Atomic force microscopy (AFM) is a type of scanning probe microscopy that images the topography of a sample surface by measuring the interaction forces between a sharp nanoscale probe attached to a flexible cantilever and the atoms on the sample.[1] The technique provides nanometer to atomic-scale resolution and can operate in various environments, including air, vacuum, and liquids, making it suitable for both conductive and insulating materials.[2] Unlike scanning tunneling microscopy, which requires electrical conductivity, AFM relies on mechanical forces such as van der Waals attractions and repulsive electrostatic interactions.[1]Invented in 1986 by Gerd Binnig, Calvin F. Quate, and Christoph Gerber, AFM combines principles from the stylus profilometer and scanning tunneling microscopy to overcome limitations in imaging non-conductive surfaces.[1] The core components include a sharp probe tip (typically 10–100 nm in radius) mounted on a cantilever, a piezoelectric scanner to raster the tip or sample, and a detection system—often a laser beam reflected from the cantilever onto a quadrant photodetector—that senses deflections as small as 0.1 nm.[3] A feedback loop adjusts the tip-sample distance to maintain constant force or oscillation amplitude, generating height maps that reveal surface features.[2]AFM operates in multiple modes to suit different applications: contact mode for direct surface contact and high-resolution topography; non-contact mode for imaging delicate samples using long-range attractive forces; and tapping mode, which oscillates the cantilever near its resonance frequency to minimize damage while enabling dynamic measurements.[2] Beyond imaging, AFM enables force spectroscopy to quantify mechanical properties like elasticity and adhesion, as well as manipulation of individual molecules.[2] Its versatility has made it indispensable in fields such as materials science for studying nanostructures, biology for visualizing biomolecules in native environments, and nanotechnology for device characterization.[2]
Introduction
Definition and basic principles
Atomic force microscopy (AFM) is a type of scanning probe microscopy (SPM) that images and measures surface properties at the nanoscale by raster-scanning a sharp probe mounted at the end of a flexible cantilever across a sample surface.[1] The technique enables high-resolution mapping of topography and mechanical, electrical, or magnetic properties without the limitations of sample conductivity required by methods like scanning tunneling microscopy.[4]In its basic operation, the probe tip interacts with the sample through short-range interatomic forces, such as attractive van der Waals forces or repulsive contact forces, causing minute deflections in the cantilever.[1] These deflections are precisely detected, often using laser reflection or interferometry, and converted into topographic data as the probe scans in a raster pattern. A feedback loop adjusts the probe-sample distance in real time to maintain constant force (in contact mode) or oscillation amplitude (in dynamic modes), ensuring stable imaging conditions.[1]The spatial resolution of AFM reaches atomic scales, typically 0.1–1 nm laterally and 0.01 nm vertically, owing to the probe tip's sharpness with a radius of about 10 nm.[5] This high precision arises from the localized nature of the tip-sample interaction, allowing visualization of individual atoms or molecules. Unlike electron microscopy techniques, which often require high-vacuum environments and can cause beam-induced damage, AFM operates non-destructively in ambient air, liquids, or vacuum, making it suitable for delicate biological and soft materials.[6]
Historical development
Atomic force microscopy (AFM) was invented in 1986 by Gerd Binnig, Calvin F. Quate, and Christoph Gerber at IBM Zurich Research Laboratory, building on the scanning tunneling microscope (STM) to enable imaging of non-conductive samples by measuring interatomic forces rather than tunneling currents. Their seminal paper in Physical Review Letters described the first AFM prototype, which used a diamond stylus on a cantilever to detect surface topography on an atomic scale through attractive and repulsive forces, achieving resolutions down to 3 nm laterally and less than 0.1 nm vertically on an insulating sapphire surface.[1] This innovation extended scanning probe techniques beyond conductive surfaces, earning indirect recognition through the 1986 Nobel Prize in Physics awarded to Binnig and Heinrich Rohrer for STM, which laid the groundwork for AFM.In the early 1990s, advancements addressed limitations of the initial contact-mode AFM, such as sample damage from lateral forces on soft materials. The tapping mode was introduced in 1993 by Qing Zhong, David Inniss, and colleagues from the Paul Hansma group at the University of California, Santa Barbara, involving oscillation of the cantilever near its resonant frequency to intermittently "tap" the surface, minimizing shear forces and enabling high-resolution imaging of delicate biological samples like DNA without deformation.[7] Concurrently, frequency modulation AFM (FM-AFM) was developed in 1991 by Franz J. Giessibl, Thomas R. Albrecht, and coworkers at IBM, using high-quality-factor cantilevers to detect minute frequency shifts caused by tip-sample interactions, which facilitated true non-contact operation with enhanced sensitivity for atomic-scale imaging in ultrahigh vacuum. These modes, detailed in key publications like Albrecht et al.'s Journal of Applied Physics article, became foundational, with FM-AFM gaining widespread adoption in the 2000s for precise force spectroscopy on surfaces.The 2000s marked further evolution toward dynamic and specialized applications, exemplified by high-speed AFM pioneered by Toshio Ando and his team at Kanazawa University, who achieved video-rate imaging of biomolecules in liquids by optimizing cantilever dynamics, scanners, and feedback systems, with a breakthrough practical system demonstrated in 2008 capable of 10 frames per second at sub-nanometer resolution. By the 2010s, cryogenic AFM emerged for studying quantum materials at low temperatures, such as superconductors and topological insulators; for instance, low-noise systems operating below 4 K enabled mapping of local electronic properties in graphene and other 2D materials without thermal drift.Up to 2025, recent innovations have integrated artificial intelligence (AI) for automated data analysis and operation, enhancing throughput and accuracy; machine learning algorithms enable real-time object detection and autonomous scanning, reducing manual intervention for large-scale biomechanical studies of cells.[8] These AI-driven tools, combined with multimodal integrations like combining AFM with optical or electrical measurements, continue to expand AFM's utility in materials science and biology, building on its foundational milestones.
Instrument Components
Probe and cantilever
The atomic force microscopy (AFM) probe is composed of a sharp nanotip mounted at the free end of a flexible microcantilever, serving as the primary force-sensing element that interacts with the sample surface.[9] The cantilever is typically a microfabricated beam with lengths ranging from 100 to 500 μm, designed to exhibit high sensitivity to minute forces while maintaining mechanical stability.[9] Common geometries include rectangular and V-shaped (or triangular) cantilevers, where rectangular designs offer uniform stiffness along the length for precise deflection measurements, and V-shaped configurations provide enhanced resistance to torsional motion, reducing lateral bending artifacts during scanning.[10] These cantilevers are fabricated from materials such as silicon (Si) or silicon nitride (Si₃N₄) using photolithography, anisotropic etching, and deposition techniques to achieve precise dimensions and low mass. To facilitate optical detection of deflection, the dorsal surface of the cantilever is coated with a thin reflective layer, usually aluminum or gold, approximately 20-30 nm thick, which enhances laser reflectivity by up to 2.5 times.[11]The spring constant k of the cantilever, which quantifies its stiffness, typically spans 0.01 to 100 N/m, allowing adaptation to various imaging conditions; softer cantilevers with k < 0.1 N/m are specialized for biological samples to minimize damage to delicate structures like cells or polymers.[12] The force F experienced by the cantilever is related to its deflection \delta through Hooke's law: F = -k \delta. This linear relationship arises from the cantilever's elastic deformation under small displacements, modeled as a simple harmonic oscillator where the restoring force opposes the applied interaction. To derive this, consider the cantilever as a beam fixed at one end; the differential equation for bending under a point load at the free end yields the deflection \delta = \frac{F L^3}{3 E I}, where L is length, E is Young's modulus, and I is the moment of inertia. The effective spring constant is then k = \frac{3 E I}{L^3}, simplifying to F = -k \delta for the quasistatic regime relevant to AFM force sensing.[13]The probe tip, integral to the cantilever, features a sharp apex with radii of 2-50 nm to achieve atomic-scale resolution, fabricated from Si, Si₃N₄, or diamond-coated variants for enhanced durability.[14] Tip fabrication often involves isotropic etching to form the initial pyramid, followed by focused ion beam (FIB) milling or oxidation sharpening to refine the apex geometry. Diamond coatings, applied via chemical vapor deposition, extend tip lifetime in abrasive environments.[14]Despite these advancements, challenges persist in probe performance, including tip wear from repeated sample interactions, which can blunt the apex and degrade resolution over scans.[15] Contamination, such as adsorbed hydrocarbons or debris pickup, alters the effective tip shape and introduces imaging artifacts, particularly in ambient conditions.[16] Additionally, tip geometry influences the measured interaction forces and topographic fidelity, as non-ideal shapes like broadened sidewalls convolute the sample profile, limiting lateral resolution.[17]
Piezoelectric scanner
The piezoelectric scanner serves as the core actuation mechanism in atomic force microscopy (AFM), enabling precise three-dimensional positioning of either the probe tip or the sample surface to facilitate high-resolution scanning. These scanners are predominantly constructed from lead zirconate titanate (PZT), a ferroelectric ceramic valued for its strong electromechanical coupling, and are designed in forms such as piezoelectric tubes or stacked actuators to generate motion along the x, y, and z axes. Piezoelectric tube scanners, featuring a cylindrical structure with four quartered outer electrodes for lateral (xy) scanning and a continuous inner electrode for vertical (z) motion, offer a compact and integrated solution for orthogonal displacement without mechanical linkages. Stack actuators, composed of layered PZT elements bonded together, provide higher force output and are often used in modular configurations for enhanced rigidity.[18][19]The fundamental operating principle of the piezoelectric scanner exploits the converse piezoelectric effect, wherein an applied electric field induces proportional mechanical strain in the material, resulting in controlled expansion or contraction. For PZT materials commonly employed in AFM, the longitudinal piezoelectric strain constant d_{33} is approximately 500 pm/V, enabling sub-nanometer positioning resolution critical for atomic-scale imaging. The resulting displacement \Delta L along the piezo element's length is given by the equation\Delta L = d \cdot V \cdot \frac{L}{t},where d represents the piezoelectric constant (e.g., d_{33}), V is the applied voltage, L is the effective length of the piezo element, and t is its thickness. This relation stems from the strain S = d \cdot E produced under the electric field E = V / t, yielding \Delta L = S \cdot L for free displacement in the absence of external loads. Typical operational ranges for such scanners include approximately 100 \mum \times 100 \mum in the xy plane and 10 \mum in the z direction, balancing field of view with nanoscale precision.[20][21][22]Despite their responsiveness, piezoelectric scanners suffer from inherent nonlinearities, including hysteresis (path-dependent displacement up to 15-20% of full range), creep (slow drift over time), and voltage nonlinearity, which can introduce positioning errors and image artifacts in AFM. These limitations are effectively mitigated through closed-loop control architectures that integrate direct position-sensing elements, such as embedded strain gauges, interferometers, or capacitive sensors, to monitor actual displacement and apply corrective voltages in real time. Such feedback systems achieve positioning accuracies below 1 nm, ensuring reliable operation for extended imaging sessions.[23][24]Piezoelectric scanners in AFM are implemented in two primary configurations: sample-scanning, where the scanner moves the specimen beneath a stationary probe, and tip-scanning, where the probe is actuated over a fixed sample. Sample-scanning designs accommodate larger or non-planar specimens and minimize probe drift influences, whereas tip-scanning configurations facilitate straightforward alignment with auxiliary optics and reduce sample perturbation. The choice between these depends on application requirements, with tube scanners often favored in tip-scanning setups for their simplicity and low mass.[25]
Detection and feedback systems
In atomic force microscopy (AFM), cantilever deflection is primarily detected using optical beam deflection, where a laser beam is directed onto the back of the cantilever and reflected onto a quadrant photodiode to measure angular changes corresponding to deflection.[26] This method offers high sensitivity, typically achieving sub-nanometer resolution, and is the most widely adopted due to its simplicity and compatibility with various environments.[26] Interferometric detection serves as an alternative, employing a laser interferometer to directly measure the cantilever's displacement through interference patterns, providing enhanced precision in low-noise conditions but requiring more complex alignment.[27] Piezoelectric strain gauges, often implemented as piezoresistive elements integrated into the cantilever, detect deflection via changes in electrical resistance induced by mechanical strain, enabling compact, self-contained sensing without external optics, though with potentially lower sensitivity compared to optical methods.[28]The detected cantilever deflection feeds into a closed-loop control system that maintains stable tip-sample interaction during scanning. A proportional-integral-derivative (PID) controller processes the signal to adjust the voltage applied to the z-axis piezoelectric scanner, ensuring the tip follows the sample topography without excessive force or separation.[29] The core of this feedback is the error signal, defined as e = setpoint - measured deflection (or oscillation amplitude in dynamic modes), which drives the PID response to minimize deviations.[30] The setpoint represents the desired interaction level, such as a constant force or amplitude, calibrated based on cantilever spring constant properties.[29]Feedback configurations vary to suit different applications: in constant force mode, the PID actively adjusts the z-piezo to keep the cantilever deflection (and thus applied force) at the setpoint, enabling gentle imaging of delicate samples.[31] Conversely, constant height mode disables z-feedback, fixing the scanner height while allowing deflection to vary with topography, which supports faster scans on flat surfaces but risks tip damage on rough ones.[31] Performance is limited by noise sources, including thermal fluctuations that excite cantilever vibrations according to the equipartition theorem, electronic noise from detection electronics degrading signal-to-noise ratio, and environmental vibrations from acoustics or building motion that couple into the system.[32][33] These are mitigated through vibration isolation, low-temperature operation, and optimized PID gains to achieve atomic-scale stability.[33]
Fundamental Principles
Interatomic forces and interactions
The interatomic forces governing atomic force microscopy (AFM) arise primarily from interactions between atoms or molecules at the probe tip and sample surface, enabling nanoscale imaging and manipulation. These forces encompass van der Waals attractions, which dominate short-range interactions; capillary forces from adsorbed moisture; electrostatic forces due to charge distributions; and magnetic forces in systems involving ferromagnetic materials.[34] Van der Waals forces, typically attractive and on the order of $10^{-9} N at separations of approximately 1 nm, stem from transient dipole-induced dipole couplings between neutral atoms.[34]Capillary forces, often the strongest in ambient conditions, result from the liquid meniscus formed by water vapor adsorbing onto hydrophilic surfaces, leading to adhesion strengths up to several nanonewtons.[35] Electrostatic forces emerge from Coulombic interactions between charged tip and sample, scaling inversely with distance squared, while magnetic forces, relevant for magnetic domain imaging, arise from dipole alignments in magnetized samples and can extend over tens of nanometers.[34]Force-distance curves, obtained by measuring cantilever deflection as the tip approaches and retracts from the sample, reveal the interplay of these forces. During approach, the curve shows an initial flat region at large separations where forces are negligible, followed by a steep "snap-in" instability when attractive forces exceed the cantilever's restoring force, causing rapid tip-sample contact.[34] The retract curve exhibits hysteresis, with a characteristic adhesion peak where pull-off occurs, often due to capillary or van der Waals bonding, and can show long-range tails from electrostatic or magnetic contributions.[34] This hysteresis quantifies energy dissipation and adhesion, with snap-in distances typically 5–20 nm in air, influenced by humidity and surface chemistry.[34]The van der Waals interaction is commonly modeled by the Lennard-Jones potential, which approximates the pairwise atomic potential energy as a function of separation r:V(r) = 4\epsilon \left[ \left( \frac{\sigma}{r} \right)^{12} - \left( \frac{\sigma}{r} \right)^6 \right]Here, \epsilon represents the binding energy (depth of the potential well), and \sigma is the distance where V(r) = 0, typically around 0.3–0.4 nm for many materials.[36] The corresponding force is the negative gradient:F(r) = -\frac{dV}{dr} = 24 \frac{\epsilon}{\sigma} \left[ 2 \left( \frac{\sigma}{r} \right)^{13} - \left( \frac{\sigma}{r} \right)^7 \right]This yields a short-range repulsive term (proportional to r^{-13}) from Pauli exclusion and orbital overlap, balancing the longer-range attractive term (proportional to r^{-7}) from dispersion forces; the minimum occurs at r \approx 1.1\sigma, with force zero at equilibrium.[36] For macroscopic tip-sample interactions, the potential is integrated over volumes, leading to a force scaling as F \propto -A R / (6 z^2) for a spherical tip of radius R at separation z, where A is the Hamaker constant.[34]The Hamaker constant A, a material-specific parameter encapsulating van der Waals strengths, is given by A = \pi^2 C \rho_1 \rho_2, with C the Lennard-Jones dispersion coefficient and \rho_{1,2} the atomic densities of tip and sample.[34] Typical values range from $10^{-20} to $10^{-19} J; for example, A \approx 6.5 \times 10^{-20} J for silica-vacuum-silica interactions, increasing in liquids due to medium effects. This constant allows prediction of interaction strengths across material pairs, essential for interpreting AFM data.[34]Tip geometry significantly modulates these forces, with long-range interactions (e.g., electrostatic, magnetic) integrating over the entire tip cone or cantilever, while short-range van der Waals and contact repulsion localize at the apex.[34] For conical or pyramidal tips, the effective radius affects force gradients; sharper tips (radii <10 nm) enhance resolution of atomic-scale features but amplify capillary adhesion via larger meniscus volumes.[35] In humid environments, water meniscus formation around the tipgeometry can increase adhesion by 10–100 times compared to dry conditions, with force scaling proportional to tip radius and contact angle.[35]
Cantilever deflection and force measurement
In atomic force microscopy (AFM), the forces between the probe tip and the sample surface cause the cantilever to deflect, enabling quantitative measurement of these interactions. The deflection δ follows Hooke's law, expressed as F = k \delta, where F is the force, k is the cantilever's springconstant, and δ represents the vertical displacement at the free end.[37] This relationship allows forces on the order of piconewtons to be detected, as typical spring constants range from 0.01 to 100 N/m depending on the cantilever design.For a rectangular cantilever under small deflections, the mechanics of bending can be described using Euler-Bernoulli beam theory, where the radius of curvature R at the base approximates R = L^2 / (2\delta), with L being the cantilever length. This curvature arises from the applied force, leading to a deflection profile that is cubic along the length. The resulting beam deflection angle \theta at the free end is given by \theta \approx 3\delta / (2L), which relates the end displacement to the angular tilt observed in detection systems.[38] In practice, deflection is quantified using the optical lever method, where a laser beam reflects off the cantilever backside onto a photodetector; the sensitivity s = \delta / V converts the photodetector voltage V to actual displacement δ, typically calibrated by pressing the tip against a rigid surface.[39]Accurate force measurement requires precise calibration of the spring constant k, often achieved via the thermal noise method, which exploits the cantilever's Brownian motion in equilibrium. The method applies the equipartition theorem to the fundamental flexural mode, yielding k = 0.817 k_B T / \langle z^2 \rangle, where k_B is the Boltzmann constant, T is the absolute temperature, and \langle z^2 \rangle is the mean-square displacement derived from the power spectral density of thermal fluctuations.[37] This nondestructive technique, introduced by Hutter and Bechhoefer, provides accuracies of 5-10% for soft cantilevers and is widely adopted due to its simplicity and applicability in various environments.[40]Quantitative force mapping involves converting raw deflection signals—obtained as voltage traces from the optical lever—into force units by applying the calibrated k and s, often pixel-by-pixel across scanned areas to generate force-volume images. This process enables spatially resolved mapping of mechanical properties, such as stiffness variations on heterogeneous samples. However, the overall accuracy of force measurements is highly sensitive to the precision of k, with uncertainties typically ranging from 10% to 30% depending on the calibrationmethod and cantilever geometry; errors in k propagate directly to force values, limiting quantitative reliability in applications like biomolecular force spectroscopy.[38]
Operational Modes
Contact mode
Contact mode is the original and simplest operational mode of atomic force microscopy (AFM), in which the sharp probe tip attached to a flexible cantilever maintains continuous physical contact with the sample surface throughout the imaging process. The cantilever deflection, caused by the tip-sample interaction force, is monitored using a detection system, typically involving a laser beam reflected off the cantilever onto a photodetector. As the piezoelectric scanner rasters the tip or sample in the xy-plane, a feedback loop continuously adjusts the z-position of the scanner to maintain a constant cantilever deflection, corresponding to a setpoint force between 10^{-9} and 10^{-7} N. This z-displacement data is recorded to construct a topographic map of the surface height variations with nanometer lateral resolution and sub-angstrom vertical sensitivity.[41]In this mode, the dominant interactions occur in the repulsive force regime, where the tip-sample separation is less than approximately 0.5 nm, compressing the electron clouds of surface atoms and generating short-range Pauli exclusion forces that outweigh longer-range attractive van der Waals interactions. The constant forcefeedback ensures the tip follows the surface contour without excessive penetration, though the exact force setpoint must be carefully tuned to avoid sample deformation on compliant materials. This static contact approach contrasts with dynamic modes by relying solely on DC deflection signals rather than oscillatory responses.[42]The primary advantages of contact mode include its straightforward instrumentation, which requires no additional oscillators or drivers, enabling high scanning speeds up to several lines per second and suitability for imaging rough or steeply sloped surfaces on rigid samples without losing contact. It also supports atomic-scale resolution on crystalline materials like graphite or mica under ultra-high vacuum conditions, making it ideal for fundamental surface studies. However, the continuous sliding contact introduces significant lateral shear forces, which accelerate tip wear and can cause frictional damage or displacement of delicate features on soft or loosely bound samples, limiting its use to hard, stable substrates.[43]Contact mode finds widespread applications in topographic characterization of robust materials, such as metals, semiconductors, and inorganic thin films, where it provides reliable height profiles for quality control in microfabrication processes. For instance, it has been employed to image silicon wafer surfaces and gold films, revealing surface roughness and step edges with sub-nanometer precision essential for semiconductor device development. While less common for biological samples due to potential damage, minimized-force variants have been adapted for select rigid biomolecules in controlled environments.[44][45]
Tapping mode
Tapping mode operates by oscillating the cantilever close to its resonance frequency, typically ranging from 30 to 400 kHz for soft cantilevers, such that the sharp tip at its end makes intermittent contact with the sample surface during each cycle. This dynamic approach, introduced in 1993, allows the tip to "tap" the surface lightly, with the amplitude of oscillation decreasing upon interaction due to short-range repulsive forces dominating near the surface.[46] The feedback loop monitors this amplitude reduction and adjusts the vertical position of the piezoelectric scanner to maintain a user-defined setpoint, enabling topographic mapping with atomic-scale resolution while the raster scan proceeds laterally.The setpoint amplitude is conventionally set to 70-90% of the free oscillation amplitude (the amplitude in air away from the sample) to balance imaging stability and minimize damage; values closer to 70% ensure sufficient contact for resolution, while higher ratios prevent the snap-in instability from capillary or van der Waals forces pulling the tip into continuous contact. Feedback can alternatively track phase shifts instead of amplitude for enhanced contrast in certain materials, though amplitude control remains standard for topography.A key advantage of tapping mode is the drastic reduction of lateral shear and frictional forces compared to contact mode, as contact occurs only briefly at the bottom of each oscillation cycle, preserving fragile structures. This makes it ideal for imaging soft, adhesive, or deformable samples such as polymers, biological tissues, and nanostructures, where continuous contact would cause deformation, wear, or displacement. For instance, tapping mode has enabled high-fidelity imaging of DNA strands and protein assemblies on substrates without altering their native conformation.[47]Phase imaging, an extension of tapping mode, captures the phase lag φ between the driving oscillation and the cantilever's response, revealing material-specific contrasts beyond topography. This lag arises from the interplay of elastic and dissipative tip-sample interactions, quantified approximately as
\phi = \tan^{-1}\left( \frac{\text{dissipative component}}{\text{elastic component}} \right),
where the dissipative term reflects viscous losses and adhesion, and the elastic term corresponds to the conservative force gradient. Lags near 90° indicate purely elastic responses (minimal energy loss), while deviations toward 0° or 180° signal dissipative behavior, enabling mapping of viscoelastic properties like stiffness variations in composite materials.The mode's performance depends on the cantilever's quality factor Q = \frac{f_0}{\Delta f}, where f_0 is the resonance frequency and \Delta f the bandwidth at half-maximum amplitude; higher Q (often 100-500 in air) amplifies small force-induced amplitude shifts, enhancing sensitivity but potentially slowing response in liquids where Q drops. Cantilever properties, such as spring constant and resonance, are optimized for tapping to ensure stable oscillations, as detailed in probe design considerations.
Non-contact mode
In non-contact atomic force microscopy (NC-AFM), the cantilever oscillates with a small amplitude typically less than 10 nm near its resonancefrequency, remaining in the attractive force regime without physical contact between the tip and sample surface. The interaction between the tip and sample modifies the effective spring constant of the cantilever due to the force gradient, causing a shift in the oscillation frequency, denoted as Δf, which serves as the primary signal for imaging. This frequency shift arises from the conservative component of the tip-sample interaction and is particularly sensitive to long-range van der Waals forces and short-range chemical bonding forces.[48]The relationship between the frequency shift and the force gradient is given by the equation\Delta f = -\frac{f_0}{2k} \frac{\partial F}{\partial z},where f_0 is the free resonance frequency of the cantilever, k is its spring constant, F is the tip-sample force, and z is the tip-sample separation. This approximation holds under the small-amplitude limit and harmonic oscillation assumption, linking the measurable Δf directly to the spatial derivative of the interaction force, \partial F / \partial z. The negative sign indicates that an attractive force gradient (typical in NC-AFM) decreases the resonance frequency. This formulation, derived from the dynamics of a driven harmonic oscillator perturbed by a force gradient, enables quantitative mapping of interaction potentials without direct forcemeasurement.[4]A prominent implementation is frequency modulation atomic force microscopy (FM-AFM), where the cantilever is driven at constant amplitude, and feedback maintains a constant frequency shift Δf to track surface topography. In this mode, ultra-high resolution, including sub-atomic contrast on insulating surfaces like silicon dioxide or graphite, has been achieved using stiff sensors such as the qPlus tuning fork, which offers low noise and high force sensitivity. For instance, bond-order discrimination and intra-atomic features have been resolved on insulators, surpassing conductive sample capabilities of scanning tunneling microscopy. Recent qPlus advancements in the 2020s, including optimized higher eigenmodes and biaxial sensing, have further enhanced stability and resolution for complex surface reconstructions.[48][4][49]NC-AFM offers key advantages, including minimal tip and sample wear due to the absence of contact, enabling prolonged imaging of delicate structures, and providing true atomic contrast from attractive interactions that reveal chemical specificity. However, it faces challenges such as high sensitivity to environmental noise, necessitating ultrahigh vacuum conditions for optimal performance, and slower scan speeds compared to contact modes due to the need for precise frequency demodulation. True NC-AFM, emphasizing FM detection in the attractive regime, differs from amplitude modulation variants, which may inadvertently enter short-range repulsive interactions with larger amplitudes, potentially compromising non-contact purity.[4]
Advanced Techniques
Force spectroscopy
Force spectroscopy in atomic force microscopy (AFM) measures the interaction forces between the AFM tip and the sample as a function of their separation distance, enabling quantitative assessment of mechanical and adhesive properties at the nanoscale. The procedure involves performing repeated approach and retraction cycles of the cantilever at a fixed lateral (x-y) position, while recording the cantilever deflection to plot force versus tip-sample separation (z). During approach, the tip moves toward the sample until contact occurs, often exhibiting an attractive regime due to van der Waals or other interatomic forces; retraction then pulls the tip away, potentially revealing adhesion through hysteresis in the curve. These force-distance curves provide insights into elasticity, adhesion, and binding events without relying on imaging modes.[50]In the contact region of the force-distance curve, where the tip indents the sample elastically, the data are analyzed by fitting to the Hertz contact model, which assumes a spherical tip and non-adhesive elastic deformation. The model describes the force-indentation relationship as:F = \frac{4}{3} E^* \sqrt{R} \, \delta^{3/2}Here, F is the applied force, E^* is the reduced elastic modulus (incorporating the Young's moduli of both sample and tip), R is the tip radius of curvature, and \delta is the indentation depth. This fitting yields the sample's Young's modulus, a key indicator of stiffness, particularly useful for soft materials like polymers or biological tissues where values range from kilopascals to megapascals. The model's validity requires small indentations (typically <10% of sample thickness) and negligible adhesion, though extensions like the Johnson-Kendall-Roberts model account for adhesive contributions when needed.[51][50]Applications of force spectroscopy include quantifying molecular binding forces and mechanical properties of materials and biomolecules. For instance, it measures unbinding forces in ligand-receptor pairs, such as biotin-avidin interactions, which occur on the piconewton scale (e.g., 100-200 pN under physiological conditions), revealing energy landscapes of specific bonds. In materials science, Young's modulus mapping via force spectroscopy assesses elasticity variations, as in nanocomposites or thin films, establishing spatial heterogeneity in mechanical response. For biological macromolecules like proteins or DNA, the technique probes unfolding dynamics using the worm-like chain (WLC) model, which interpolates the entropic elasticity of semi-flexible polymers:F(x) = \frac{k_B T}{p} \left[ \frac{1}{4} \left(1 - \frac{x}{L}\right)^{-2} - \frac{1}{4} + \frac{x}{L} \right]where F is the force, x the end-to-end extension, L the contour length, p the persistence length, k_B Boltzmann's constant, and T temperature. Fitting WLC to retraction curves from proteins like titin quantifies domain stability and folding pathways. As of 2025, advances in high-speed nanomechanical mapping have enabled quantitative assessment of dynamic properties in biological systems with improved spatial resolution and speed.[52][53][54]Variants of force spectroscopy enhance its efficiency for mapping properties over areas. Force-volume mode extends single-point measurements by acquiring force-distance curves on a grid of points, enabling 3D maps of elasticity, adhesion, and dissipation while avoiding lateral shear forces that could damage samples. PeakForce Tapping, a high-speed variant, oscillates the cantilever at low amplitudes (e.g., 100-500 Hz) to perform rapid approach-retraction cycles, allowing real-time extraction of mechanical parameters like modulus and adhesion at scan rates up to 8 lines per second, ideal for dynamic or heterogeneous samples. These techniques prioritize quantitative property extraction over topography, with interatomic forces influencing the non-contact portions of curves as detailed elsewhere.[50][55]
High-resolution atomic imaging
High-resolution atomic imaging in atomic force microscopy (AFM) relies on advanced non-contact modes, particularly frequency modulation AFM (FM-AFM) conducted in ultra-high vacuum (UHV) environments to minimize contamination and noise. In FM-AFM, the cantilever oscillates at its resonance frequency, and short-range tip-sample interactions cause detectable frequency shifts, enabling visualization of surface atomic structures with sub-angstrom precision. This technique has become the standard for true atomic resolution in vacuum, as it sensitively probes conservative and dissipative forces without physical contact, achieving lateral resolutions below 0.1 nm.[4]A pivotal advancement involves the use of functionalized tips, such as those terminated with carbon monoxide (CO), which enhance contrast for chemical identification and bond-order discrimination. By positioning the CO molecule at the tip apex, researchers can resolve intramolecular features based on Pauli repulsion and bonding interactions, distinguishing single, double, and triple carbon-carbon bonds in polycyclic aromatic hydrocarbons through variations in frequency shift images. This approach, demonstrated in 2012 by Gross et al., allows for the identification of atomic species and molecular configurations on insulating surfaces like NaCl.[56] FM-AFM with such tips has imaged complex lattices, including the reconstructed Si(111)-(7×7) surface with atomic resolution of approximately 0.6 nm laterally and 0.01 nm vertically, revealing adatom arrangements and corner holes. Similarly, the hexagonal lattice of graphene has been resolved at the atomic scale, confirming its honeycomb structure and defects like vacancies.[48][57]Frequency shift maps in these setups provide chemical contrast by mapping variations in short-range forces, enabling differentiation between atoms of distinct electronic properties, such as metal (e.g., iron) versus insulator atoms, where metallic atoms exhibit stronger attractive interactions at larger separations. For instance, spectra of frequency shifts and forces distinguish Fe from Co adatoms based on their orbital signatures. However, challenges persist, including maintaining tip apex stability to prevent reconfiguration or dulling during scanning, and mitigating drift from thermal fluctuations or piezo creep, which can shift images by several angstroms over minutes. Low-temperature operation (below 5 K) and qPlus sensors help stabilize the tip, while active feedback reduces drift effects.Recent progress includes 3D force mapping, where frequency shifts are recorded across a range of tip-sample distances to reconstruct volumetric data, revealing subsurface atomic positions. This has enabled imaging of buried atoms in materials like graphite or thin films with nanoscale depth resolution, extending visualization beyond surface layers. A landmark example is the 2009 demonstration by Custance, Pérez, and Morita of AFM-based atomic manipulation on Si(111)-(7×7), where controlled lateral forces repositioned single silicon adatoms without subsurface damage, highlighting the technique's precision for building atomic structures.[58] As of 2025, high-speed AFM techniques with flexible fitting methods have advanced real-time observation of dynamic processes at atomic resolution.
Data Acquisition and Processing
Topographic image formation
In atomic force microscopy (AFM), topographic images are generated through a raster scanning process, where the probe systematically traverses the sample surface in a two-dimensional grid. The scan proceeds along a fast axis (typically the x-direction) in a continuous back-and-forth motion, with discrete steps along the slow axis (y-direction) between each line. This pattern mimics the line-by-line acquisition of pixels in digital imaging, allowing the instrument to map surface features point by point. A feedback control system plays a central role, continuously monitoring the cantilever-sample interaction—such as deflection in contact mode or oscillation amplitude in tapping mode—and adjusting the vertical position (z) of the scanner to maintain a predefined setpoint. At each (x, y) coordinate, the required z-displacement is recorded, forming a height matrix z(x, y) that quantitatively represents the sample's topography. This matrix directly encodes the surface elevation variations, enabling the reconstruction of three-dimensional surface profiles.The collected height data is rendered into visual topographic maps, often displayed as grayscale or false-color images where color intensity or hue corresponds to height values, facilitating intuitive interpretation of surface undulations. For instance, warmer colors might denote peaks while cooler tones indicate valleys, with the scale bar providing quantitative height references. Complementary channels enhance visualization: deflection images capture the raw cantilever bending signal, revealing slope and edge contrasts beyond pure height; error images, derived from the difference between the setpoint and actual interaction, highlight feedback deviations and subtle topographic transitions. These multi-channel representations are typically generated during or immediately after data acquisition, allowing users to assess image quality in near real-time.Lateral resolution in topographic imaging is fundamentally constrained by the AFM tipgeometry and operational parameters. Tipconvolution arises from the finite size of the probe, which geometrically dilates sharp surface features, broadening their apparent dimensions in the image. For a spherical tip of radius R imaging a step-like feature of height h (where h \ll R), the dilation effect results in an apparent width w approximated by the equation:w = 2 \sqrt{2 R h}This formula illustrates how larger tip radii exacerbate broadening, limiting the ability to resolve fine details; for example, a 10 nm radius tip might widen a 1 nm high step to approximately 8.9 nm. Scan speed further influences resolution, as faster rates can introduce dynamic artifacts like overshoot or undersampling if the feedback loop's bandwidth is exceeded, though optimized speeds balance imaging time and fidelity. Tip sharpness, often below 10 nm for standard probes, thus sets the practical lateral limit, typically achieving sub-10 nmresolution on ideal samples.Visualization of topographic data occurs in both real-time and post-processing modes. During scanning, preliminary images update progressively on the instrument's display, providing immediate feedback for adjustments like setpoint tuning or scan area refinement. Post-acquisition processing, however, refines these raw matrices through leveling, filtering, or 3D rendering software, yielding polished visualizations that better isolate true topography from scanner nonlinearities. This dual approach ensures efficient data collection while maximizing interpretive accuracy.
Image analysis and artifacts
Image analysis in atomic force microscopy (AFM) involves processing raw topographic data to extract meaningful surface features while identifying and mitigating artifacts that can distort interpretations. Artifacts arise from instrumental limitations, such as scanner nonlinearity or probe geometry, and noise from environmental or electronic sources, necessitating post-acquisition corrections to ensure reliable quantitative results.[59]Common artifacts include piezo creep, which causes nonlinear scanning distortions due to the viscoelastic response of piezoelectric actuators, leading to bowed or stretched features in images, particularly during initial scans or direction changes.[60]Tip wear results in broadening of surface features, as progressive dulling of the probe tip convolves with the sample topography, exaggerating widths and reducing lateral resolution over repeated imaging sessions.[59]Feedback oscillations, or ringing, occur when the feedback loop gain is excessively high, producing wavy patterns or stripes in the image as the scanner overcorrects height variations.Correction methods address these issues through specialized processing techniques. Line-by-line or planar flattening removes tilt and curvature from piezo creep by subtracting fitted polynomials or planes from the height data, restoring accurate sample geometry without altering local features.[61] For tip-induced broadening, blind tip reconstruction algorithms deconvolve the probe shape from the image, estimating the tipgeometry from sharp sample edges and reconstructing the true surface profile, often using iterative erosion models to minimize convolution effects.[62]Noise in AFM images primarily consists of thermal noise, arising from cantilever fluctuations, with an amplitude spectral density approximated by \sqrt{\frac{4 k_B T k}{\omega Q}}, where k_B is Boltzmann's constant, T is temperature, k is the spring constant, \omega is the angular frequency, and Q is the quality factor; this limits resolution in dynamic modes.[63] Additionally, 1/f noise, or flicker noise, introduces low-frequency drifts from electronic instabilities or sample charging, manifesting as gradual baseline shifts across the scan.[64] Filtering via fast Fourier transform (FFT) mitigates these by isolating and attenuating specific frequency bands, such as removing high-frequency thermal components or low-frequency 1/f trends while preserving topographic signals.[65]Quantitative analysis of corrected images employs specialized software like Gwyddion, an open-source tool for scanning probe microscopy data that computes surface roughness parameters such as the arithmetic average roughness R_a (mean deviation from the centerline) and root mean square roughness R_q (standard deviation of heights), enabling statistical evaluation of texture over defined areas.[66] Similarly, MountainsMap provides advanced areal roughness analysis compliant with ISO standards, integrating R_a and R_q calculations with 3D visualization for profilometric data from AFM scans.[67]Validation of processed AFM images involves cross-checking with complementary techniques, such as scanning electron microscopy (SEM) for morphological confirmation, where SEM's chemical contrast verifies feature dimensions against AFM topography, or computational simulations that model expected image distortions to benchmark correction efficacy.[68][69]
Atomic force microscopy (AFM) plays a pivotal role in materials science and nanotechnology by enabling high-resolution topographic characterization of nanostructures such as nanoparticles, thin films, and graphene sheets. For nanoparticles, AFM provides precise measurements of size, shape, and distribution on substrates, often revealing surface features down to the sub-nanometer scale that are critical for optimizing their integration into composite materials.[70] In thin films, AFM topographic imaging assesses uniformity, roughness, and layer thickness, which are essential for applications in coatings and semiconductors, with typical resolutions achieving angstrom-level accuracy.[71] For graphene, AFM identifies defects like vacancies, grain boundaries, and wrinkles, which influence electrical and mechanical properties; for instance, studies have shown that defect densities can be mapped to correlate with reduced carrier mobility in graphene devices.[72]Beyond topography, AFM-derived techniques probe material properties at the nanoscale, particularly in tribology and electrical conductivity. Friction force microscopy (FFM), a lateral force variant of AFM, quantifies frictional forces between the tip and sample surface, providing insights into nanotribology of materials like thin films and nanoparticles; seminal work demonstrated atomic-scale friction variations on layered materials such as graphite, establishing FFM as a tool for studying wear and lubrication mechanisms in engineering surfaces.[73] Conducting AFM (c-AFM) measures local electrical conductivity by applying a bias voltage between a conductive tip and the sample, mapping variations in nanomaterials like graphene and carbon nanotubes; this has revealed conductive pathways in graphene with spatial resolution below 10 nm, aiding the design of nanoelectronics.[74]AFM also facilitates direct manipulation and patterning in nanotechnology, enabling precise control over nanostructures. Dip-pen nanolithography (DPN), where an AFM tip acts as a "pen" to deposit molecular "inks" onto substrates, allows creation of patterns with features down to approximately 15 nm, applied to fabricate nanoparticle arrays and functional thin films for sensors and devices.[75][76] Additionally, AFM tips can push and position individual nanoparticles or atoms on surfaces, as demonstrated in controlled relocation of gold nanoparticles with nanometer precision, which supports assembly of hybrid nanostructures for plasmonics. Examples include AFM-assisted alignment of carbon nanotubes, where tip-induced forces orient bundles for improved electrical connectivity in composites, and sizing of quantum dots, revealing size distributions that determine their optical emission properties.[77][78]Recent advancements highlight AFM's role in verifying the quality of 2D materials produced via mechanical exfoliation. For molybdenum disulfide (MoS2), AFM confirms monolayer thickness and assesses exfoliation uniformity by measuring step heights of approximately 0.7 nm, ensuring defect-free sheets for transistor applications; this verification has been crucial in scaling up production of large-area 2D crystals.[79]
Biological and biomedical uses
Atomic force microscopy (AFM) has become a vital tool in biology and biomedicine for visualizing and manipulating structures at the molecular and cellular levels in their native, hydrated environments. Operating in liquid media, AFM enables non-destructive imaging of dynamic biological processes, such as protein conformational changes and cellular interactions, without the need for labeling or vacuum conditions that could alter sample integrity.[80] This capability is particularly valuable for studying soft, fragile biomolecules that maintain their physiological hydration shells, contrasting with drier imaging techniques.[81]In biological imaging, AFM excels at resolving protein structures like DNA origami nanostructures in liquid environments, where it captures three-dimensional topologies with sub-nanometer precision. For instance, DNA origami fiducials have been used to correct tip artifacts and achieve accurate 3D reconstructions of these self-assembled scaffolds under buffer conditions.[82] Similarly, AFM in liquid mode images intact cells, revealing surface topographies, membrane features, and cytoskeletal elements in real-time, preserving their native state and functionality.[83]Force spectroscopy, a key AFM mode, measures mechanical forces at the single-molecule scale to probe biological interactions. In protein unfolding experiments, AFM applies controlled tensile forces, producing characteristic sawtooth patterns in force-extension curves that correspond to sequential domain unravels, providing insights into folding energy landscapes and mechanicalstability.[84] For cell adhesion, AFM quantifies binding forces between cells or cell-matrix components, such as integrins, revealing how mechanical cues influence migration and signaling in tissues.[83]Notable examples include AFM studies of amyloid fibrils, where it visualizes fibril assembly, polymorphism, and mechanical properties at the nanoscale, aiding understanding of neurodegenerative diseases like Alzheimer's.[85] Membrane proteins, embedded in lipid bilayers, have been imaged and unfolded in situ using AFM, exposing transmembrane topologies and interaction sites critical for drug targeting.[86] High-speed AFM, pioneered by Toshio Ando in the 2010s, captures these dynamics at video rates, filming processes like myosin walking on actin or bacteriorhodopsin conformational shifts in physiological solutions.[87]Challenges in biological AFM include managing hydration layers around samples, which can introduce repulsive forces and blur resolution, necessitating optimized liquid setups to minimize artifacts.[88] Thermal drift in liquids also distorts images over time, addressed through automated correction algorithms that align frames based on morphological features.[89] To enhance specificity, tips are functionalized with biomolecules like antibodies or ligands, enabling targeted recognition of cellular receptors while preserving imaging fidelity.[90]In biomedical applications, AFM characterizes drug delivery nanoparticles by mapping their surface topography, elasticity, and ligand distribution in liquid, ensuring stability and targeting efficiency for therapies like cancer treatment.[91] It also assesses tissuemechanics, quantifying stiffness variations in extracellular matrices to correlate biomechanical properties with disease progression, such as fibrosis or tumor invasion.[92]
Other interdisciplinary applications
In chemistry, atomic force microscopy (AFM) enables the detection and localization of single molecular recognition events by measuring specific binding forces between ligand-receptor pairs on self-assembled monolayers (SAMs). This approach has revealed unbinding forces on the order of 20–100 pN for interactions such as biotin-streptavidin, providing insights into molecular affinity and specificity at the nanoscale.[93] AFM imaging of SAMs, such as alkanethiol films on gold, has also characterized defect densities and phase segregation, with defect coverage typically below 5% in well-ordered monolayers, aiding the design of functional surfaces for sensors and coatings.[94]In physics, magnetic force microscopy (MFM), a derivative of AFM, visualizes magnetic domains in ferromagnetic materials by detecting stray magnetic fields from the tip, achieving resolutions down to 10–50 nm. MFM has mapped domain walls in thin films of materials like Co/Pt multilayers, revealing stripe patterns with widths of 100–200 nm influenced by film thickness and anisotropy.[95] Similarly, scanning thermal microscopy (SThM), integrated with AFM, probes local thermal properties by sensing temperature-dependent resistance changes in the tip, enabling nanoscale mapping of thermal conductivity in heterogeneous structures. For instance, SThM has quantified thermal conductivities varying from 0.1 to 10 W/m·K across polymer composites, highlighting interfaces as thermal bottlenecks.[96]In environmental science, AFM facilitates in situ observation of polymer degradation processes, such as the hydrolysis of poly(ε-caprolactone) films in aqueous media, where pit formation and surface roughening progress at rates of 1–10 nm/day under neutral conditions. This reveals degradation mechanisms like chain scission at hydrophilic domains, informing the environmental persistence of plastics.[97] AFM also characterizes soil nanoparticles, including natural clays and anthropogenic metal oxides, by measuring size distributions (typically 10–100 nm) and surface topography, which influence sorption behaviors and pollutant mobility in soils. Such analyses show irregular morphologies with root-mean-square roughness of 5–20 nm, correlating with enhanced reactivity toward contaminants like heavy metals.[98]Beyond these fields, AFM supports forensic trace analysis by quantifying adhesion forces of explosive particles, such as RDX, to surfaces, with pull-off forces ranging from 10–50 nN depending on substrate wettability, aiding in residue transfer modeling.[99] In art conservation, AFM maps pigment distributions in historical paints, identifying nanoscale binder-pigment interactions in oil paintings, where lead white particles (50–200 nm) exhibit distinct mechanical moduli of 1–5 GPa compared to organic matrices.[100]An emerging interdisciplinary application integrates AFM with tip-enhanced Raman spectroscopy (TERS), where the AFM tip amplifies Raman signals by factors up to 10^4–10^6, enabling chemical mapping at 1–10 nm resolution. TERS-AFM has identified molecular vibrations in SAMs and pigments, such as distinguishing phthalocyanine aggregates in artworks through characteristic peaks at 1300–1600 cm⁻¹, combining topography with spectroscopic identification for non-destructive analysis.[101]
Advantages and Limitations
Key advantages
One of the primary strengths of atomic force microscopy (AFM) is its versatility, enabling imaging and analysis of a wide range of samples regardless of conductivity, in diverse environments such as air, vacuum, or liquid. Unlike electron microscopy techniques that require conductive samples or high vacuum, AFM operates effectively on both conductive and non-conductive materials without such constraints.[102][103] This adaptability extends to biological specimens in their native aqueous environments, facilitating studies under physiologically relevant conditions.[104]AFM achieves high lateral resolution down to the atomic scale, often without the need for vacuum or sample labeling, providing detailed surface information that surpasses the limitations of optical methods. In ultra-high vacuum or liquid settings, resolutions approaching 0.1 nm have been demonstrated, allowing visualization of individual atoms or molecules on various substrates.[43] This capability stems from the probe's direct interaction with the sample surface, independent of electromagnetic radiation.[105]The technique's multifunctionality allows simultaneous acquisition of topographic data alongside mechanical, electrical, or other properties, enhancing its utility in comprehensive material characterization. For instance, conductive AFM tips can map surface topography while probing local conductivity or mechanical stiffness in a single scan.[106][107] This integrated approach provides multidimensional insights without requiring multiple instruments.AFM is inherently non-destructive, enabling real-time imaging of dynamic processes, and is generally more cost-effective than transmission electron microscopy (TEM) due to lower equipment and operational costs, absence of vacuum requirements, and minimal sample preparation.[108][104] High-speed variants further support in situ observations of surface evolution.[109]Quantitatively, AFM excels in measuring interaction forces, elastic moduli, and viscoelastic properties at the nanoscale through force spectroscopy and nanoindentation modes, yielding precise values such as Young's modulus for soft materials like cells or polymers.[110][111] These measurements, often calibrated against known standards, offer reliable mechanical characterization essential for nanotechnology and biomedicine.[112]
Principal disadvantages
One major limitation of atomic force microscopy (AFM) arises from artifacts introduced during imaging, particularly tip-induced damage and convolution effects. In contact and tapping modes, the probe tip can deform or damage soft samples, such as biological membranes or polymers, leading to inaccurate height measurements and altered surface morphology. For instance, excessive force application may displace lipid vesicles or cause wear on delicate nanostructures, resulting in spurious features in the topography. Additionally, the finite size and shape of the AFM tip cause convolution effects, where the imaged feature is broadened by the tip geometry, reducing lateral resolution to approximately the tip radius (typically 5–20 nm) and distorting sharp edges or narrow trenches. These artifacts are prevalent in high-resolution imaging and require careful interpretation, as detailed in image analysis techniques.The scanning speed of conventional AFM represents another significant drawback, often requiring several minutes to acquire a single image due to the mechanical constraints of the piezoelectric scanner and feedback loop. Typical scan rates are limited to 0.1–2 μm/s for atomic resolution, particularly in vacuum or liquid environments, where hydrodynamic damping further slows the process. This sluggish pace hinders real-time observation of dynamic processes, such as protein conformational changes or material phase transitions, although high-speed variants can achieve 10–20 frames per second over small areas (e.g., 240 × 120 nm).AFM is inherently restricted to surface characterization of small, flat samples, typically scanning areas up to 100–150 μm × 100–150 μm, with sample stages supporting dimensions up to 150–200 mm in diameter and heights up to 15–20 mm, depending on the system. It cannot probe internal structures or volumes, limiting its use to external topography and precluding applications like subsurface defect analysis common in techniques such as X-ray tomography. Moreover, samples must be mounted on flat, stable substrates to minimize tilt or curvature, which complicates preparation for irregular or macroscopic objects.Environmental factors pose substantial challenges for AFM operation, as the instrument is highly sensitive to external vibrations and temperature fluctuations. Even minor acoustic noise or building vibrations (e.g., from footsteps or HVAC systems) can introduce drift or noise in the cantilever deflection signal, degrading image quality and resolution. Temperature variations, as small as 0.1°C, affect the thermal expansion of the scanner and sample, causing positional inaccuracies up to several nanometers over extended scans.Quantitative measurements in AFM suffer from calibration variability, which undermines reproducibility and accuracy across instruments and users. Cantilever spring constants and sensitivity factors often deviate by factors of up to two from manufacturer specifications due to manufacturing inconsistencies and environmental influences, necessitating individual calibration via methods like thermal noise analysis. This variability affects force spectroscopy and mechanical property mapping, where errors in inverse optical lever sensitivity (InvOLS) can lead to 20–50% inaccuracies in modulus or adhesion values.
Comparisons with Other Techniques
Relation to scanning probe microscopies
Atomic force microscopy (AFM) belongs to the broader family of scanning probe microscopies (SPM), which achieve nanoscale resolution through raster scanning of a sharp probe across a sample surface while employing feedback loops to maintain a precise interaction distance or force.[113] These techniques typically resolve features down to the nanometer scale or better, relying on local probe-sample interactions rather than far-field optics or beams.[3] AFM exemplifies this approach by using a flexible cantilever with a tip to detect surface topography via mechanical deflections.[1]A primary distinction within SPM lies between AFM and scanning tunneling microscopy (STM), the inaugural technique in the family.[113] STM measures quantum tunneling currents between a conductive tip and sample, limiting its application to electrically conducting or semiconducting materials under vacuum conditions.[114] In contrast, AFM senses short-range repulsive or long-range attractive forces, such as van der Waals interactions, enabling imaging of non-conductive samples like insulators, polymers, and biological specimens without requiring electrical conductivity.[115] This mechanical detection principle allows AFM to operate in ambient air, liquids, or even physiological environments, broadening SPM's utility beyond STM's constraints.[2]Hybrids of AFM extend its capabilities by incorporating additional SPM modalities, such as conductive AFM (c-AFM), which combines topographic mapping with local electrical measurements.[116] In c-AFM, a conductive tip applies a bias voltage to the sample, enabling simultaneous acquisition of current maps and height profiles to study charge transport in thin films or nanostructures on insulating substrates.[117] This mode bridges AFM's mechanical sensitivity with STM-like electrical probing, facilitating analysis of device heterostructures where conductivity varies spatially.[118]AFM's development marked a pivotal evolution in SPM, addressing STM's sample limitations by enabling high-resolution imaging of insulators and operation in liquid media for dynamic processes like biomolecular interactions.[1] Introduced in 1986, it expanded the technique's scope to diverse materials and environments previously inaccessible to electron-based probes.[105]Beyond AFM and STM, the SPM family includes specialized variants like scanning near-field optical microscopy (SNOM), which integrates a probe with optical illumination to achieve sub-diffraction-limited resolution by exploiting evanescent fields near the sample surface.[119] Similarly, magnetic force microscopy (MFM) employs a magnetized AFM tip to detect stray magnetic fields, mapping domain structures in ferromagnetic materials with nanoscale precision.[120] These techniques underscore SPM's versatility in probing mechanical, electrical, optical, and magnetic properties through tailored tip-sample interactions.[121]
Differences from electron and optical microscopies
Atomic force microscopy (AFM) differs fundamentally from scanning electron microscopy (SEM) and transmission electron microscopy (TEM) in its operational principles and requirements. Unlike SEM and TEM, which rely on electron beams accelerated in a high-vacuum environment to image surfaces or internal structures, AFM uses a mechanical probe to raster-scan the sample surface, enabling operation in ambient air, liquids, or controlled gaseous atmospheres without the need for vacuum systems. This environmental flexibility allows AFM to image samples in near-native conditions, such as hydrated biological specimens, whereas SEM and TEM necessitate dehydration, conductive coating for non-conductive samples in SEM, or ultrathin sectioning (often <100 nm) for TEM, which can introduce artifacts or alter sample integrity.A key advantage of AFM over electron microscopies is its independence from sample conductivity and labeling; it measures nanoscale forces between the probe tip and surface atoms, providing true three-dimensional topographic data with vertical resolution down to 0.1 nm and lateral resolution approaching 0.1 nm under optimal conditions. In contrast, SEM offers surface morphology imaging with resolutions typically 1–10 nm but provides only a pseudo-3D appearance from secondary electron signals, while TEM achieves sub-angstrom resolution for internal structures but requires electron-transparent samples and cannot directly yield surface topography. However, electron microscopies excel in speed and field of view: SEM and TEM use parallel electron beams to image larger areas (up to micrometers) rapidly, whereas AFM's serial point-by-point scanning limits it to smaller scan areas (often 100 nm to 100 μm) and longer acquisition times, making it slower for broad surveys.AFM lacks the elemental analysis capabilities inherent to SEM when coupled with energy-dispersive X-ray spectroscopy (EDS), which identifies chemical composition by detecting characteristic X-rays from the sample; TEM can similarly integrate with electron energy loss spectroscopy (EELS) for elemental mapping. These analytical features make electron microscopies preferable for compositional studies, while AFM focuses on mechanical, frictional, and magnetic properties at the nanoscale. Despite these limitations, AFM and electron microscopies are often used complementarily in correlative approaches, such as overlaying AFM topography with SEM images to combine surface height data with high-resolution morphological and chemical details.Compared to optical microscopy, AFM circumvents the diffraction limit of visible light, which restricts resolution to approximately 200 nm, by directly probing surface contours rather than relying on photonscattering or absorption. This enables AFM to resolve features down to the atomic scale, providing quantitative height profiles and 3D reconstructions that optical methods, even super-resolution variants like STED or PALM, struggle to match for topographic accuracy without additional fluorescence labeling. Optical microscopy, however, offers advantages in speed, larger fields of view (from micrometers to millimeters), and multicolor imaging for labeling-based identification of cellular components or molecular species, features absent in AFM's monochromatic topographic outputs.The trade-offs in resolution stem from AFM's single-probe serial scanning versus the parallel illumination in optical systems, which allows rapid, wide-area imaging but at the cost of depth information and sub-diffraction precision. While optical microscopy operates in ambient conditions similar to AFM, it cannot measure mechanical properties like stiffness or adhesion directly, whereas AFM modes such as force spectroscopy provide such data. Correlative optical-AFM setups enhance this by integrating fluorescence or phase-contrast optical images with AFM topography, enabling multimodal analysis of dynamic processes in live samples.