Fact-checked by Grok 2 weeks ago

Scanning tunneling microscope

The scanning tunneling microscope (STM) is an instrument used to image conductive surfaces at the atomic scale by detecting the quantum tunneling current between a sharp metallic tip and the sample, enabling three-dimensional topographic mapping with sub-angstrom resolution. Invented in 1981 by physicists Gerd Binnig and Heinrich Rohrer at IBM's Zurich Research Laboratory, the STM marked a breakthrough in microscopy by overcoming the limitations of optical and electron microscopes in resolving atomic structures without damaging samples. Their development, which built on earlier concepts like field emission microscopy, earned Binnig and Rohrer the 1986 Nobel Prize in Physics, shared with Ernst Ruska for contributions to electron optics. At its core, the STM operates on the principle of quantum tunneling, where electrons from the sample or tip "tunnel" through the insulating vacuum gap—typically a few angstroms wide—under an applied bias voltage, producing a measurable current that decays exponentially with distance. This sensitivity allows the instrument to maintain a constant tunneling current via a feedback loop that adjusts the tip's height using piezoelectric actuators, scanning the tip raster-style across the surface to generate height profiles that reveal atomic arrangements. Key innovations included vibration isolation through suspension systems and precise tip sharpening, enabling stable operation in ambient or ultra-high vacuum conditions. The invention stemmed from efforts in the late 1970s to perform local spectroscopy on surfaces, with the first prototype achieving clear images of a gold crystal surface on March 16, 1981, after overcoming challenges like thermal drift and mechanical noise. Early demonstrations included resolving the 7×7 reconstruction of silicon(111) surfaces and the missing-row structure of gold(110), confirming theoretical models and opening avenues for studying adsorption, catalysis, and superconductivity. While primarily suited for conductive materials, extensions like combining STM with atomic force microscopy have broadened its applicability to insulators. The STM's impact extends to fields like nanotechnology, materials science, and biology, facilitating atomic manipulation—such as spelling "IBM" with xenon atoms in 1990—and advancing semiconductor design by visualizing self-assembled structures and molecular orientations. Its non-destructive, real-time imaging capabilities have driven innovations in microelectronics, surface engineering, and even biomolecular studies, such as DNA helices and viral components, underscoring its role as a foundational tool in modern nanoscience.

History and Development

Invention and Early Work

In the mid-20th century, quantum tunneling had been experimentally demonstrated in various contexts, notably through Ivar Giaever's 1960 work measuring energy gaps in superconductors via electron tunneling across thin insulating barriers, which laid foundational evidence for tunneling phenomena in solids. At the IBM Zurich Research Laboratory in the late 1970s and early 1980s, physicists Gerd Binnig and Heinrich Rohrer sought to overcome the resolution limits of existing surface analysis techniques, such as low-energy electron diffraction and field ion microscopy, which could not reliably image atomic-scale structures on surfaces. Their motivation stemmed from a desire to perform local spectroscopy on surface inhomogeneities smaller than 100 Å in diameter, building on earlier tunneling spectroscopy ideas to enable direct visualization of atomic arrangements. By 1981, they had shifted focus toward developing a scanning instrument that exploited vacuum tunneling between a sharp tip and a sample surface to map topography at unprecedented scales. The duo's first , constructed with assistance from Christoph Gerber, faced significant technical hurdles, particularly from mechanical vibrations and acoustic that could disrupt the delicate tunneling current. They addressed these by implementing a soft system within a , including superconducting levitation for isolation, allowing stable operation on March 16, 1981, when they first observed tunneling signals between a tip and a calcium iridium stannide sample. Progress accelerated in mid-1981 with initial topographic images of monosteps on CaIrSn₄ surfaces, and by spring 1982, they achieved atomic resolution on a gold (110) surface reconstruction. A landmark demonstration came in autumn 1982 with the first atomic-resolution image of the (111) surface, revealing its complex 7×7 reconstruction—a structure that resolved long-standing debates in . This success was detailed in their seminal 1983 publication in , marking the public debut of the scanning tunneling microscope () as a viable tool for atomic-scale surface imaging. Their earlier 1982 paper had already introduced the technique's principles through vacuum tunneling demonstrations on various surfaces.

Key Milestones and Recognition

The 1986 Nobel Prize in Physics was awarded to Gerd Binnig and Heinrich Rohrer for their invention of the scanning tunneling microscope, shared with Ernst Ruska for his work on electron microscopy. This recognition, just five years after the first STM images were obtained, underscored the instrument's transformative potential in surface science and propelled its adoption beyond IBM laboratories. Commercialization of STM accelerated in the mid-1980s, transitioning the technology from custom-built prototypes to accessible instruments for research labs worldwide. Early efforts included QuanScan, formed in 1986 by Caltech researchers based on their , which evolved into TopoMetrix by 1990 and facilitated broader distribution. By the late 1980s, companies like Digital Instruments began selling user-friendly air-compatible STMs in 1987, significantly expanding the base, while introduced its first STM systems in 1988. Park Scientific Instruments, founded in 1989, also contributed to early commercialization efforts alongside its focus on related probe technologies. In the late 1980s, advancements enabled low-temperature STMs, often operated in at cryogenic conditions, which were crucial for studying phenomena. These systems allowed precise measurements of superconducting gaps and states in high-temperature superconductors discovered in 1986, enhancing resolution and for delicate surface investigations. A landmark achievement came in 1990 when Donald Eigler and Erhard K. Schweizer at Almaden Research Center demonstrated the first controlled manipulation of individual atoms using an STM at low temperatures. By applying voltage pulses to position atoms on a surface, they spelled out "," proving the microscope's capability for atomic-scale engineering and opening avenues in nanoscience. Key conferences and publications further disseminated STM knowledge and fostered collaboration during this period. The First International Conference on Scanning Tunneling Microscopy, held in Santiago de Compostela, Spain, in July 1986, brought together pioneers to discuss applications and challenges, with proceedings capturing early progress. The subsequent Adriatico Research Conference on STM in Trieste, Italy, in 1987, emphasized theoretical and experimental advancements, while seminal reviews, such as Binnig and Rohrer's Nobel lecture, provided foundational overviews of the technique's evolution.

Theoretical Principles

Quantum Tunneling Fundamentals

Quantum tunneling arises from the wave-particle duality of matter, a fundamental principle of where particles such as electrons exhibit both particle-like and wave-like properties. This duality implies that the position and momentum of a particle cannot be precisely known simultaneously, leading to probabilistic descriptions of particle behavior. The provides the mathematical framework for this, describing how the wave function \psi(x) evolves and yields the probability density |\psi(x)|^2 of finding a particle at position x. Solving the time-independent reveals a non-zero probability for particles to penetrate regions classically forbidden, enabling tunneling. In classical mechanics, a particle with energy E cannot cross a potential barrier V(x) if V(x) > E, as it lacks sufficient kinetic energy, resulting in reflection and zero transmission probability. Quantum mechanically, however, the wave function extends into the barrier region, allowing a finite probability of transmission. This is governed by the time-independent Schrödinger equation: -\frac{\hbar^2}{2m} \frac{d^2\psi}{dx^2} + V(x)\psi = E\psi where \hbar is the reduced Planck's constant, m is the particle mass, and V(x) is the potential. Inside the barrier where V > E, the solution yields evanescent waves that decay exponentially rather than oscillate, characterized by \psi(x) \propto e^{-\kappa x} with decay constant \kappa = \sqrt{2m(V - E)/\hbar^2}. The transmission coefficient T, representing the tunneling probability, approximates to T \approx e^{-2\kappa d} for thick barriers of width d, highlighting the exponential sensitivity to barrier thickness and height. A simple example is the one-dimensional , where V(x) = V_0 for $0 < x < d and zero elsewhere. For an electron with E < V_0, the incident wave from the left region matches boundary conditions across the barrier, yielding an exact transmission coefficient involving hyperbolic functions, but the approximation T \approx 16 \frac{E}{V_0} (1 - \frac{E}{V_0}) e^{-2\kappa d} holds for \kappa d \gg 1. For instance, an electron of 7 eV incident on a 10 eV, 5 nm barrier has T \approx 3 \times 10^{-16}, but reducing d to 1 nm increases T to about $3 \times 10^{-3}, demonstrating how nanoscale dimensions enhance tunneling. In the context of electron transport between conductors separated by a vacuum gap, classical current flow is prevented because electrons lack the energy to overcome the and enter the vacuum, as their kinetic energy is insufficient to escape the potential well. Quantum tunneling circumvents this by allowing electrons to probabilistically cross the vacuum barrier when the gap is narrow, on the order of angstroms, without requiring classical emission.

Tunneling Current Models

The tunneling current in a (STM) is modeled using a rectangular potential barrier approximation, where the vacuum gap between the metallic tip and sample acts as an insulating barrier through which electrons tunnel quantum mechanically. In this model, electrons from the occupied states below the of one electrode tunnel to unoccupied states above the [Fermi level](/page/Fermi level) of the other electrode when a bias voltage is applied. This simplified approach assumes a one-dimensional potential barrier of height approximately equal to the average and width equal to the tip-sample separation. The tunneling current I in the rectangular barrier model is given by I \propto V e^{-2\kappa d}, where V is the applied bias voltage, d is the tip-sample separation, and \kappa is the decay constant characterizing the exponential decay of the electron wave function in the barrier. For small bias voltages, the current is approximately linear in V, reflecting the opening of an energy window for electron transmission proportional to the voltage. This expression arises from the transmission probability through the barrier, integrated over the available electron states. The decay constant \kappa depends on the work function \phi of the materials and is defined as \kappa = \sqrt{8\pi^2 m \phi / h^2}, where m is the electron mass and h is . More precisely, \kappa = \sqrt{2m(\phi - E)} / \hbar (with \hbar = h / 2\pi and E the electron energy relative to the barrier bottom), but for electrons near the , \phi dominates as the effective barrier height, typically around 4 eV for metals. Higher work functions increase \kappa, reducing the tunneling probability and thus the current for a given separation. In the context of two conductors, the tunneling current flows due to the alignment of their Fermi levels, which are shifted by the bias voltage V. Without bias, the Fermi levels align, and no net current flows as occupied states match unoccupied ones symmetrically. Applying a bias eV (where e is the electron charge) shifts the energy window, allowing electrons from occupied states in the lower-potential electrode (e.g., up to E_F) to tunnel into unoccupied states in the higher-potential electrode (from E_F + eV), with the current magnitude determined by the overlap of density of states in this window. This process is elastic, conserving electron energy during tunneling. The exponential dependence on distance in the model confers extreme sensitivity to the tip-sample separation d, with the current typically changing by a factor of 10 for every 1 Å variation in d. For a work function of 4 eV, \kappa \approx 1 \, \AA^{-1}, so a 1 Å increase in separation reduces the current by approximately an order of magnitude, enabling atomic-scale resolution in STM imaging. This sensitivity arises directly from the e^{-2\kappa d} term, where small changes in d amplify the decay dramatically. The rectangular barrier model's validity relies on assumptions such as low temperatures (e.g., below 100 K) to minimize thermal broadening of the Fermi distribution and small bias voltages (e.g., |eV| \ll \phi) to maintain a nearly rectangular barrier shape without significant band bending or image potential effects. At higher temperatures or biases, more advanced models incorporating trapezoidal barriers or surface states are needed for accuracy.

Bardeen's Perturbation Theory

John Bardeen developed a foundational quantum mechanical formalism for electron tunneling in 1961, applying time-dependent perturbation theory to calculate tunneling rates between two electrodes separated by a potential barrier. This approach treats the tunneling Hamiltonian as a perturbation on the unperturbed systems of the tip and sample, enabling the computation of transition probabilities without explicitly solving the full coupled wave equation across the barrier. The tunneling current in this framework is expressed using Fermi's golden rule as I = \frac{2\pi e}{\hbar} \sum_{k,p} |M_{kp}|^2 \delta(E_k - E_p - eV) [f(E_k) - f(E_p)], where the sum is over initial states k in one electrode and final states p in the other, M_{kp} is the tunneling matrix element, \delta enforces energy conservation, f is the , e is the electron charge, \hbar is the , and V is the applied bias voltage. The matrix element M_{kp} = \langle \psi_k | H' | \psi_p \rangle captures the coupling via the perturbation Hamiltonian H', but direct evaluation is avoided by separating the tip and sample wavefunctions \psi_k and \psi_p at a dividing surface in the vacuum barrier region. This separation transforms the volume integral into a surface integral over the flux of the wavefunctions, leveraging the divergence theorem: M_{kp} \approx -\frac{\hbar^2}{2m} \int_S \left( \psi_p^* \nabla \psi_k - \psi_k \nabla \psi_p^* \right) \cdot d\mathbf{S}, where m is the electron mass and S is the dividing surface. Integrating over the density of states (DOS) in each electrode further refines the current expression, yielding I \propto \int \rho_\text{tip}(E) \rho_\text{sample}(E + eV) |M(E)|^2 [f(E) - f(E + eV)] \, dE, where \rho denotes the DOS; this highlights how spatial and energetic variations in the sample DOS modulate the tunneling contrast in STM images. The theory assumes weak coupling and low bias, such that the matrix element remains approximately constant and electron-electron interactions are negligible. However, it breaks down under high bias voltages or strong tip-sample coupling, where inelastic processes or barrier distortion become significant, necessitating more advanced non-perturbative methods.

Instrumentation and Setup

Core Components

The core of a scanning tunneling microscope (STM) consists of several essential hardware elements designed to enable precise control and measurement at the atomic scale. The sharp metallic tip serves as the primary probe, typically fabricated from or to ensure conductivity and mechanical stability. Tungsten tips are commonly prepared by electrochemical etching in a solution such as , achieving atomic sharpness at the apex where a single atom often dominates the tunneling process. Platinum-iridium tips, favored for their resistance to oxidation in ambient conditions, are similarly etched using cyanide-based electrolytes or mechanically sheared for initial shaping. Piezoelectric scanners provide the fine x-y-z motion required for raster scanning the tip over the sample surface. These scanners are constructed from stacks or tubes of (PZT) ceramics, which expand or contract proportionally to applied voltage, offering sub-angstrom precision—typically on the order of 0.01 nm in the z-direction and up to several nanometers in x-y. In the original design, a tripod configuration of three piezoelectric elements allowed controlled displacements of about 1 angstrom per 0.1 V. The sample holder secures the specimen in close proximity to the tip, while integrated vibration isolation systems minimize external disturbances to maintain stability. Holders are often mounted on spring-suspended platforms, with eddy current damping achieved via copper plates interacting with permanent magnets to suppress oscillations effectively. This setup, as in early implementations, used stacked springs for multi-stage isolation to achieve displacements below 0.001 nm. Supporting electronics include a bias voltage supply and a high-sensitivity current amplifier. The bias voltage, applied between the tip and sample, ranges from millivolts to several volts to induce the tunneling current, which depends exponentially on the tip-sample distance. Current amplifiers, often transimpedance types, detect signals from femtoamperes to nanoamperes with low noise, using feedback resistors in the gigaohm range for logarithmic amplification. These components collectively enable the STM's hallmark resolution, with typical lateral precision below 0.1 nm and vertical sensitivity under 0.01 nm, allowing atomic-scale imaging.

Environmental Considerations

The operation of a scanning tunneling microscope (STM) demands stringent environmental controls to achieve atomic-scale resolution, primarily through ultra-high vacuum (UHV) conditions with base pressures below $10^{-10} Torr. This level of vacuum prevents adsorption of residual gases onto the sample surface, which could otherwise lead to contamination layers that distort the tunneling current and degrade image quality. Oxidation of reactive surfaces, such as metals or semiconductors, is similarly averted, ensuring stable and reproducible measurements of clean atomic structures. Low-temperature environments, typically ranging from 4 K to 300 K, are implemented using cryostats or dilution refrigerators to mitigate thermal effects that compromise STM performance. At these reduced temperatures, thermal drift—arising from differential expansion or contraction of components—is minimized, allowing for precise tip positioning over extended scan times. Additionally, phonon excitations, which contribute to inelastic scattering and broaden spectroscopic features, are suppressed, enabling sharper resolution of electronic states near the Fermi level. Acoustic and mechanical isolation is essential to shield the instrument from external vibrations, which could otherwise introduce noise exceeding the angstrom-level precision required for tunneling. Common setups employ heavy granite tables mounted on pneumatic air tables or suspended via bungee cords, providing multi-stage damping with cutoff frequencies below 3 Hz. Superconducting magnets or eddy-current dampers may further isolate magnetic and electromagnetic interference in cryogenic systems. STM hardware is frequently assembled in cleanrooms (Class 100 or better) to eliminate dust and particulates that might adhere to sensitive components, with final integration occurring under UHV. In-situ tip preparation is critical for maintaining apex sharpness and conductivity; techniques include controlled field emission—where high bias voltages gently etch the tip—or thermal annealing at 800–1200 K to desorb contaminants without altering the underlying structure. These methods ensure the tip ends in a single-atom configuration, vital for reliable tunneling. A key trade-off in STM environments involves UHV's restriction to conductive or semiconductive samples, as insulators often require surface charging mitigation unavailable in such vacuums; ambient-pressure STMs, while enabling broader material compatibility including coated insulators, suffer from rapid contamination and reduced resolution compared to UHV setups. This clean vacuum is foundational to the reliable observation of quantum tunneling effects, as detailed in current models.

Operational Procedure

Scanning and Feedback Mechanisms

The scanning process in a scanning tunneling microscope (STM) employs a raster scan pattern, where the sharp metallic tip is moved across the sample surface in a systematic, line-by-line manner over a predefined area. This motion occurs in the x-y plane and is controlled by applying precise voltages to piezoelectric actuators, which expand or contract in response to the electric field, enabling sub-angstrom resolution positioning. The tip typically scans from left to right along each line, then advances to the next line, forming a grid of data points that map the surface topography. To maintain a stable tip-sample interaction, a feedback loop continuously monitors the tunneling current and adjusts the tip height accordingly. This loop utilizes a proportional-integral-derivative (PID) controller, which processes the difference between the measured current and a setpoint (typically on the order of nanoamperes) to generate corrective signals for the z-direction . The PID algorithm minimizes errors in current by balancing proportional response to the immediate deviation, integral accumulation of past errors, and derivative anticipation of future changes, ensuring rapid and stable adjustments without oscillations. In constant current mode, the dominant operational approach for topographic imaging, the feedback loop dynamically varies the z-position of the tip to compensate for surface undulations, keeping the tunneling current fixed at the setpoint. As the tip scans, these vertical adjustments are recorded as a function of x-y position, yielding a height map that reflects the sample's surface profile; this mode is particularly effective for following atomic-scale topography due to the exponential dependence of tunneling current on tip-sample distance. Prior to scanning, the tip must be approached to the sample surface through a two-stage procedure to establish tunneling contact without causing damage. The coarse approach brings the tip from an initial distance of several millimeters to micrometers using mechanical actuators, such as micrometers or inchworm motors, which provide larger displacements at slower speeds (e.g., 2 µm/s) while monitoring for initial current onset. Once proximate, the fine approach engages the piezoelectric elements for sub-micrometer to nanometer precision, incrementally closing the gap (from ~1 mm to ~1 nm) via voltage ramps until the tunneling current reaches the setpoint, with safeguards like current thresholds to prevent crashes. During imaging, data acquisition can include spectroscopic measurements by pausing the raster scan at selected points to perform voltage ramps. At these pauses, the bias voltage is swept across a range (e.g., ±1 V) while measuring the resulting current, generating current-voltage (I-V) curves that reveal local electronic properties; the feedback loop is often disabled momentarily to isolate the spectroscopic signal.

Imaging Modes

The (STM) operates in various imaging modes that extend its capabilities beyond basic topographic mapping, allowing the acquisition of spectroscopic, magnetic, and dynamic data about surface properties. These modes modify the standard scanning procedure—typically constant current topography with feedback control—to prioritize specific measurements, such as electronic structure or temporal evolution, while maintaining atomic-scale resolution under ultrahigh vacuum conditions. In constant height mode, the tip height remains fixed during lateral raster scanning, without active feedback adjustment, enabling rapid imaging by directly recording variations in tunneling current as a function of lateral position and bias voltage. This approach maps both surface topography and local electronic density of states through current intensity changes, which reflect tip-sample distance and orbital features like defects or molecular states. It is particularly advantageous for atomically flat samples, where the absence of z-piezo feedback minimizes mechanical delays and supports scan speeds up to several lines per second, though it risks tip crashes on rough terrains. Spectroscopic modes, often implemented as scanning tunneling spectroscopy (STS), focus on the electronic structure by mapping the differential conductance \frac{dI}{dV}, which is proportional to the local density of states (LDOS) at the . A lock-in amplifier facilitates this by superimposing a small AC modulation (typically 1-10 mV at 100 Hz to several kHz) on the DC bias voltage, while the feedback loop is opened or paused; the amplifier then extracts the first-harmonic current response at the modulation frequency to yield \frac{dI}{dV} maps with sub-meV energy resolution. These maps reveal spatial variations in electronic states, such as band gaps or orbital hybridizations, across the scanned area, with acquisition times on the order of minutes for 100x100 pixel images at low temperatures. Bias spectroscopy complements spatial mapping by acquiring current-voltage (I-V) curves at discrete, fixed tip positions, where the bias is ramped (e.g., from -2 V to +2 V) while holding the tip stationary and disabling feedback to measure steady-state tunneling current. These curves directly probe local energy levels, identifying features like molecular orbitals or superconducting gaps through peaks and onsets in conductance that correspond to the sample's density of states integrated over the bias window. Performed under cryogenic conditions to sharpen the , this mode achieves energy resolutions below 1 meV and is essential for site-specific analysis of quantum confinement or doping effects. Spin-polarized STM (SP-STM) adapts the standard setup with a ferromagnetic or antiferromagnetic tip to enable magnetic imaging, where the spin-dependent tunneling current varies based on the relative alignment of tip and sample spin polarizations. Magnetic tips, often prepared by coating etched tungsten with iron or chromium films, provide a spin-polarized electron source or drain, producing contrast in topographic or spectroscopic images that reflects magnetic domains or spin textures down to single-atom scale. The tunneling magnetoresistance effect underlies the signal, with current modulations up to 20-50% for well-polarized tips (polarization >30%) on magnetic surfaces like Fe or Mn films. Pulse techniques introduce time resolution to STM by applying ultrafast voltage or optical pulses to excite the tip-sample , capturing transient in tunneling current with to precision. In terahertz-STM, for instance, a THz transiently biases the (peak fields ~1 MV/cm), inducing non-equilibrium measurable via time-gated current detection, while enable pump-probe schemes for studying vibrational or relaxations. These methods preserve atomic but require specialized setups like photoconductive antennas or streak cameras, achieving temporal resolutions as low as 30 for processes like hot- decay in adsorbates.

Applications and Capabilities

Surface Atomic Imaging

The scanning tunneling microscope (STM) enables direct visualization of atomic-scale surface structures through measurement of the tunneling current between a sharp tip and the sample surface, achieving lateral resolutions down to 0.1 nm and vertical resolutions of 0.01 nm. This capability has revolutionized by providing real-space images of atomic lattices that were previously inferred only from reciprocal-space techniques like . A demonstration occurred in 1983, when STM resolved the complex 7×7 reconstruction of the Si(111) surface, revealing a containing 49 silicon atoms arranged in a dimer-adatom-stacking-fault , confirming theoretical models and highlighting STM's role in resolving long-standing structural debates. STM images do not strictly map topographic height but rather the local density of states (LDOS) near the , leading to variations that reflect electronic structure rather than purely geometric features. According to the Tersoff-Hamann theory, the tunneling current is proportional to the integral of the sample's LDOS over an energy window around the voltage, causing atoms with higher LDOS—such as those with partially filled d-orbitals or dangling bonds—to appear brighter, while those with lower LDOS appear . For instance, on surfaces like (111), individual atoms exhibit uniform brightness due to delocalized s-p states, whereas on semiconductors like (100), reconstructed dimer rows show alternating from asymmetric charge distribution. Interpreting STM images requires distinguishing real-space topography from reciprocal-space influences, as Fourier transforms of images can reveal periodic modulations akin to diffraction patterns. Real-space imaging directly captures local atomic arrangements and defects, offering intuitive visualization, whereas reciprocal-space analysis via fast Fourier transform uncovers long-range order and moiré superstructures, bridging STM with traditional diffraction methods. This dual perspective has been crucial for studying surface reconstructions, such as the herringbone pattern on Au(111), where STM reveals stress-induced uniaxial distortions in real space, complemented by spot-splitting in reciprocal space. Extensive STM investigations have elucidated adsorbates, defects, and reconstructions on both metal and surfaces, providing atomic-level insights into surface chemistry and physics. On semiconductors like Ge(001), adsorbates such as or alkali metals induce local reconstructions, with STM visualizing end-bond and bridge configurations that alter the 2×1 dimer rows. Defects, including vacancies and steps on metal surfaces like Pt(111), appear as dark spots or lines due to reduced LDOS, influencing catalytic activity by trapping impurities. Reconstructions on semiconductors, such as the √3×√3 phase of Bi on Si(111), show ordered adsorbate islands with hexagonal symmetry, while on metals like Ir(111), oxygen-induced faceting creates nanoscale pyramids observable in real space. A prominent example is the imaging of , where routinely resolves the honeycomb lattice with carbon-carbon bond lengths of 0.142 nm, revealing sublattice contrast from π-orbital LDOS asymmetry. Early high-resolution studies on exfoliated sheets supported on insulating substrates confirmed the Bernal stacking in bilayer regions through moiré patterns, demonstrating 's sensitivity to van der Waals interfaces. Recent applications as of include atomic-scale mapping of MXene surfaces, such as Ti3C2Tx, revealing termination groups and electronic properties for and . In biological contexts, has been applied to image DNA helices and viral components on conductive substrates, though challenges with non-conductive samples often require hybridization with ; a 2023 study visualized functionalized gold nanoparticles for biomedical targeting.

Atomic Manipulation

Atomic manipulation with the (STM) involves the controlled repositioning of individual atoms or molecules on a surface by exploiting interactions between the STM tip and the adsorbate. This technique enables the construction of custom atomic-scale structures, leveraging the STM's atomic precision to apply forces that displace adatoms without causing or damage to the . The process typically occurs under and low temperatures to minimize perturbations, allowing for deliberate atom-by-atom . Three primary modes of lateral manipulation—pulling, pushing, and sliding—govern the movement of atoms based on the tip-sample interaction and geometry. In the pulling mode, the atom adheres to the receding tip via a short-range attractive force, following it until the interaction weakens, resulting in discontinuous jumps observable in tip-height traces. Pushing occurs through repulsive interactions as the tip advances toward the atom, causing it to roll or slide ahead continuously. Sliding involves a smooth, long-range attraction where the atom glides laterally with the tip at a constant distance, often distinguished by gradual changes in tunneling current during manipulation. These modes are identified by analyzing approach and retraction curves of the tip height relative to the surface, which reveal distinct thresholds for atom displacement. A landmark demonstration of atomic was achieved in , when researchers positioned 35 (Xe) atoms on a (Ni(110)) surface at 4 K to spell out "" in a 5 nm × 14 nm arrangement, showcasing sub-angstrom precision in atom placement. The atoms were individually nudged by bringing the STM tip into close proximity (~4-6 ) and applying a voltage of approximately 0.1 V, allowing van der Waals forces to guide the displacement while maintaining reversible control. This experiment highlighted the STM's capability for deliberate atomic engineering, with each Xe atom manipulated sequentially over hours without unintended movement. The forces arise from van der Waals and short-range chemical between the and the adatom, tunable by tip-sample separation and voltage. Van der Waals forces dominate in sliding modes at larger separations (>5 ), providing gentle lateral drag, while chemical —forming transient covalent-like interactions—enables pulling at closer distances (~3-4 ), with typical voltages of 0.1-1 V to balance attraction and avoid inelastic . These interactions must be calibrated to prevent irreversible processes, such as adatom pickup by the or substrate . Manipulation thresholds distinguish reversible from irreversible outcomes: below a critical tip height (e.g., ~2-3 for many systems), elastic interactions allow the atom to return to its original site upon tip retraction, enabling iterative positioning. Exceeding this threshold triggers irreversible events, like atom to the or excitation-induced , often detected by sudden jumps in manipulation curves. Such has facilitated applications in fabrication, including the creation of quantum corrals—circular arrays of 48 iron atoms on Cu(111) that confine surface electrons, producing observable patterns via quantum . These corrals, approximately 14 nm in diameter, demonstrate how atomic can engineer electronic properties for potential quantum devices. Recent advances as of 2025 include precise of atoms and molecules in two-dimensional materials like and dichalcogenides, enabling the fabrication of custom quantum structures.

Spectroscopic Techniques

Scanning tunneling spectroscopy (STS) extends the capabilities of the scanning tunneling microscope by measuring the differential conductance, \frac{dI}{dV}, as a function of bias voltage V, which is proportional to the local density of states (LDOS) at the Fermi level of the sample surface. This technique allows probing of electronic properties with atomic-scale spatial resolution and energy resolution down to approximately 1 meV. In practice, STS is performed by pausing the scan at selected positions and ramping the bias voltage while recording the tunneling current, often in constant-height or constant-current modes to capture spectroscopic data. Inelastic electron tunneling spectroscopy (IETS), a variant of STS, detects vibrational modes by observing steps or peaks in the \frac{d^2I}{dV^2} spectra corresponding to phonon energies on the order of millielectronvolts (meV). These features arise when tunneling electrons excite molecular or lattice vibrations, opening additional inelastic channels that enhance conductance at specific energies. IETS has been particularly useful for identifying bond-specific vibrations in adsorbed molecules, providing insights into adsorbate-substrate interactions with sub-nanometer spatial resolution. STS enables detailed mapping of band structures and gap states in semiconductors by revealing spatial variations in the LDOS across the bandgap. For instance, conductance peaks near band edges highlight defect-induced states, while the absence of states within the gap confirms insulating behavior, as observed in materials like or surfaces. In applications to , STS has mapped the superconducting energy gap in cuprates such as Bi₂Sr₂CaCu₂O₈₊δ, revealing d-wave with gap magnitudes around 20-30 meV and spatial inhomogeneities linked to pseudogap phenomena. Similarly, in organic molecules, STS visualizes s by correlating \frac{[dI](/page/Di)}{[dV](/page/DV)} peaks with highest occupied (HOMO) and lowest unoccupied molecular orbital (LUMO) energies, as demonstrated in studies of porphyrins and fullerenes on metal substrates, where orbital hybridization shifts are resolved to ~0.1 nm spatially. As of , extensions like electron resonance scanning tunneling microscopy (ESR-STM) have advanced the probing of spin states in individual atoms and molecules, supporting the development of atomic-scale quantum platforms and qubits.

Limitations and Advancements

Technical Challenges

One major technical challenge in (STM) arises from artifacts, particularly when the probe exhibits multiple apexes or mini-tips due to imperfect preparation or wear. These irregularities can produce ghost images or distorted features in the scan, as the tunneling current interacts with several protruding points simultaneously rather than a single atomic-scale apex, thereby degrading to the nanometer scale instead of precision. For instance, a blunt with clustered mini-tips leads to replicated patterns in the , complicating accurate surface interpretation. Thermal drift and further limit STM performance, especially at , where relative motion between the tip and sample—driven by thermal gradients and expansions in the —restricts stable imaging to scan times of only a few minutes. This drift, often on the order of angstroms per second, blurs atomic features and necessitates frequent recalibration, while and from exacerbates signal instability, reducing the effective in ambient or non-optimized setups. Such issues are mitigated somewhat in environments, but they remain inherent constraints for routine operation. STM's reliance on quantum tunneling imposes strict sample requirements, confining imaging to electrically conductive or semiconductive surfaces where electrons can flow between tip and substrate. Insulating materials, such as oxides or polymers, cannot sustain the necessary tunneling current without modification, typically requiring a thin conductive like or a conductive to enable , though this adds preparation complexity and potential artifacts from the overlayer. Contamination from adsorbates, such as hydrocarbons or water molecules, poses another challenge by altering local tunneling paths and electron density, often manifesting as unexpected protrusions or depressions in images that mimic surface features. These unintended layers, common in non-vacuum conditions, can self-assemble into organized structures on substrates like graphene, interfering with the probe's ability to resolve true topography and requiring pristine preparation to avoid. Finally, speed limitations stem from the feedback system's , typically around 1-10 kHz for standard STMs, which constrains the rate of tip adjustments to maintain or height during scanning. As a result, acquiring a full image—such as a 100 × 100 area—takes from seconds for coarse scans to hours for high-resolution atomic mapping, bottlenecked by the need for precise, iterative corrections to avoid crashes or distortions.

Modern Improvements and Variants

Since the early , advancements in () have addressed key limitations in , environmental compatibility, and operational speed, enabling the study of dynamic processes at the atomic scale. , developed around 2010, integrates pulses with traditional setups to probe dynamics on to timescales. This technique employs pump-probe schemes where an optical pump excites the sample, and the tip detects transient tunneling currents, achieving spatiotemporal resolutions down to 1 and 100 . Recent extensions include lightwave-driven () in 2024, which provides real-space access to single-molecule dynamics on scales. For instance, coherent of tunneling in a junction has been demonstrated using carrier-envelope phase-stable pulses, revealing attosecond-scale manipulations of . Electrochemical (EC-) represents for in environments, overcoming the constraints of conventional to investigate electrochemical interfaces. Operating in electrolytes, EC- maintains while applying controlled potentials, allowing of surface reconstructions during . This pivotal in studying processes, such as the selective of alloys, where -scale pitting and passivation layers are visualized. In , EC- elucidates electrode-electrolyte interactions, including solid-electrolyte formation on lithium-ion anodes, providing insights into degradation mechanisms at the nanoscale. These capabilities stem from specialized electrochemical cells that isolate the from faradaic currents, ensuring tunneling. Efforts to achieve video-rate imaging have focused on enhancing piezoelectric actuators and feedback systems, enabling scan rates exceeding 100 frames per second while preserving atomic resolution. Advanced piezotube designs with higher resonant frequencies, often exceeding 10 kHz, minimize mechanical resonances and drift, allowing continuous of surface diffusion and reactions. For example, ultrahigh-vacuum STMs with optimized shear-mode piezos have captured real-time videos of metal surfaces at speeds up to 10^5 nm/s tip velocity. Microelectromechanical system (MEMS)-based scanners further boost Z-axis bandwidth to over 100 kHz, facilitating high-speed electrochemical variants for dynamic processes in liquids. Integration of STM with complementary techniques has expanded its applicability in complex environments. Combined STM-scanning electron microscopy (SEM) systems allow correlative imaging, where SEM provides overview topography and STM delivers atomic detail, useful for nanorobotics and in situ manipulation. Similarly, STMs operable in high magnetic fields as high as 35 T (as of 2025) enable the study of spin-polarized states and quantum materials, with custom cryogenic designs maintaining stability against field-induced vibrations. Recent enhancements, such as subsurface atomic structure imaging demonstrated in 2025, allow visualization of buried interfaces in materials. Adaptations of qPlus sensors, originally developed for non-contact atomic force microscopy, have been incorporated into STM setups to enhance sensitivity in hybrid modes. These quartz-based resonators, with quality factors over 10,000, allow simultaneous tunneling and frequency-shift detection, improving for weakly interacting surfaces without lateral forces. Such integrations are particularly effective in electrochemical or low-temperature environments, where qPlus provides robust for non-contact enhancements.

Atomic Force Microscopy

The atomic force microscope (AFM) was invented in 1986 by , Christoph Gerber, and Calvin F. Quate as a means to extend scanning probe imaging to non-conductive materials, building on the principles of the scanning tunneling microscope (STM) but replacing electronic tunneling with detection. In AFM, a sharp tip mounted on a flexible is raster-scanned over the sample surface, and the cantilever's deflection—caused by short-range interatomic s between the tip and sample—is precisely measured using an optical beam bounce technique, enabling topographic mapping with high sensitivity. This -sensing approach allows AFM to image insulators and other non-conducting surfaces where STM's requirement for electron tunneling fails, achieving resolutions down to approximately 10 piconewtons (pN). AFM operates in several key modes to optimize imaging for different sample types and conditions, each exploiting variations in tip-sample interaction forces. In contact mode, the tip maintains constant light contact with the surface, measuring repulsive forces as the cantilever bends, which is suitable for rigid samples but can cause wear on soft materials. Non-contact mode involves oscillating the cantilever far above the surface to detect attractive van der Waals forces, minimizing damage and enabling imaging in vacuum or gaseous environments. Tapping mode, a dynamic variant, oscillates the cantilever at its resonance frequency to intermittently "tap" the surface, balancing resolution and gentleness for delicate samples like biological structures. Unlike STM's electronic sensing of tunneling currents, which is limited to conductive or semiconducting samples, AFM's mechanical sensing of atomic-scale forces provides atomic resolution on insulators, such as imaging individual atoms on mica or silicon oxide surfaces. This capability complements STM by enabling studies of diverse materials, including biomolecules like DNA and proteins, as well as polymers, where AFM reveals nanoscale topography, elasticity, and adhesion without conductivity constraints. For instance, AFM has mapped the helical structure of collagen fibers in polymers and force-induced unfolding of protein domains in piconewton ranges. Hybrid STM-AFM systems integrate both techniques on a shared piezoelectric , allowing simultaneous acquisition of (tunneling ) and mechanical () for comprehensive sample , such as probing charge states in nanoparticles or subsurface defects in semiconductors. These combined instruments enhance correlative analysis, revealing how local properties influence mechanical behavior at the .

Other Probe-Based Methods

Magnetic force microscopy (MFM) is a scanning probe technique that images magnetic domains and stray fields on sample surfaces by detecting long-range magnetic interactions between a magnetized tip and the sample. Developed in , MFM operates in a non-contact mode where the tip, coated with a ferromagnetic material, oscillates above the surface, and variations in magnetic gradient cause shifts in the cantilever's resonance frequency. This allows spatial resolution down to approximately 10-50 nm for magnetic features, making it valuable for studying magnetic recording media and nanomaterials. Unlike STM, which relies on quantum tunneling of electrons, MFM probes magnetic dipole interactions without requiring electrical conductivity in the sample. Scanning near-field optical microscopy (SNOM), also known as near-field scanning optical microscopy (NSOM), achieves optical with resolution below the limit of , typically 20-100 , by using a sub-wavelength or tip to confine in the near-field zone. Pioneered in , the technique illuminates the sample through a nanoscale , such as a pulled optical fiber coated with metal, and collects evanescent waves that decay rapidly with distance. SNOM enables spectroscopic analysis, including and Raman mapping, on non-conductive samples like biological tissues or polymers, providing chemical contrast complementary to topographic data. In contrast to STM's electronic sensitivity, SNOM exploits optical interactions for label-free of dielectric properties. Kelvin probe force microscopy (KPFM) maps variations in surface potential and work function at the nanoscale by nullifying the electrostatic force between a conductive tip and sample through an applied bias voltage. Introduced in 1991, KPFM typically employs a two-pass scanning mode: the first pass acquires topography via , and the second pass measures contact potential difference (CPD) while lifting the tip to minimize van der Waals influences. This non-destructive achieves resolutions of 10-20 and is widely used for characterizing semiconductors, , and processes. KPFM differs from by focusing on capacitive electrostatic forces rather than tunneling currents, allowing measurements on insulators. Ballistic electron emission microscopy (BEEM) extends STM principles to probe subsurface interfaces in layered structures by injecting hot s from a scanning tip and detecting those that traverse the sample ballistically to reach a buried collector . First demonstrated in 1988, BEEM maps heights and interface quality with nanometer lateral resolution and energy selectivity up to 0.1 eV. It is particularly suited for studying metal-semiconductor junctions in devices like transistors, revealing inaccessible to surface-only techniques. While sharing STM's injection , BEEM incorporates a three-terminal setup to isolate ballistic currents from scattered ones. These probe-based methods, including MFM, SNOM, KPFM, and BEEM, all employ raster scanning over the sample surface similar to STM's imaging procedure, but they diverge in the physical interactions exploited—MFM via magnetic forces, SNOM through near-field , KPFM by electrostatic potentials, and BEEM with ballistic —enabling diverse characterizations beyond purely tunneling.

References

  1. [1]
    [PDF] Gerd Binnig and Heinrich Rohrer - Nobel Lecture
    We present here the historic development of Scanning Tunneling Microscopy; the physical and technical aspects have already been covered in a few recent.
  2. [2]
    September 1981: Invention of the scanning tunneling microscope
    The first researchers to succeed in building an STM were Gerd Binnig and Heinrich Rohrer at IBM Research Laboratories in Zurich, Switzerland, largely because of ...
  3. [3]
    Scanning tunneling microscope - IBM
    Binnig and Rohrer's first STM experiment involved inspecting the surface structure of a gold crystal. The STM provided unsurpassed spatial resolution, ...Missing: principles | Show results with:principles
  4. [4]
    [PDF] The Scanning Tunneling Microscope - Physics 122
    The tunneling current depends both on the tunnel distance and the electronic structure of the surface and on the fact that each atomic element has an ...<|control11|><|separator|>
  5. [5]
    The Nobel Prize in Physics 1986 - NobelPrize.org
    The Nobel Prize in Physics 1986 was divided, one half awarded to Ernst Ruska ... Binnig and Heinrich Rohrer "for their design of the scanning tunneling microscope ...
  6. [6]
    Press release: The 1986 Nobel Prize in Physics - NobelPrize.org
    The other half of this year's prize has been awarded to Gerd Binnig and Heinrich Rohrer for “their design of the scanning tunneling microscope”.
  7. [7]
    Scanning Probe Microscopy - Engineering and Technology History ...
    May 19, 2015 · Demographic and Policy Shifts. The late 1980s also saw the founding of commercial probe microscope manufacturers, especially in the US West and ...
  8. [8]
    About Us - Scienta Omicron
    Two years after the Nobel Prize award, Omicron started to offer the first scanning tunneling microscope. ... Both companies started their operations early 1980s.Missing: commercialization | Show results with:commercialization
  9. [9]
    History - Park Systems
    Park upon his graduation took this technology out of the lab and founded Park Scientific Instruments (PSI), where it pioneered the commercialization of AFM.
  10. [10]
    Scanning tunneling spectroscopy of high-temperature ...
    Mar 13, 2007 · Tunneling spectroscopy has played a central role in the experimental verification of the microscopic theory of superconductivity in classical superconductors.
  11. [11]
    7.6 The Quantum Tunneling of Particles through Potential Barriers
    Sep 29, 2016 · We substitute Equation 7.79 into Equation 7.71 and obtain the exact expression for the transmission coefficient for the barrier,. T ( L , E ) ...
  12. [12]
    [PDF] Ivar Giaever - Nobel Lecture
    This was one of the first clues that we were dealing with tunneling rather than shorts. In the early experiments we used a relatively thick oxide, thus very ...
  13. [13]
  14. [14]
  15. [15]
    Single-electron tunneling force spectroscopy of an individual ...
    Jun 28, 2006 · Since the tunneling probability is a strong function of z ⁠, varying by almost an order of magnitude per angstrom, an electron may be deposited ...
  16. [16]
    Tunnelling from a Many-Particle Point of View | Phys. Rev. Lett.
    Ivar Giaever, Phys. Rev. Letters 5, 147, 464 (1960); J. Nicol, S. Shapiro, and P. H. Smith, Phys. Rev. Letters 5, 461 (1960); J. Bardeen, L. N. Cooper, ...Missing: John | Show results with:John
  17. [17]
    [PDF] Scanning Tunneling Microscope (STM) (II)
    Scanning Tunneling Microscopy (STM) (II). Instrumentation: The following figure shows essential elements of STM. A probe tip, usually made of W or Pt-Ir.Missing: core | Show results with:core
  18. [18]
    Short CommunicationThe electrochemical etching of tungsten STM ...
    STM tips were prepared from electrochemically etched, high-purity tungsten wire (0.25 mm diameter, 99.95% purity; Goodfellow, Cambridge, U.K.). Sodium hydroxide ...Missing: microscope | Show results with:microscope
  19. [19]
    [PDF] Pt-Ir Tip Etching Techniques for Scanning Tunneling Microscopy
    A microscope stage was modified to hold a copper clip to hold the tip in place, a platform to mount the tungsten loop and an additional two-way adjustable stage ...
  20. [20]
  21. [21]
    [PDF] Lecture 6 Scanning Tunneling Microscopy (STM)
    This (approximate) equation shows that the tunneling current obeys Ohm's law, i.e. the current I is proportional to the voltage V. • The current depends ...Missing: rectangular exp kappa
  22. [22]
    A high-stability scanning tunneling microscope achieved by an ...
    Nov 14, 2013 · It is equipped with a small but powerful GeckoDrive piezoelectric motor which drives a miniature and detachable scanning part to implement ...
  23. [23]
    Fast low-noise transimpedance amplifier for scanning tunneling ...
    Jul 1, 2020 · The usual bias voltage range of an STM is ±10 V, which is higher than the permissible supply voltage of the first-stage OP1 (6 V between the ...
  24. [24]
    FEMTO® - DLPCA-200 switchable transimpedance amplifier
    The DLPCA-200 offers variable transimpedance gain from 10 3 to 10 11 V/A allowing the sensitive measurement of currents in the sub-pico to milli amp range.
  25. [25]
    [PDF] Scanning Tunneling Microscopy S - Rutgers University, GLASS |
    For this reason, the core of a STM, where the tip-sample junction is located, is always equipped with one or several types of vibration damping systems. These ...
  26. [26]
    [PDF] a practical guide to scanning probe microscopy - MIT
    In the early 1980's scanning probe microscopes (SPMs) dazzled the world with the first real-space images of the surface of silicon. Now, SPMs are used in a.Missing: commercialization Omicron
  27. [27]
    Designing Advanced Scanning Probe Microscopy Instruments | NIST
    Sep 22, 2011 · The progress to lower temperatures has been slow because of the competing requirements of isolation from the environment for low noise ...
  28. [28]
    [PDF] A Liquid-helium-free High-stability Cryogenic Scanning Tunneling ...
    Thermal drift analysis showed that under optimized conditions, the lateral stability of the STM scanner can be as low as 0.18. Å/hour. STS measurements (based ...
  29. [29]
    [PDF] INFORMATION TO USERS The most advanced technology has ...
    When the scanning tunneling microscope (STM) was invented in 1982 by ... as placing our STM on a floating granite table ... excellent way to isolate ones STM from ...
  30. [30]
    [PDF] Identification of Functional Groups Using Scanning Tunneling ...
    Jan 1, 2000 · Also, the microscope is placed on a bungee swing supported by an air table, which provides vibration isolation, necessary to reduce noise lines ...
  31. [31]
    [PDF] J. Chae , S. Y. Jung , S. J. Woo , H. J. Yang , H. Baek , J. Ha , Y. J. ...
    Three air table legs support the granite table, at 500 kg, and the cut off frequency is below ~3Hz. A helium Dewar is attached to three additional air dampers ...
  32. [32]
    [PDF] In-situ Ag tip preparation and validation techniques for scanning ...
    Here we present several in-situ Ag tip preparation and validation techniques for the microscopist to use depending on their intended application, including ...
  33. [33]
    Design and performance of an ultra-high vacuum scanning ...
    Oct 3, 2013 · However, most instruments operate at tempera- tures above 1 K, limiting access to exotic electronic phases and quantum effects expected at lower ...
  34. [34]
    [PDF] Dielectric Constant Measurements Using Atomic Force Microscopy ...
    Aug 23, 2012 · The AFM also allowed insulating materials to be analyzed. Here, very small sharp probing tip is scanned very closely above the sample surface.
  35. [35]
  36. [36]
    Scanning Tunneling Microscopy (STM): An Overview
    Scanning Tunneling Microscopy allows researchers to map a conductive sample's surface atom by atom with ultra-high resolution, without the use of electron ...
  37. [37]
    8.3: Scanning Tunneling Microscopy - Chemistry LibreTexts
    Aug 28, 2022 · In a typical STM operation process, the tip is scanning across the surface of sample in x-y plain, the instrument records the x-y position of ...Missing: procedure coarse
  38. [38]
    Scanning Tunneling Microscopy (STM) - Park Systems
    STM measures the tunneling current between tip and sample, giving highly accurate sub-nanometer scale images you can use to gain insights into sample ...
  39. [39]
    [PDF] Surface Tunneling Microscopy and Spectroscopy STM and STS
    STM provides atomic resolution surface images via electron tunneling. STS measures local electronic structure. STM uses a sharp tip and voltage between tip and ...
  40. [40]
    Interpreting tunneling spectroscopic maps of a dinuclear Co(II ...
    For measuring the differential conductance d I / d V a lock-in amplifier was used and a sinusoidal modulation voltage (0.5 mV rms at 667.8 Hz) was applied to ...
  41. [41]
    introduction to STM and STS
    STM is known to be used for the investigation of surface topography. However, we are more interested in the electronic properties of our samples.
  42. [42]
    Combined I(V) and dI(V)/dz scanning tunneling spectroscopy
    Jul 13, 2018 · The I(V) spectrum is recorded by ramping the bias voltage, while the feedback loop of the scanning tunneling microscope is disabled.
  43. [43]
    [PDF] Scanning Tunneling Microscopy and Spectroscopy
    One of these phenomena is the tunnel effect, which describes the ability of an electron to tunnel through a vacuum barrier from one electrode to the other.
  44. [44]
    Spin-polarized Scanning Tunneling Microscopy (SP-STM)
    Mar 30, 2020 · SP-STM is STM performed with a magnetic tip. When both surface and tip are magnetic, the imbalance of spin-up and spin-down electrons around the Fermi level.
  45. [45]
    Preparation of magnetic tips for spin-polarized scanning tunneling ...
    Apr 24, 2015 · Here we demonstrate preparation of magnetic tips, both ferromagnetic and antiferromagnetic, on single crystals of FeTe.Abstract · Article Text
  46. [46]
    Bulk ferromagnetic tips for spin-polarized scanning tunneling ...
    Jan 25, 2019 · In a conventional method, a spin-polarized tip is made by depositing ferromagnetic (e.g., Fe,6,7 Gd8) or antiferromagnetic (e.g., Cr,9 Mn10) ...
  47. [47]
    Time resolution of terahertz scanning tunneling microscopy ...
    Apr 17, 2025 · A terahertz scanning tunneling microscope (THz-STM) is an STM that allows us to perform time-resolved measurements of tunneling currents ...INTRODUCTION · II. EXPERIMENTAL SETUP · III. RESULTS AND DISCUSSION
  48. [48]
    Ultrafast Dynamics Revealed with Time-Resolved Scanning ...
    Mar 17, 2023 · A THz pulse applies a transient voltage at the STM junction to induce ultrafast electron tunneling. This voltage variance from a strong electric ...
  49. [49]
    None
    Nothing is retrieved...<|control11|><|separator|>
  50. [50]
    7 × 7 Reconstruction on Si(111) Resolved in Real Space
    Jan 10, 1983 · The 7 × 7 reconstruction on Si(111) was observed in real space by scanning tunneling microscopy. The experiment strongly favors a modified adatom model.Missing: original | Show results with:original
  51. [51]
    Transition from reciprocal-space to real-space surface science ...
    Shahrivar 11, 1382 AP · The scanning tunneling microscope enabled a transition to real-space imaging, making surface science visual and thus much more accessible. ... STM ...
  52. [52]
    Theory of the scanning tunneling microscope | Phys. Rev. B
    '' The tunneling current is found to be proportional to the local density of states of the surface, at the position of the tip.
  53. [53]
    [PDF] Scanning Tunneling Microscopy of Metals and Semiconductors ...
    Applications of scanning tunneling microscopy (STM) to surfaces of both metals and semiconductors have rapidly expanded during the last fifteen years.Missing: core | Show results with:core
  54. [54]
    [PDF] Imaging atoms and molecules on surfaces by scanning tunnelling ...
    Nov 4, 2011 · This review discusses nearly 30 years of scanning tunnelling microscopy (STM) work on high resolution imaging of numerous materials systems, ...
  55. [55]
    Surface point defects on bulk oxides: atomically-resolved scanning ...
    Mar 17, 2017 · This review gives an overview of using Scanning Probe Microscopy (SPM), in particular Scanning Tunneling Microscopy (STM), to study the changes in the local ...
  56. [56]
    High-resolution scanning tunneling microscopy imaging of ... - PNAS
    We present scanning tunneling microscopy (STM) images of single-layer graphene crystals examined under ultrahigh vacuum conditions.
  57. [57]
    Positioning single atoms with a scanning tunnelling microscope
    Apr 5, 1990 · Here we report the use of the STM at low temperatures (4 K) to position individual xenon atoms on a single-crystal nickel surface with atomic pre-cision.
  58. [58]
    Atom manipulation with the STM: nanostructuring, tip ...
    Repulsive (pushing) as well as discontinuous (pulling) and continuous (sliding) attractive manipulation modes could be distinguished on Cu(2 1 1) for CO ...
  59. [59]
    [PDF] STM Single Atom/Molecule Manipulation and Its Application ... - arXiv
    (b) STM tip-height manipulation curves correspond to (1) pulling, (2) pushing, and (c) sliding modes. In the “pushing” mode, the tip first climbs up the contour.
  60. [60]
    Confinement of Electrons to Quantum Corrals on a Metal Surface
    A method for confining electrons to artificial structures at the nanometer lengthscale is presented. Surface state electrons on a copper(111) surface were ...Missing: original | Show results with:original
  61. [61]
    Scanning Tunneling Spectroscopy - Annual Reviews
    Nov 21, 2008 · the tunneling current is measured over each pixel of the STM picture. ... work function, and F is the field strength in volts per angstrom.
  62. [62]
    Recent advances in inelastic electron tunneling spectroscopy
    Inelastic electron tunneling spectroscopy (IETS) based on scanning tunneling microscopy (STM) opens a new avenue for vibrational spectroscopy at single bond ...
  63. [63]
    Inelastic electron tunneling spectroscopy for probing strongly ...
    Mar 6, 2020 · We present an extension of the tunneling theory for scanning tunneling microscopy (STM) to include different types of electron-vibrational couplings.
  64. [64]
    (PDF) Recent developments in scanning tunneling spectroscopy of ...
    Aug 7, 2025 · Several recent developments in scanning tunneling spectroscopy (STS) of semiconductor surfaces are reviewed. First, the normalization of spectra is discussed.
  65. [65]
    Scanning tunneling spectroscopy study of the electronic structure of ...
    Both surfaces show the same tunneling behavior. The presence of a semiconducting band gap at the magnetite surface is evident from the spectra. (b): ...
  66. [66]
    Evaluation of curves in scanning tunneling spectroscopy of organic ...
    Jun 21, 2007 · For this purpose, scanning tunneling spectroscopy (STS) is used to obtain information about the density of electronic states (DOS) of the ...
  67. [67]
    Removal of multiple-tip artifacts from scanning tunneling microscope ...
    Nov 14, 2015 · Our analysis clarifies why crystallographic averaging works well in removing the effects of a blunt STM tip (that consists of multiple mini-tips) ...
  68. [68]
    Invited Review Article: Multi-tip scanning tunneling microscopy
    Oct 15, 2018 · In order to use a scanning tunneling microscope (STM) for electrical transport measurements at nanostructures, more than one tip is required.
  69. [69]
    Real-space post-processing correction of thermal drift and ...
    Jan 26, 2017 · ... STM can be obscured by probe-tip fluctuations and by electronic noise. This is particularly true with room-temperature STM imaging, where ...
  70. [70]
    (PDF) Scanning Tunneling Microscopy: A Review - ResearchGate
    This article tries to explain the principles and applications of the STM based on attractions and repulsions of atoms at very small distances to collect ...
  71. [71]
    Scanning Tunneling Microscopy of Biological Structures: An Elusive ...
    Aug 31, 2022 · Scanning tunneling microscopy (STM) is a technique that can be used to directly observe individual biomolecules at near-molecular scale.<|control11|><|separator|>
  72. [72]
    The composition and structure of the ubiquitous hydrocarbon ...
    Nov 9, 2022 · Here we show that the common ambient contamination on the surface of: graphene, graphite, hBN and MoS 2 is composed of a self-organized molecular layer.
  73. [73]
    Colloquium: Time-resolved scanning tunneling microscopy
    May 17, 2010 · ... STM to measure dynamic events is increased by a factor of 10 3 as compared to conventional STM imaging techniques ( Swartzentruber, 1996
  74. [74]
    Ultrafast Dynamics Revealed with Time-Resolved Scanning ...
    Mar 17, 2023 · STM reads this density through tunneling current change. The tunnel current induced by the second optical pulse depends on the delay time ...Missing: factor | Show results with:factor<|control11|><|separator|>
  75. [75]
    Attosecond coherent manipulation of electrons in tunneling ...
    We demonstrate coherent control over electrons in a tunnel junction of a scanning tunneling microscope by means of precise tuning of the carrier-envelope phase.
  76. [76]
    In Situ Scanning Tunneling Microscopy of Corrosion of Silver-Gold ...
    An in situ scanning tunneling microscope (STM) was used to observe the morphological changes accompanying the selective dissolution of Ag from low-Ag ...<|separator|>
  77. [77]
    Combining Electrochemical Scanning Tunneling Microscopy with ...
    We introduce the reader to the simultaneous operation of electrochemical scanning tunneling microscopy and force microscopy using a qPlus sensor.
  78. [78]
    Electrochemical Scanning Tunneling Microscopy - ScienceDirect.com
    Electrochemical scanning tunneling microscopy (EC-STM) has been shown to be a powerful tool for the detailed investigation of (dynamic) oxidation processes.
  79. [79]
    An Ultrahigh Vacuum High Speed Scanning Tunneling Microscope
    Oct 31, 2015 · The prototype operates with a tip scan speed up to 10(5) nm/s. At this high tip speed, we have made real time videos of nickel and gold at rates ...
  80. [80]
    A high bandwidth microelectromechanical system-based ...
    Jul 31, 2019 · A number of attempts have been made to increase the Z axis bandwidth of scanning probe microscopes by redesigning the piezopositioners. A common ...Missing: piezos | Show results with:piezos
  81. [81]
    (PDF) A novel AFM/STM/SEM system - ResearchGate
    This AFM/STM/SEM system enables us to image a sample conventionally by SEM as well as to investigate the local surface topography by AFM or STM. This device ...
  82. [82]
    Atomic-Resolution Imaging of Micron-Sized Samples Realized by ...
    Jan 22, 2023 · In this paper, we present the design and construction of an atomic-resolution STM that can operate in a 9 T high magnetic field. More ...<|separator|>
  83. [83]
    A qPlus-based scanning probe microscope compatible with optical ...
    Apr 1, 2022 · We design and develop a scanning probe microscope (SPM) system based on the qPlus sensor for atomic-scale optical experiments.INTRODUCTION · QPLUS SENSOR · STM performance · Optical performance
  84. [84]
    Imaging modes of atomic force microscopy for application ... - PubMed
    Apr 6, 2017 · Here, we review the basic principles, advantages and limitations of the most common AFM bioimaging modes, including the popular contact and dynamic modes.
  85. [85]
    [cond-mat/0305119] Advances in atomic force microscopy - arXiv
    May 6, 2003 · Vacuum AFM opens up new classes of experiments, ranging from imaging of insulators with true atomic resolution to the measurement of forces ...
  86. [86]
    Atomic Force Microscopy - Nanoscience Instruments
    Gerd Binnig and Heinrich Rohrer developed the first working STM while working at IBM Zurich Research Laboratories in Switzerland. This instrument would later ...
  87. [87]
    Review: Advanced Atomic Force Microscopy Modes for Biomedical ...
    Dec 2, 2022 · The term “tapping mode” is often used when the interaction force is within the repulsive region of the Lennard-Jones potential between the probe ...
  88. [88]
    Sensing current and forces with SPM - ScienceDirect.com
    Atomic force microscopy (AFM) and scanning tunneling microscopy (STM) are well established techniques to image surfaces and to probe material properties at ...
  89. [89]
    Magnetic Force Microscopy - an overview | ScienceDirect Topics
    The first MFM images were obtained by Martin and Wickramasinghe [52] in 1987. Here, data on the field profile of a thin-film magnetic recording head were ...Missing: original | Show results with:original
  90. [90]
    Near-Field Scanning Optical Microscopy (NSOM) - History and ...
    Jun 23, 2010 · SNOM is the acronym for Scanning Near-Field Optical Microscopy, an alternative name for NSOM (Near-Field Scanning Optical Microscopy).Missing: seminal | Show results with:seminal
  91. [91]
    [PDF] Kelvin probe force microscopy and its application
    Kelvin probe force microscopy (KPFM) is a tool that enables nanometer-scale imaging of the surface potential on a broad range of materials.
  92. [92]
    Ballistic electron emission microscopy and spectroscopy
    Jul 20, 2016 · The original model10 for BEEM spectroscopy involved tunnel injection and simple assumptions for carrier transport through the sample structure.
  93. [93]
    Kelvin Probe Force Microscopy: Developments and Applications
    May 3, 2023 · Kelvin probe force microscopy (KPFM) is an AFM mode that extracts the contact potential difference between a probe and sample by observing ...