flat cylinder.

Oliver and Pharr, therefore, did not consider adhesive force, but only elastic force, so they concluded:

Considering adhesive force [41]

Introducing as the adhesion energy and as the work of adhesion:

obtaining

In conclusion:

The consequences of the additional term of adhesion is visible in the following graph:

Load-displacement curves that shows the effect of adhesion force Cattura di schermata (39).png
Load-displacement curves that shows the effect of adhesion force

During loading, indentation depth is higher when adhesion is not negligible: adhesion forces contributes to the work of indentation; on the other hand, during unloading process, adhesion forces opposes indentation process.

Adhesion is also related to capillary forces acting between two surfaces when in presence of humidity. [42]

Applications of adhesion studies

This phenomenon is very important in thin films, because a mismatch between the film and the surface can cause internal stresses and, consequently interface debonding.

When a normal load is applied with an indenter, the film deforms plastically, until the load reaches a critical value: an interfacial fracture starts to develop. The crack propagates radially, until the film is buckled. [40]

On the other hand, adhesion was also investigated for its biomimetic applications: several creatures including insects, spiders, lizards and geckos have developed a unique climbing ability that are trying to be replicated in synthetic materials .

It was shown that a multi-level hierarchical structure produces adhesion enhancement: a synthetic adhesive replicating gecko feet organization was created using nanofabrication techniques and self-assembly. [43]

Wear

Wear is related to the removal and the deformation of a material caused by the mechanical actions. At the nanoscale, wear is not uniform. The mechanism of wear generally begins on the surface of material. The relative motion of two surfaces can cause indentations obtained by the removal and deformation of surface material. Continued motion can eventually grow in both width and depth these indentations [21] . [27]

At the macro scale wear is measured by quantifying the volume (or mass) of material loss or by measuring the ratio of wear volume per energy dissipated. At the nanoscale, however, measuring such volume can be difficult and therefore, it is possible to use evaluate wear by analyzing modifications in surface topology, generally by means of AFM scanning. [44] [2]

See also

Related Research Articles

<span class="mw-page-title-main">Friction</span> Force resisting sliding motion

Friction is the force resisting the relative motion of solid surfaces, fluid layers, and material elements sliding against each other. Types of friction include dry, fluid, lubricated, skin, and internal.

Force spectroscopy is a set of techniques for the study of the interactions and the binding forces between individual molecules. These methods can be used to measure the mechanical properties of single polymer molecules or proteins, or individual chemical bonds. The name "force spectroscopy", although widely used in the scientific community, is somewhat misleading, because there is no true matter-radiation interaction.

<span class="mw-page-title-main">Atomic force microscopy</span> Type of microscopy

Atomic force microscopy (AFM) or scanning force microscopy (SFM) is a very-high-resolution type of scanning probe microscopy (SPM), with demonstrated resolution on the order of fractions of a nanometer, more than 1000 times better than the optical diffraction limit.

<span class="mw-page-title-main">Superlubricity</span> Regime of motion with little or no friction

In physics, superlubricity is a regime of motion in which friction vanishes or very nearly vanishes. However, a "vanishing" friction level is not clear, which makes the term vague. As an ad hoc definition, a kinetic coefficient of friction less than 0.01 can be adopted. This definition also requires further discussion and clarification.

Scanning probe microscopy (SPM) is a branch of microscopy that forms images of surfaces using a physical probe that scans the specimen. SPM was founded in 1981, with the invention of the scanning tunneling microscope, an instrument for imaging surfaces at the atomic level. The first successful scanning tunneling microscope experiment was done by Gerd Binnig and Heinrich Rohrer. The key to their success was using a feedback loop to regulate gap distance between the sample and the probe.

<span class="mw-page-title-main">Nanoelectromechanical systems</span> Class of devices for nanoscale functionality

Nanoelectromechanical systems (NEMS) are a class of devices integrating electrical and mechanical functionality on the nanoscale. NEMS form the next logical miniaturization step from so-called microelectromechanical systems, or MEMS devices. NEMS typically integrate transistor-like nanoelectronics with mechanical actuators, pumps, or motors, and may thereby form physical, biological, and chemical sensors. The name derives from typical device dimensions in the nanometer range, leading to low mass, high mechanical resonance frequencies, potentially large quantum mechanical effects such as zero point motion, and a high surface-to-volume ratio useful for surface-based sensing mechanisms. Applications include accelerometers and sensors to detect chemical substances in the air.

<span class="mw-page-title-main">Magnetic force microscope</span>

Magnetic force microscopy (MFM) is a variety of atomic force microscopy, in which a sharp magnetized tip scans a magnetic sample; the tip-sample magnetic interactions are detected and used to reconstruct the magnetic structure of the sample surface. Many kinds of magnetic interactions are measured by MFM, including magnetic dipole–dipole interaction. MFM scanning often uses non-contact AFM (NC-AFM) mode.

Nanoindentation, also called instrumented indentation testing, is a variety of indentation hardness tests applied to small volumes. Indentation is perhaps the most commonly applied means of testing the mechanical properties of materials. The nanoindentation technique was developed in the mid-1970s to measure the hardness of small volumes of material.

<span class="mw-page-title-main">Chemical force microscopy</span> Method of microscopy which measures chemical bonding between the probe and surface

In materials science, chemical force microscopy (CFM) is a variation of atomic force microscopy (AFM) which has become a versatile tool for characterization of materials surfaces. With AFM, structural morphology is probed using simple tapping or contact modes that utilize van der Waals interactions between tip and sample to maintain a constant probe deflection amplitude or maintain height while measuring tip deflection. CFM, on the other hand, uses chemical interactions between functionalized probe tip and sample. Choice chemistry is typically gold-coated tip and surface with R−SH thiols attached, R being the functional groups of interest. CFM enables the ability to determine the chemical nature of surfaces, irrespective of their specific morphology, and facilitates studies of basic chemical bonding enthalpy and surface energy. Typically, CFM is limited by thermal vibrations within the cantilever holding the probe. This limits force measurement resolution to ~1 pN which is still very suitable considering weak COOH/CH3 interactions are ~20 pN per pair. Hydrophobicity is used as the primary example throughout this consideration of CFM, but certainly any type of bonding can be probed with this method.

<span class="mw-page-title-main">Local oxidation nanolithography</span>

Local oxidation nanolithography (LON) is a tip-based nanofabrication method. It is based on the spatial confinement on an oxidation reaction under the sharp tip of an atomic force microscope.

<span class="mw-page-title-main">Conductive atomic force microscopy</span> Method of measuring the microscopic topography of a material

In microscopy, conductive atomic force microscopy (C-AFM) or current sensing atomic force microscopy (CS-AFM) is a mode in atomic force microscopy (AFM) that simultaneously measures the topography of a material and the electric current flow at the contact point of the tip with the surface of the sample. The topography is measured by detecting the deflection of the cantilever using an optical system, while the current is detected using a current-to-voltage preamplifier. The fact that the CAFM uses two different detection systems is a strong advantage compared to scanning tunneling microscopy (STM). Basically, in STM the topography picture is constructed based on the current flowing between the tip and the sample. Therefore, when a portion of a sample is scanned with an STM, it is not possible to discern if the current fluctuations are related to a change in the topography or to a change in the sample conductivity.

<span class="mw-page-title-main">Thermal scanning probe lithography</span>

Thermal scanning probe lithography (t-SPL) is a form of scanning probe lithography (SPL) whereby material is structured on the nanoscale using scanning probes, primarily through the application of thermal energy.

<span class="mw-page-title-main">Photoconductive atomic force microscopy</span> Type of atomic force microscopy

Photoconductive atomic force microscopy (PC-AFM) is a variant of atomic force microscopy that measures photoconductivity in addition to surface forces.

Nanosensors is a company that manufactures probes for use in atomic force microscopes (AFM) and scanning probe microscopes (SPM).

<span class="mw-page-title-main">Colloidal probe technique</span>

The colloidal probe technique is commonly used to measure interaction forces acting between colloidal particles and/or planar surfaces in air or in solution. This technique relies on the use of an atomic force microscope (AFM). However, instead of a cantilever with a sharp AFM tip, one uses the colloidal probe. The colloidal probe consists of a colloidal particle of few micrometers in diameter that is attached to an AFM cantilever. The colloidal probe technique can be used in the sphere-plane or sphere-sphere geometries. One typically achieves a force resolution between 1 and 100 pN and a distance resolution between 0.5 and 2 nm.

<span class="mw-page-title-main">Non-contact atomic force microscopy</span>

Non-contact atomic force microscopy (nc-AFM), also known as dynamic force microscopy (DFM), is a mode of atomic force microscopy, which itself is a type of scanning probe microscopy. In nc-AFM a sharp probe is moved close to the surface under study, the probe is then raster scanned across the surface, the image is then constructed from the force interactions during the scan. The probe is connected to a resonator, usually a silicon cantilever or a quartz crystal resonator. During measurements the sensor is driven so that it oscillates. The force interactions are measured either by measuring the change in amplitude of the oscillation at a constant frequency just off resonance or by measuring the change in resonant frequency directly using a feedback circuit to always drive the sensor on resonance.

<span class="mw-page-title-main">Infrared Nanospectroscopy (AFM-IR)</span> Infrared microscopy technique

AFM-IR or infrared nanospectroscopy is one of a family of techniques that are derived from a combination of two parent instrumental techniques. AFM-IR combines the chemical analysis power of infrared spectroscopy and the high-spatial resolution of scanning probe microscopy (SPM). The term was first used to denote a method that combined a tuneable free electron laser with an atomic force microscope equipped with a sharp probe that measured the local absorption of infrared light by a sample with nanoscale spatial resolution.

A probe tip is an instrument used in scanning probe microscopes (SPMs) to scan the surface of a sample and make nano-scale images of surfaces and structures. The probe tip is mounted on the end of a cantilever and can be as sharp as a single atom. In microscopy, probe tip geometry and the composition of both the tip and the surface being probed directly affect resolution and imaging quality. Tip size and shape are extremely important in monitoring and detecting interactions between surfaces. SPMs can precisely measure electrostatic forces, magnetic forces, chemical bonding, Van der Waals forces, and capillary forces. SPMs can also reveal the morphology and topography of a surface.

<span class="mw-page-title-main">Robert Carpick</span>

Robert William Carpick is a Canadian mechanical engineer. He is currently director of diversity, equity, and inclusion and John Henry Towne Professor in the Department of Mechanical Engineering and Applied Mechanics at the University of Pennsylvania. He is best known for his work in tribology, particularly nanotribology.

Bimodal Atomic Force Microscopy is an advanced atomic force microscopy technique characterized by generating high-spatial resolution maps of material properties. Topography, deformation, elastic modulus, viscosity coefficient or magnetic field maps might be generated. Bimodal AFM is based on the simultaneous excitation and detection of two eigenmodes (resonances) of a force microscope microcantilever.

References

  1. 1 2 Bhushan, Bharat (1999). Handbook of Micro/Nanotribology (2nd ed.). Boca Raton, Florida: CRC Press. pp. 1–833. ISBN   9780849384028.
  2. 1 2 3 4 Bhushan, Bharat (2017). Nanotribology and Nanomechanics: An Introduction (4th ed.). Springer. pp. 1–928. ISBN   978-3-319-51433-8.
  3. Krim, J. (1996). "Friction at the Atomic Scale". Scientific American. 275 (4): 74–80. Bibcode:1996SciAm.275d..74K. doi:10.1038/scientificamerican1096-74. JSTOR   24993406.
  4. Ringlein, James; Robbins, Mark O. (2004). "Understanding and illustrating the atomic origins of friction". Am. J. Phys. 72 (7): 884. Bibcode:2004AmJPh..72..884R. doi:10.1119/1.1715107.
  5. Muser, M.H.; Urbackh, M.; Robbins, M.O. (2003). "Statistical Mechanics of Static and Low-velocity kinetic friction". Advances in Chemical Physics. 126: 187. doi:10.1002/0471428019.ch5. ISBN   9780471235828.
  6. 1 2 Bhushan, B.; Israelachvili, J.N.; Landman, U. (1995). "Nanotribology: friction, wear and lubrication at the atomic scale". Nature. 374 (6523): 607–616. Bibcode:1995Natur.374..607B. doi:10.1038/374607a0. S2CID   4263053.
  7. Tabor, D.; Winterton, R. H. S. (1969-09-30). "The Direct Measurement of Normal and Retarded van der Waals Forces". Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences. 312 (1511): 435–450. Bibcode:1969RSPSA.312..435T. doi:10.1098/rspa.1969.0169. ISSN   1364-5021. S2CID   96200833.
  8. Binnig, G. (1982-01-01). "Surface Studies by Scanning Tunneling Microscopy". Physical Review Letters. 49 (1): 57–61. Bibcode:1982PhRvL..49...57B. doi: 10.1103/PhysRevLett.49.57 .
  9. Binnig, G.; Quate, C. F.; Gerber, Ch. (1986-03-03). "Atomic Force Microscope". Physical Review Letters. 56 (9): 930–933. Bibcode:1986PhRvL..56..930B. doi: 10.1103/PhysRevLett.56.930 . PMID   10033323.
  10. Neubauer, G.; Cohen, S.R.; Mcclelland, G.M.; Hajime, S. (1990). "Nanotribology of Diamond Films Studied by Atomic Force Microscopy". MRS Proceedings. 188: 219. doi:10.1557/PROC-188-219.
  11. Krim, J.; Solina, D.H.; Chiarello, R. (1991-01-14). "Nanotribology of a Kr monolayer: A Quartz-Crystal Microbalance Study of Atomic-Scale Friction". Physical Review Letters. 66 (2): 181–184. Bibcode:1991PhRvL..66..181K. doi:10.1103/PhysRevLett.66.181. PMID   10043531. S2CID   40001657.
  12. 1 2 3 Israelachvili, J; Min, Y; Akbulut, M; Alig, A; Carver, G; Greene, W; Kristiansen, K; Meyer, E; Pesika, N (2010). "Recent advances in the surface forces apparatus (SFA) technique". Reports on Progress in Physics. 73 (3): 036601. Bibcode:2010RPPh...73c6601I. doi:10.1088/0034-4885/73/3/036601. S2CID   53958134.
  13. 1 2 3 4 5 Szlufarska, Izabela; Chandross, Michael; Carpick, Robert W (2008). "Recent advances in single-asperity nanotribology". Journal of Physics D: Applied Physics. 41 (12): 123001. Bibcode:2008JPhD...41l3001S. doi:10.1088/0022-3727/41/12/123001. S2CID   11348039.
  14. Bhushan, Bharat (1995). "Nanotribology: Friction, wear and lubrication at the atomic scale". Nature. 374 (6523): 607. Bibcode:1995Natur.374..607B. doi:10.1038/374607a0. S2CID   4263053.
  15. Lucas, Marcel; Zhang, Xiaohua; Palaci, Ismael; Klinke, Christian; Tosatti, Erio; Riedo, Elisa (November 2009). "Hindered rolling and friction anisotropy in supported carbon nanotubes". Nature Materials. 8 (11): 876–881. arXiv: 1201.6487 . Bibcode:2009NatMa...8..876L. doi:10.1038/nmat2529. ISSN   1476-4660. PMID   19749768. S2CID   3844211.
  16. Smith, J. R.; Larson, C.; Campbell, S. A. (2011-01-01). "Recent applications of SEM and AFM for assessing topography of metal and related coatings — a review". Transactions of the IMF. 89 (1): 18–27. doi:10.1179/174591910X12922367327388. ISSN   0020-2967. S2CID   137321931.
  17. 1 2 Alvarez-Asencio, Rubén. "Nanotribology, Surface Interactions and Characterization: An AFM Study" (PDF).
  18. Liu, Yu. "Atomic Force Microscopy for Better Probing Surface Properties at Nanoscale: Calibration, Design and Application".
  19. Joyce, Stephen A.; Houston, J. E. (1991-03-01). "A new force sensor incorporating force-feedback control for interfacial force microscopy". Review of Scientific Instruments. 62 (3): 710–715. Bibcode:1991RScI...62..710J. doi:10.1063/1.1142072. ISSN   0034-6748.
  20. Joyce, Stephen A.; Houston, J. E.; Michalske, T. A. (1992-03-09). "Differentiation of topographical and chemical structures using an interfacial force microscope". Applied Physics Letters. 60 (10): 1175–1177. Bibcode:1992ApPhL..60.1175J. doi:10.1063/1.107396. ISSN   0003-6951.
  21. 1 2 3 4 5 6 Bhushan, Bharat (2013). Principles and applications of tribology, 2nd edition. New York: John Wiley & Sons, Ltd., Publication. ISBN   978-1-119-94454-6.
  22. Bhushan, Bharat (2013). Principles and applications of tribology, 2nd edition. New York: John Wiley & Sons, Ltd., Publication. pp. 711–713. ISBN   978-1-119-94454-6.
  23. Oliver, Warren C. (January 2004). "Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding and refinements to methodology". Journal of Materials Research. 19 (1): 3. Bibcode:2004JMatR..19....3O. doi:10.1557/jmr.2004.19.1.3. S2CID   135628097.
  24. Pethica, J. B.; Sutton, A. P. (1988-07-01). "On the stability of a tip and flat at very small separations". Journal of Vacuum Science & Technology A. 6 (4): 2490–2494. Bibcode:1988JVSTA...6.2490P. doi:10.1116/1.575577. ISSN   0734-2101.
  25. Pérez, Rubén; Štich, Ivan; Payne, Michael C.; Terakura, Kiyoyuki (1998-10-15). "Surface-tip interactions in noncontact atomic-force microscopy on reactive surfaces: Si(111)". Physical Review B. 58 (16): 10835–10849. Bibcode:1998PhRvB..5810835P. doi:10.1103/PhysRevB.58.10835.
  26. Abdurixit, A; Baratoff, A; Meyer, E (2000-04-02). "Molecular dynamics simulations of dynamic force microscopy: applications to the Si(111)-7×7 surface". Applied Surface Science. 157 (4): 355–360. arXiv: cond-mat/0003004 . Bibcode:2000ApSS..157..355A. doi:10.1016/S0169-4332(99)00566-8. S2CID   95706125.
  27. 1 2 3 4 Bhushan, Bharat (2013). Introduction to Tribology (2nd ed.). New York: Wiley. ISBN   9781118403259.
  28. Bhushan, Bharat; Israelachvili, Jacob N.; Landman, Uzi (1995-04-13). "Nanotribology: friction, wear and lubrication at the atomic scale". Nature. 374 (6523): 607–616. Bibcode:1995Natur.374..607B. doi:10.1038/374607a0. S2CID   4263053.
  29. Bhushan, Bharat (2008-04-28). "Nanotribology, nanomechanics and nanomaterials characterization". Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences. 366 (1869): 1351–1381. Bibcode:2008RSPTA.366.1351B. doi:10.1098/rsta.2007.2163. ISSN   1364-503X. PMID   18156126. S2CID   25593355.
  30. Hod, Oded (2012-08-20). "Superlubricity - a new perspective on an established paradigm". Physical Review B. 86 (7): 075444. arXiv: 1204.3749 . Bibcode:2012PhRvB..86g5444H. doi:10.1103/PhysRevB.86.075444. ISSN   1098-0121. S2CID   119251977.
  31. Hirano, Motohisa; Shinjo, Kazumasa; Kaneko, Reizo; Murata, Yoshitada (1997-02-24). "Observation of Superlubricity by Scanning Tunneling Microscopy". Physical Review Letters. 78 (8): 1448–1451. Bibcode:1997PhRvL..78.1448H. doi:10.1103/PhysRevLett.78.1448.
  32. Bennewitz, Roland (2007-01-01). "Friction Force Microscopy". In Gnecco, Dr Enrico; Meyer, Professor Dr Ernst (eds.). Fundamentals of Friction and Wear. NanoScience and Technology. Springer Berlin Heidelberg. pp. 1–14. doi:10.1007/978-3-540-36807-6_1. ISBN   9783540368069.
  33. Dienwiebel, Martin (2004-01-01). "Superlubricity of Graphite". Physical Review Letters. 92 (12): 126101. Bibcode:2004PhRvL..92l6101D. doi:10.1103/PhysRevLett.92.126101. PMID   15089689. S2CID   26811802.
  34. Socoliuc, Anisoara; Gnecco, Enrico; Maier, Sabine; Pfeiffer, Oliver; Baratoff, Alexis; Bennewitz, Roland; Meyer, Ernst (2006-07-14). "Atomic-Scale Control of Friction by Actuation of Nanometer-Sized Contacts". Science. 313 (5784): 207–210. Bibcode:2006Sci...313..207S. doi:10.1126/science.1125874. ISSN   0036-8075. PMID   16840695. S2CID   43269213.
  35. Popov, Valentin L. (2020). "Contacts With Negative Work of "Adhesion" and Superlubricity". Front. Mech. Eng. 5: 73. doi: 10.3389/fmech.2019.00073 . S2CID   210840278.
  36. Zhou, Yunong; Wang, Anle; Müser, Martin H. (2019). "How Thermal Fluctuations Affect Hard-Wall Repulsion and Thereby Hertzian Contact Mechanics". Frontiers in Mechanical Engineering. 5. doi: 10.3389/fmech.2019.00067 .
  37. Jinesh, K. B.; Krylov, S. Yu.; Valk, H.; Dienwiebel, M.; Frenken, J. W. M. (2008-10-30). "Thermolubricity in atomic-scale friction". Physical Review B. 78 (15): 155440. Bibcode:2008PhRvB..78o5440J. doi:10.1103/PhysRevB.78.155440.
  38. 1 2 Oliver, W.c.; Pharr, G.m. (1992-06-01). "An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments". Journal of Materials Research. 7 (6): 1564–1583. Bibcode:1992JMatR...7.1564O. doi:10.1557/JMR.1992.1564. ISSN   2044-5326. S2CID   137098960.
  39. Sneddon, Ian N. (1965-05-01). "The relation between load and penetration in the axisymmetric boussinesq problem for a punch of arbitrary profile". International Journal of Engineering Science. 3 (1): 47–57. doi:10.1016/0020-7225(65)90019-4.
  40. 1 2 Matthewson, M. J. (1986-11-24). "Adhesion measurement of thin films by indentation". Applied Physics Letters. 49 (21): 1426–1428. Bibcode:1986ApPhL..49.1426M. doi:10.1063/1.97343. ISSN   0003-6951.
  41. 1 2 Sirghi, L.; Rossi, F. (2006-12-11). "Adhesion and elasticity in nanoscale indentation". Applied Physics Letters. 89 (24): 243118. Bibcode:2006ApPhL..89x3118S. doi:10.1063/1.2404981. ISSN   0003-6951.
  42. Szoszkiewicz, Robert; Riedo, Elisa (2005-09-22). "Nucleation Time of Nanoscale Water Bridges". Physical Review Letters. 95 (13): 135502. Bibcode:2005PhRvL..95m5502S. doi:10.1103/PhysRevLett.95.135502. hdl: 1853/45727 . PMID   16197146.
  43. Bhushan, Bharat (2007-01-01). "Adhesion of multi-level hierarchical attachment systems in gecko feet". Journal of Adhesion Science and Technology. 21 (12–13): 1213–1258. doi:10.1163/156856107782328353. ISSN   0169-4243. S2CID   137062774.
  44. Achanta, Satish; Celis, Jean-Pierre (2007-01-01). Gnecco, Dr Enrico; Meyer, Professor Dr Ernst (eds.). Nanotribology of MEMS/NEMS. NanoScience and Technology. Springer Berlin Heidelberg. pp. 521–547. doi:10.1007/978-3-540-36807-6_23. ISBN   9783540368069.