Fact-checked by Grok 2 weeks ago

Evanescent field

An evanescent field, also known as an evanescent wave, is a non-propagating that occurs near the between two media with different refractive indices, particularly during when travels from a higher-index medium to a lower-index one at an angle greater than the . This field exhibits an in perpendicular to the , typically over a on the of the , without transporting net in that . It arises to satisfy boundary conditions at the , resulting in a with an imaginary component that leads to the decaying behavior rather than . Evanescent fields are generated when an incident wave undergoes total internal reflection, such as a laser beam in glass encountering an air interface, where the transmitted wave into the rarer medium becomes evanescent instead of refracting. Mathematically, the field in the lower-index medium can be described by a form like e^{i k_x x - \beta z}, where k_x is the real propagation constant parallel to the interface, \beta is positive and real (leading to exponential decay e^{-\beta z} with distance z into the medium, and phase progression e^{i k_x x} parallel to the interface), and no phase progression occurs in the decay direction. The penetration depth, often around \lambda / (2\pi) where \lambda is the wavelength, increases as the incidence angle approaches the critical angle or with smaller refractive index contrasts. Key properties of evanescent fields include their confinement to the near-field region, zero net in the perpendicular direction (indicating no energy flow across the boundary), and potential for transverse spin angular momentum. These characteristics enable diverse applications in and , such as evanescent wave coupling in optical fibers and resonators for light transfer between waveguides, near-field scanning optical (NSOM) for subwavelength resolution imaging down to 10-100 nm, and biosensors that detect chemical or biological interactions via field perturbations. Additionally, they are utilized in fluorescence (TIRF) , surface plasmon resonance sensors, and atomic traps.

Fundamentals

Definition and Terminology

An evanescent field, also known as an evanescent wave, is a non-propagating electromagnetic or acoustic field that arises near an or between two and decays exponentially with distance away from that , without transporting net energy in the direction perpendicular to the . This field exists only in the near-field region and does not radiate or propagate as a traveling wave into the second medium. The term "evanescent" originates from the Latin ēvānēscēns, meaning "disappearing" or "vanishing," aptly describing the field's rapid attenuation. The concept traces its roots to early wave in the Newtonian , with the specific terminology emerging in 19th-century studies of behavior at interfaces. For instance, in 1891, Paul Drude and employed the phrase "standing evanescent wave" in their pioneering experiments on excitation via such fields. Over time, the nomenclature evolved, with "evanescent wave" becoming the standard synonym in optical literature to emphasize its wave-like yet non-propagating nature. Evanescent fields differ fundamentally from propagating waves, which have real components allowing energy transport over arbitrary distances; in contrast, evanescent fields possess an imaginary perpendicular component, leading to and the formation of near-field patterns without far-field radiation. This distinction underscores their role as bound or reactive fields rather than free-propagating ones. The terminology is most prevalent in and electromagnetics, but analogous evanescent phenomena appear in acoustics (e.g., wave boundaries), (e.g., tunneling through barriers), and dynamics.

Physical Properties

The amplitude of an evanescent field exhibits an profile perpendicular to the , described by E(z) \propto e^{-\kappa z}, where z is the from the into the lower-index medium and \kappa is the decay constant that determines the field's . This decay arises in scenarios such as at a , confining the field to a thin layer near the . The decay constant \kappa depends on the refractive indices of the media, the angle of incidence, and the , typically resulting in penetration depths on the order of the . Evanescent fields are inherently near-field phenomena, existing only in close proximity to or —generally within a comparable to the —beyond which they rapidly vanish without contributing to propagating radiation. Unlike far-field electromagnetic waves that radiate indefinitely, evanescent fields do not propagate freely and are non-radiative in the direction normal to the , as evidenced by the time-averaged having zero real component perpendicular to the boundary. However, can be stored within or transferred laterally along the through interactions with nearby structures. The behavior of evanescent fields shows dependence, with transverse electric () and transverse magnetic (TM) modes exhibiting differences in their field orientations relative to the . However, the is the same for both transverse electric () and transverse magnetic (TM) modes at isotropic under . Additionally, the is influenced by and ; shorter wavelengths lead to shallower decay, such as approximately 100-200 nm for visible light at a glass-air . Experimental confirmation of evanescent fields relies on indirect methods like excitation, where molecules near the are selectively illuminated within the field's decay range, enabling high-resolution imaging in techniques such as . Scattering observations, such as those from particles or biomolecules in the evanescent zone, further verify the field's presence and profile by measuring intensity variations that match the . These techniques demonstrate the field's confinement and non-propagating nature without direct far-field detection.

Mathematical Description

Derivation from Wave Equations

The mathematical description of evanescent fields begins with the , which governs time-harmonic electromagnetic waves in source-free, linear, isotropic media. Derived from under the assumption of time-harmonic fields \mathbf{E}(\mathbf{r}, t) = \mathbf{E}(\mathbf{r}) e^{-i\omega t} and \mathbf{H}(\mathbf{r}, t) = \mathbf{H}(\mathbf{r}) e^{-i\omega t}, with no free charges or currents, the curl equations simplify to \nabla \times \mathbf{E} = i\omega \mu \mathbf{H} and \nabla \times \mathbf{H} = -i\omega \epsilon \mathbf{E}. Taking the curl of the first and substituting the second yields \nabla \times (\nabla \times \mathbf{E}) = \omega^2 \epsilon \mu \mathbf{E}, which, using the vector identity \nabla \times (\nabla \times \mathbf{E}) = \nabla (\nabla \cdot \mathbf{E}) - \nabla^2 \mathbf{E} and \nabla \cdot \mathbf{E} = 0, results in the vector : \nabla^2 \mathbf{E} + k^2 \mathbf{E} = 0, where k = \omega \sqrt{\epsilon \mu} is the wave number. For plane waves at a , assume a two-dimensional with the interface at z = 0, separating medium 1 (z < 0, refractive index n_1 = \sqrt{\epsilon_1 \mu_1 / \epsilon_0 \mu_0}) from medium 2 (z > 0, n_2 < n_1). The incident wave in medium 1 propagates as \mathbf{E}_i = \mathbf{E}_0 e^{i (k_{1x} x + k_{1z} z)}, with k_{1x} = k_1 \sin \theta_i and k_{1z} = k_1 \cos \theta_i, where k_1 = n_1 k_0 and k_0 = \omega / c. The reflected and transmitted waves must satisfy the Helmholtz equation in their respective media and boundary conditions: continuity of tangential \mathbf{E} and \mathbf{H} (or normal \mathbf{D} and \mathbf{B}) at z = 0. Phase matching along x requires k_{1x} = k_{2x} = \beta, so the transmitted wave in medium 2 has form e^{i (\beta x + k_{2z} z)}, with k_{2z} = \sqrt{k_2^2 - \beta^2} and k_2 = n_2 k_0. When the angle of incidence \theta_i exceeds the critical angle \theta_c = \sin^{-1}(n_2 / n_1), total internal reflection occurs, as \beta > k_2. This makes k_{2z} imaginary: k_{2z} = i \kappa, where \kappa = \sqrt{\beta^2 - k_2^2} > 0 is real, ensuring the transmitted field decays rather than propagates in z > 0. This follows from extending , n_1 \sin \theta_i = n_2 \sin \theta_t, to complex \theta_t when \sin \theta_i > n_2 / n_1, yielding \sin \theta_t = (n_1 / n_2) \sin \theta_i > 1 and \cos \theta_t = i \sqrt{\sin^2 \theta_t - 1}, so k_{2z} = k_2 \cos \theta_t = i \kappa. The boundary conditions then determine the (magnitude 1, phase shift) and (zero power transmission). The general solution for the transmitted evanescent field in medium 2 is thus \mathbf{E}(x, z) = \mathbf{E}_0 e^{i \beta x} e^{-\kappa z} for z > 0, representing along the (phase progression in x) and exponential decay perpendicular to it (no net power flow in z). This form satisfies the , as \nabla^2 (e^{i \beta x} e^{-\kappa z}) + k_2^2 (e^{i \beta x} e^{-\kappa z}) = (-\beta^2 - \kappa^2 + k_2^2) e^{i \beta x} e^{-\kappa z} = 0 since \kappa^2 = \beta^2 - k_2^2. For vector electromagnetic fields, (TE or TM) imposes additional constraints from , such as the relation between electric and magnetic components. Evanescent fields arise similarly in scalar wave contexts, such as acoustics, where the p satisfies the scalar \nabla^2 p - \frac{1}{c^2} \frac{\partial^2 p}{\partial t^2} = 0, leading to the time-harmonic \nabla^2 p + k^2 p = 0 with k = \omega / c. At interfaces between media with different speeds c_1 < c_2, the same boundary conditions (continuity of and normal velocity) yield analogous evanescent solutions beyond the critical angle, without vector constraints.

Key Parameters and Behaviors

The penetration depth \delta of an evanescent field, defined as the distance over which the field amplitude decays to $1/e of its value at the interface, is given by \delta = 1/\kappa, where \kappa is the decay constant perpendicular to the interface. In optical total internal reflection scenarios, this is typically expressed as \delta = \frac{\lambda}{2\pi \sqrt{\sin^2 \theta - n^2}}, with \lambda the wavelength in the higher-index medium, \theta the angle of incidence, and n = n_2 / n_1 the ratio of refractive indices across the interface (n_1 > n_2). This parameter governs the spatial extent of the field into the lower-index medium, typically ranging from tens to hundreds of nanometers, and decreases as the incidence angle increases beyond the critical angle or as the refractive index contrast grows. The \beta, representing the wave vector component parallel to the , is \beta = k \sin \theta, where k = 2\pi n_1 / \lambda is the wave number in the incident medium. This determines the lateral phase progression of the evanescent field along the , with the v_p = \omega / \beta = c / (n_1 \sin \theta) appearing subluminal relative to the medium's but exhibiting anomalies in evanescent contexts. A notable effect arising from the evanescent field is the Goos-Hänchen shift, a lateral displacement of the reflected beam parallel to the interface upon total internal reflection. For TE (transverse electric) modes, this shift is approximated as \delta_x = \frac{2}{\kappa} \tan \theta, reflecting the field's brief penetration and phase delay. The shift scales with the penetration depth and incidence angle, typically on the order of the wavelength, and originates from the evanescent field's influence on the reflected wavefront. Dispersion relations for evanescent fields, derived from the , link \kappa and \beta via \beta^2 - \kappa^2 = k_2^2 in planar structures, showing how these parameters vary with \omega, incidence \theta, and material properties like refractive indices. In waveguides, for instance, increasing typically increases \kappa (reducing ) while \beta approaches the core medium's wave number limit, with material dispersion further modulating the through frequency-dependent indices. These variations highlight the field's sensitivity to structural and environmental changes. In evanescent regions, phase velocities can exceed the in (v_p > c) under certain conditions, such as near-cutoff modes, yet no net energy transport occurs due to the imaginary perpendicular component, ensuring and preventing information superluminality. The , associated with energy flow, remains subluminal or zero in purely evanescent cases. Evanescent fields exhibit heightened to material changes near the , amplifying small variations \Delta n through shifts in effective mode index or decay rate, often quantified as S = dN_\mathrm{eff} / dn_\mathrm{cover}, where N_\mathrm{eff} is the mode's effective index and n_\mathrm{cover} the covering medium's index. This arises because the field's exponential tail overlaps strongly with the adjacent medium, enabling detection of \Delta n on the order of $10^{-6} RIU in optimized structures.

Generation Mechanisms

Total Internal Reflection

(TIR) occurs at the between two media when propagates from a medium with higher n_1 to one with lower n_2 (n_1 > n_2) and the angle of incidence \theta_i exceeds the \theta_c = \sin^{-1}(n_2 / n_1). Above this angle, the refracted wave cannot propagate in the second medium and instead forms an evanescent field that decays exponentially away from the . For \theta_i > \theta_c, the reflection coefficient reaches unity in magnitude, resulting in 100% reflection of the incident energy back into the first medium, with all non-reflected energy converted into the evanescent field in the second medium. The Fresnel coefficients describe this behavior; for transverse electric (TE) polarization, the reflection coefficient is r_{TE} = 1, while the transmission coefficient for the evanescent wave is given by t_{TE} = \frac{2 \cos \theta_i \cdot i \kappa}{\cos \theta_i + i \kappa}, where \kappa = \sqrt{\sin^2 \theta_i - (n_2 / n_1)^2} is the imaginary component of the transverse wave number in the second medium. Energy conservation is maintained because the time-averaged Poynting vector component normal to the interface in the second medium carries no net power across the boundary, ensuring all incident power is reflected despite the evanescent penetration. Early observations of phenomena related to TIR date to Isaac Newton's (1704), where he described experiments with prisms demonstrating frustrated total internal reflection, in which appeared to "tunnel" through a narrow air gap between two prisms. Later quantitative work, including studies on the intensity and polarization of reflected at grazing angles, was advanced by George Gabriel Stokes in his 1849 paper on the dynamical theory of diffraction, which provided foundational insights into wave behavior at interfaces. Beyond , TIR analogs exist in other wave systems; for in fluids, total internal reflection arises at interfaces between media with differing sound speeds when the incidence angle exceeds a critical value analogous to \theta_c, generating evanescent pressure fields in the lower-speed medium. In , the evanescent wave in TIR serves as a classical analog to particle tunneling through potential barriers, where the wave function decays exponentially in the forbidden region but enables transmission under frustrated conditions.

Guided Wave Structures

In guided wave structures, evanescent fields arise from the confinement of electromagnetic or acoustic waves within a core region surrounded by a lower-index cladding, where the propagation constant β exceeds the wave number in the cladding, k_clad, leading to exponential decay outside the core. This confinement is fundamental to planar waveguides, such as symmetric dielectric slab waveguides, where guided modes exhibit oscillatory fields inside the high-index core and evanescent tails in the cladding regions on both sides. For transverse electric (TE) modes in such structures, the effective index n_eff = β / k_0 > n_clad ensures the field decays as e^{-κ y}, with κ = √(β² - k_clad²), enabling tight mode confinement while allowing limited penetration into the cladding for potential interactions. Cutoff conditions in slab waveguides determine the minimum frequency for guided modes, below which fields become radiative rather than evanescent. For a symmetric slab of thickness d and core wave number k_core, the even TE modes satisfy the transcendental equation tan(k_core d / 2) = (κ_clad / k_core), where κ_clad is the decay constant in the cladding; this equation arises from matching boundary conditions at the core-cladding interfaces, ensuring continuity of the field and its derivative. At cutoff, κ_clad approaches zero, corresponding to β = k_clad, and higher-order modes require larger d or higher frequencies to support evanescent tails. Odd modes follow a similar form with cotangent, highlighting the role of symmetry in mode classification and evanescent behavior. In fiber optics, evanescent fields extend beyond the core-cladding boundary of step-index fibers, particularly for single-mode operations where the core diameter is comparable to the . These fields decay exponentially in the surrounding medium, with on the order of the , facilitating weak but controllable interactions with external environments without significant loss of guided power. This property is exploited in structures like tapered or D-shaped fibers, where partial removal of the cladding exposes the evanescent tail for enhanced overlap with adjacent media. Photonic crystals and gratings generate evanescent modes through bandgap engineering, where periodic structures create frequency bands forbidding , leading to evanescent in the gaps. At defect sites, such as line defects in two-dimensional photonic crystal waveguides, guided modes emerge with evanescent fields confined to the defect region, decaying rapidly into the surrounding crystal due to the photonic bandgap. Gratings, as one-dimensional photonic crystals, similarly induce evanescent modes near band edges or defects, enabling slow-light with strong field localization. Surface plasmon polaritons (SPPs) represent a electromagnetic at metal- , featuring evanescent fields decaying perpendicularly from the on both sides. The for the dielectric side is given by κ = √(β² - ε k₀²), where β is the SPP , ε is the , and k₀ is the free-space wave number; this results in sub-wavelength confinement, with typical lengths of 100-200 for visible wavelengths. SPPs thus provide a mechanism for evanescent field generation in submicron-scale waveguides, distinct from all-dielectric structures due to their dispersive, lossy nature. Acoustic and mechanical analogs to optical evanescent fields occur in phononic waveguides, where periodic structures create phononic bandgaps leading to evanescent modes at interfaces or defects. In finite phononic crystals, evanescent Bloch waves decay exponentially away from the core, analogous to optical cases, with the decay governed by imaginary wave vectors in the bandgap; this enables sound wave confinement in structures like sonic crystals or elastic metamaterials. Such modes are crucial for designing acoustic analogs of photonic devices, including with controlled evanescent penetration for .

Interactions and Effects

Field Penetration and Decay

In evanescent fields penetrating absorbing media, the perpendicular component of the wave vector, denoted as κ, becomes , expressed as κ = κ_r + i κ_i, where the real part κ_r governs the oscillatory behavior and the imaginary part κ_i dictates the . This complexity arises because the of the absorbing medium is itself complex, leading to a field profile that exhibits damped oscillations rather than pure , as the wave penetrates a short distance into the medium before being significantly attenuated. Such behavior is particularly relevant in scenarios like at interfaces with lossy dielectrics, where the evanescent tail interacts with mechanisms, modifying the field's . The presence of adjacent media or nearby objects can distort the evanescent field profile, inducing perturbations that shift the effective decay constant κ. For instance, a scatterer or dielectric perturbation in close proximity alters the local boundary conditions, effectively changing the field's exponential decay rate through induced dipoles or refractive index gradients. This shift in κ can be modeled using perturbation theory, where the interaction strength scales with the inverse cube of the distance to the perturbing object, enabling sensitive detection of nanoscale features. In practical terms, such distortions are observed in near-field optics, where the evanescent field's homogeneity is disrupted by environmental inhomogeneities, leading to asymmetric decay profiles. In the time domain, transient evanescent fields excited by pulsed sources exhibit dispersive decay, where the field's temporal evolution depends on the frequency content of the pulse and the medium's dispersion. For ultrashort pulses, the evanescent component disperses as different frequency modes decay at varying rates, resulting in pulse broadening and a non-exponential temporal decay profile. This dispersive behavior is pronounced in structures like waveguides under pulsed illumination, where the evanescent tail's decay time must exceed the pulse duration to maintain subwavelength resolution in time-resolved measurements. At high intensities, nonlinear effects such as (SPM) in evanescent fields arise due to the Kerr nonlinearity, altering the profiles by inducing an intensity-dependent phase shift that modifies the effective . In evanescent configurations like polaritons, SPM causes the to vary along the field, leading to a modified spatial that can enhance or suppress depending on the and . This effect is leveraged in nonlinear , where high-intensity evanescent tails in subwavelength structures enable tunable rates through SPM-induced spectral broadening. Quantum aspects of evanescent fields manifest in atom-light s, where the subwavelength confinement allows precise control over states at nanoscale distances from the . In setups using optical nanofibers, the evanescent field couples to dipoles, enabling coherent manipulation of quantum states with resolution below the limit, as the field's rapid localizes the volume. Two-color evanescent fields, for example, facilitate atom trapping and guiding by balancing attractive and repulsive potentials, achieving subwavelength positioning of atoms for processing. Evanescent field microscopy techniques, such as total internal reflection fluorescence (TIRF) microscopy, directly probe the decay profile by exciting fluorophores within the evanescent tail and measuring the resulting emission intensity as a function of depth. Calibration methods using tilted fluorescent layers or known scatterers quantify the penetration depth, confirming exponential decay with deviations attributable to medium properties. Plasmonic near-field scanning optical microscopy (p-NSOM) further enhances resolution, mapping the evanescent decay with sub-10 nm precision by adiabatically focusing the field onto a probe tip. These methods verify the basic decay form, typically e^{-κ z} as derived from wave equations, while revealing subtle modifications in complex environments.

Coupling Between Structures

Evanescent-wave coupling occurs when the evanescent fields from two adjacent s overlap, enabling between guided modes without direct physical contact between the cores. This phenomenon is fundamental to directional couplers, where light launched into one partially transfers to the adjacent one over an interaction length determined by the strength. The theory is described by coupled-mode theory, which models the amplitude evolution of modes in each as a system of coupled differential equations. In a basic directional coupler with two parallel waveguides, the power transferred from the input waveguide to the output waveguide is given by P = \sin^2(\kappa L), where \kappa is the coupling coefficient and L is the interaction length. The coupling coefficient \kappa quantifies the overlap of the evanescent fields and depends on factors such as the gap distance between waveguides and the mode overlap integral, often expressed as \kappa = \frac{\omega \epsilon_0}{4P} \iint (n^2 - n_1^2) \mathbf{E}_1 \cdot \mathbf{E}_2 \, dx \, dy, with n the refractive index perturbation, \mathbf{E}_1 and \mathbf{E}_2 the mode fields, and P the mode power normalization. The coupling length L_c, defined as the distance for complete power transfer (P = 1), is L_c = \frac{\pi}{2\kappa}, which decreases as the gap narrows due to stronger field overlap. For efficient coupling, synchronous conditions must be met, where the propagation constants satisfy \beta_1 \approx \beta_2 (phase-matching), enabling full power oscillation; asynchronous coupling (\Delta \beta = \beta_1 - \beta_2 \neq 0) results in incomplete transfer, with maximum power P = \frac{\kappa^2}{\kappa^2 + (\Delta \beta / 2)^2} \sin^2(\sqrt{\kappa^2 + (\Delta \beta / 2)^2} L). To enhance broadband performance and reduce sensitivity to fabrication variations, tapered and adiabatic introduce gradual changes in geometry, such as linearly varying widths or separations, allowing modes to evolve without exciting higher-order or radiating modes. In adiabatic designs, the follows a slow variation that preserves mode purity, achieving near-100% transfer efficiency over shorter lengths compared to uniform couplers, particularly for multimode systems. These structures minimize losses by avoiding abrupt discontinuities that could scatter light into radiation modes. Evanescent coupling enables wavelength-selective devices like modal filters, where the coupling coefficient's wavelength dependence—arising from dispersion in mode overlap and phase-matching—allows selective transfer of specific modes or wavelengths; for instance, a 3 dB coupler (50% transfer) at L = L_c / 2 can filter LP11 modes from double-mode fibers with high selectivity (>20 dB rejection). However, imperfect overlap or mismatches introduce losses, including radiation loss from mode leakage into the cladding (scaling with gap size and length, often <0.1 dB/cm in optimized silicon waveguides) and absorption from material imperfections, exacerbated in asynchronous regimes or non-adiabatic tapers.

Applications

Optical Sensing and Biosensors

Evanescent fields play a crucial role in optical sensing by enabling the detection of changes in the or presence of analytes near the surface through interactions with the exponentially decaying outside the guiding structure. This surface-sensitive approach confines the sensing volume to a thin layer adjacent to the or , allowing for selective probing of environmental perturbations without interference from the bulk medium. In (SPR) sensors, the evanescent field from at a metal-dielectric excites surface plasmons, whose condition shifts with variations in the analyte's , typically measured via angular interrogation. This configuration achieves high sensitivity, with limits of detection around 10^{-6} refractive index units (RIU) through observable shifts in the resonance angle. For instance, silver-based SPR setups demonstrate angle shifts of approximately 0.68 degrees for a 0.01 RIU change, enabling precise biomolecular interaction monitoring. Fiber-optic evanescent sensors utilize the evanescent field by or removing the cladding of optical fibers, exposing to the surrounding medium for direct interaction and detection of changes or chemical species. These sensors are particularly effective for inline monitoring in harsh environments, such as measuring concentrations in aqueous solutions where evanescent correlates with . Etched germanium-doped single-mode fibers have been demonstrated for distributed sensing with resolutions suitable for environmental applications. Biosensors leveraging evanescent fields enable label-free detection of biomolecules by monitoring perturbations to the field caused by binding events on the sensor surface, often quantified through phase or intensity shifts. In Mach-Zehnder interferometer (MZI) configurations, the evanescent field in one arm interacts with the , producing an interferometric signal proportional to the bound mass, achieving detection limits as low as 10^{-7} RIU for protein or DNA hybridization assays. Silicon nitride slot-waveguide MZIs, for example, have shown high sensitivity for real-time monitoring of biomolecular interactions without fluorescent labels. The primary limitation of evanescent field sensors arises from the shallow of the field into the , typically 100-200 nm, which restricts the effective sensing volume to surface-bound and excludes deeper contributions. This confinement, while advantageous for selectivity, can reduce signal strength for low-concentration or sparsely distributed targets. Compared to bulk optical methods, evanescent field sensors offer superior by localizing detection to the near-surface region, minimizing and enabling high specificity for surface-adsorbed analytes. They also support real-time monitoring of dynamic processes, such as binding kinetics, in a compact and integrable format suitable for point-of-care applications. Recent advances since 2020 have focused on integrating evanescent field sensors with to create portable point-of-care devices, enhancing sample handling and for clinical diagnostics. For example, evanescent wave-based platforms combined with microfluidic channels have improved detection speed and sensitivity, incorporating hybrid designs for on-chip processing of complex biological samples. These developments emphasize scalable fabrication and all-optical label-free operation for rapid disease screening. As of 2025, comprehensive reviews highlight ongoing innovations in evanescent wave-based optical biosensors for medical diagnostics and , including highly sensitive surface-enhanced (SERS) probes for trace analyte detection.

Integrated Photonics and Devices

In integrated , evanescent fields play a crucial role in enabling compact and within photonic integrated circuits (PICs), particularly on silicon-on-insulator (SOI) platforms. These fields facilitate interactions between closely spaced waveguides, allowing for efficient power transfer without direct physical contact. Directional couplers and splitters, for instance, rely on evanescent to achieve precise power division, such as 3 dB splitting in symmetric photonic designs where the evanescent field overlap between waveguides determines the coupling length and efficiency. In one optimized implementation, a compact directional coupler with a 300 nm gap and 670 µm coupling length achieves measured coupling efficiencies of 73% at 1530 nm and 77% at 1653.7 nm, enabling low-loss multiplexing for telecom wavelengths. Such devices are foundational for optical signals in PICs, with the evanescent field decay controlled to minimize . Evanescent fields also underpin electro-optic modulators, where modulation arises from changes in the induced by applied overlapping the evanescent region. In hybrid silicon evanescent Mach-Zehnder modulators (MZMs), the optical mode in a evanescently couples to carrier-depletion regions in bonded III-V quantum wells, enabling high-speed operation through phase shifts. The first such device demonstrated up to 40 Gb/s with an 11.4 dB extinction ratio and a voltage-length product of 0.85 V·cm, achieving a π-phase shift via carrier density in the evanescently accessed active layer. This approach leverages the strong evanescent field confinement in SOI waveguides to integrate 's passive routing with III-V's active electro-optic effects, supporting data rates essential for telecom applications. For (WDM), arrayed gratings (AWGs) utilize evanescent mode overlaps between adjacent waveguides in the to disperse and route multiple wavelengths. The evanescent coupling strength, governed by waveguide spacing, directly influences channel ; numerical analyses indicate a minimum separation of 10–11 µm is required to limit below -20 in AWGs operating at telecom bands. Low- designs incorporating parabolic tapers at junctions further enhance performance, achieving insertion losses under 5 across 16 channels spaced at 100 GHz. These evanescent interactions ensure precise phase control for demultiplexing, forming a of scalable WDM filters in PICs. Fabrication challenges in these devices include losses from evanescent field , particularly at bends where the field extends into rough sidewalls. Strip waveguides exhibit higher propagation losses (e.g., ~1.5 dB/cm at mid-infrared wavelengths) compared to waveguides (~0.1 dB/cm), as the former's full etch exposes more sidewall area to scattering in the evanescent tail. losses can reach 0.02 dB per 90° turn for 10 µm radii in SOI structures, mitigated by designs that partially confine the and reduce evanescent exposure to etched surfaces. These considerations guide material and geometry choices to maintain low overall insertion losses in integrated systems. Scalability of evanescent-field-based PICs has advanced significantly on SOI platforms since the , transitioning from small-scale (SSI) with 1–10 components to large-scale (LSI) exceeding 500 elements for transceivers. Low-loss SOI waveguides, introduced in the , enable dense packing with evanescent for reduced , supporting commercial modulators and multiplexers in data centers by the mid-. Modern SOI chips integrate thousands of components for WDM routing, with evanescent fields enabling heterogeneous bonding and low-crosstalk layouts critical for co-packaged . In quantum photonics, evanescent facilitates integration of single-photon sources into PICs, such as by tapering nanowires to overlap evanescent fields with waveguides. This hybrid approach achieves efficient collection of photons from quantum dots or defects, with demonstrated coupling efficiencies exceeding 50% for on-chip routing in quantum networks. Such evanescently coupled sources enable scalable processing, leveraging SOI compatibility for low-loss propagation. Additionally, as of 2025, low-power evanescent field atom guides based on nanofibers have been developed for quantum inertial sensors, enhancing applications in cold atom .

Microscopy and Imaging Techniques

Total internal reflection fluorescence (TIRF) microscopy utilizes the evanescent field generated at a glass-water to selectively excite fluorophores within approximately 100 nm of the surface, enabling high-contrast of cellular structures near substrates without illuminating the bulk sample. This technique, pioneered by Daniel Axelrod in the early 1980s through key developments including the 1981 demonstration of cell-substrate contact visualization, confines excitation to a thin optical section, reducing and background fluorescence in deeper regions. In biological applications, TIRF excels at live-cell of plasma membranes and adhesion sites, such as monitoring vesicle fusion or protein dynamics at the cell-substratum , where the evanescent field's shallow penetration—typically 50-150 nm depending on and refractive indices—preserves viability by minimizing photodamage to intracellular components. Near-field scanning optical microscopy (NSOM), also known as scanning near-field optical microscopy (SNOM), employs a subwavelength to access evanescent fields, achieving resolutions below λ/10 by confining light interaction to the near-field zone, far surpassing diffraction-limited far-field . Developed in the 1980s by Dieter W. Pohl and colleagues at , the technique raster-scans a nanometer-sized over the sample, coupling evanescent waves to reveal nanoscale optical properties and simultaneously. NSOM's ability to evanescent fields enables detailed surface imaging in and , such as mapping molecular distributions on cell membranes with resolutions down to 20-50 nm. Evanescent wave lithography leverages the rapid decay of evanescent fields to pattern nanoscale features on surfaces, creating patterns or standing waves that expose photoresists in subwavelength dimensions without propagating light's constraints. This method, demonstrated in early works using setups, achieves feature sizes as small as 26 nm at 193 nm wavelengths by exploiting the field's to control depth and resolution in the near-field regime. Applications include fabricating periodic nanostructures for , where the evanescent provides high-fidelity patterns limited primarily by the field's . Evanescent fields break classical diffraction limits in super-resolution imaging by enabling near-field interactions that capture high spatial frequencies otherwise lost in far-field propagation, as seen in variants of stimulated emission depletion (STED) microscopy adapted for total internal reflection. In TIR-STED, the evanescent excitation combined with a depletion beam confines the effective point spread function to tens of nanometers axially and laterally, achieving resolutions below 50 nm for surface-bound fluorophores. This approach enhances selectivity in imaging, such as tracking single molecules on live cell surfaces, while the evanescent penetration depth—detailed in fundamental parameters—ensures minimal out-of-focus excitation.

References

  1. [1]
    Evanescent Waves - RP Photonics
    Evanescent waves have rapidly decaying amplitudes without energy transport. They are used in various optical applications.
  2. [2]
    5.12: Evanescent Waves - Engineering LibreTexts
    Sep 12, 2022 · When total internal reflection occurs, the transmitted field is an evanescent wave; ie, a surface wave which conveys no power and whose magnitude decays ...Missing: definition | Show results with:definition
  3. [3]
    Evanescent Wave - an overview | ScienceDirect Topics
    An evanescent wave is defined as an electromagnetic field that exists only within a short distance from the surface of a material, generated during the ...
  4. [4]
    What are Evanescent Waves? - GoPhotonics.com
    Jun 12, 2023 · Evanescent waves are electromagnetic waves that exist only within a very short distance from a boundary or interface between two materials ...
  5. [5]
    5.12: Evanescent Waves - Physics LibreTexts
    May 9, 2020 · When total internal reflection occurs, the transmitted field is an evanescent wave; ie, a surface wave which conveys no power and whose magnitude decays ...
  6. [6]
    evanescent - Wiktionary, the free dictionary
    English. WOTD – 29 February 2020. Etymology. Borrowed from French évanescent (“evanescent”), from Latin ēvānēscēns (“disappearing, vanishing”), present ...
  7. [7]
    Chapter 5 The history of near-field optics - ScienceDirect.com
    For example, already in 1891 Paul Drude and Walther Nernst investigated experimentally the excitation of fluorescence by a standing evanescent wave (Drude and ...
  8. [8]
    Evanescent fields in physics and their interpretations in terms of ...
    Evanescent waves and fields play an important role in microwaves, quantum mechanics, optics and elastic waves. Because electromagnetic waves in waveguides.
  9. [9]
  10. [10]
    Self-referenced directional enhanced Raman scattering using ...
    Jan 24, 2018 · The distinct penetration depths (δ) of the evanescent field for the transverse electric (TE) and transverse magnetic (TM) modes result in ...
  11. [11]
    Total Internal Reflection Fluorescence (TIRF) Microscopy - PMC - NIH
    Depending on the excitation wavelength and objective numerical aperture, the thickness of the excitation depth, which is called the evanescent field, can be ...
  12. [12]
    Evanescent scattering imaging of single protein binding kinetics and ...
    Apr 28, 2022 · The incident angle was fixed in all experiments to ensure stable penetration depth of evanescent field (Supplementary Note 1). Traditional ...
  13. [13]
    Maxwell's Equations and the Helmholtz Wave Equation
    Maxwell's Equations and the Helmholtz Wave Equation. There are four ... Helmholtz equation: $latex \displaystyle -\nabla^2 \vec{E} = – \epsilon \mu ...
  14. [14]
  15. [15]
    Total Internal Reflection - Richard Fitzpatrick
    When total internal reflection takes place, the evanescent transmitted wave penetrates a few wavelengths into the lower refractive index medium. The ...
  16. [16]
    [PDF] Evanescent Waves
    Apr 21, 2011 · In optics and acoustics, evanescent waves are formed when waves traveling in a medium undergo total internal reflection at its boundary because ...
  17. [17]
    Total Internal Reflection Fluorescence (TIRF) Microscopy
    Formula 6 - Penetration Depth. d = λ(i)/4π × (n(1) 2sin2θ(1) - n(2)2) -1 ... evanescent field penetration depth. With the objective system, the point of ...Physical Basis of TIRFM · Basic Instrumental... · TIRFM Applications
  18. [18]
    Goos-Hänchen effect - Scholarpedia
    ### Summary of Goos-Hänchen Shift Formula
  19. [19]
    [PDF] Evanescent-wave bonding between optical waveguides
    The dispersion relation for y-odd modes (odd with respect to y=0 plane) of the two- waveguide system is shown in Figure 1(b). Modes lying outside the light cone ...
  20. [20]
    [PDF] evanescent waves and superluminal behavior of matter - arXiv
    Abstract An evanescent wave is a non-propagating wave with an imaginary wave vector. In this study, we prove that these are solutions of the tachyon-like ...
  21. [21]
    [PDF] Sensitivity enhancement in Evanescent optical waveguide sensors
    Here the sensitivity is defined as the ratio of the effective index change of the guided mode to the refractive index change of the core layer environment ...
  22. [22]
    High-sensitivity, evanescent field refractometric sensor based on a ...
    A maximum sensitivity of 1913 nm / RIU is achieved for RI ∼ 1.44 , resulting in a resolvable index change of 5.23 × 10 − 6 for a resolvable wavelength change of ...
  23. [23]
    Frustrated total internal reflection: the Newton experiment revisited
    Frustrated total internal reflection (FTIR) has been studied since the time of Newton and Fresnel. We review its history and applications in modern optics.
  24. [24]
    [PDF] ACOUSTIC REFLECTION COEFFICIENT - PSU-ETD
    we have total internal reflection as the grazing angle decreases beyond this. This. Table 5.1. Properties of the fluid media. Density(g/cm3) Sound speed(m/s).
  25. [25]
    Tunneling delay time in frustrated total internal reflection | Phys. Rev. A
    This phenomenon of frustrated total internal reflection (FTIR) is the electromagnetic analog of quantum tunneling, with the gap between the prisms serving as ...<|control11|><|separator|>
  26. [26]
    [PDF] 7.3 Dielectric Slab Waveguide - BYU
    The first is that, due to the symmetry of the geometry, the fields will either be symmetric or anti-symmetric about the x-z plane. The second is that in order ...
  27. [27]
    Slow-light and evanescent modes at interfaces in photonic crystal ...
    For example, evanescent waves enable efficient excitation of slow-light modes in photonic crystal waveguides without a transition region [5]. Moreover, all ...
  28. [28]
    [PDF] Nano-optics of surface plasmon polaritons - Duke Physics
    ... evanescent field decay above a metal surface. The resolution of SNOM mapping of the SPP field with uncoated fiber tips routinely reaches about 100 nm at the ...<|control11|><|separator|>
  29. [29]
    Evanescent Bloch waves and the complex band structure of ...
    Sep 18, 2009 · Evanescent Bloch waves are involved in the diffraction of acoustic phonons at the interfaces of finite phononic crystal structures.
  30. [30]
    [PDF] arXiv:cond-mat/0206110v2 [cond-mat.mes-hall] 10 Jan 2003
    Jan 10, 2003 · We compare the behavior of propagating and evanescent light waves in absorbing media with that of electrons in the presence of inelastic ...
  31. [31]
    Resolving the wave vector in negative refractive index media
    The evanescent regions could be absorbing or amplifying as explained in the text, depending on the signs of ϵ ′ and μ ′ . The sign of the square root for ...
  32. [32]
    Propagating and evanescent waves in absorbing media
    Jun 1, 2003 · We compare the behavior of propagating and evanescent light waves in absorbing media with that of electrons in the presence of inelastic ...
  33. [33]
    Quantitative Imaging of Rapidly Decaying Evanescent Fields Using ...
    Sep 30, 2013 · Recent studies of surface plasmon polaritons (SPPs) have led to new opportunities for manipulating light at deep sub-wavelength length scales.
  34. [34]
    Interplay between disorder-induced scattering and local field effects ...
    Local field effects are shown to increase the predicted disorder-induced scattering loss and result in significant resonance shifts of the propagating waveguide ...
  35. [35]
    (PDF) Quantitative Imaging of Rapidly Decaying Evanescent Fields ...
    Sep 11, 2013 · Non-propagating evanescent fields play an important role in the development of nano-photonic devices. While detecting the evanescent fields ...
  36. [36]
    Superlens in the Time Domain | Phys. Rev. Lett.
    We find that the resolution can be improved if the pulse duration is shorter than the surface wave decay time. At first glance, the idea of improving spatial ...
  37. [37]
    Super-shear evanescent waves for non-contact elastography of soft ...
    In our study, we analyze the Green's function for a pulsed excitation to define SEW propagation speed and produce complementary numerical simulations and ...
  38. [38]
    Self-phase-modulation of surface plasmon polaritons | Phys. Rev. A
    Jan 31, 2014 · We present an approach for calculating the nonlinear propagation of surface plasmon polaritons in one-dimensional planar waveguides ...
  39. [39]
    Self-phase modulation-induced modulation instability in silicon-on ...
    We present an investigation of the impact of two-photon absorption and free-carrier effects on modulation instability (MI) in silicon-on-insulator (SOI) ...Missing: intensity altering
  40. [40]
    Atom trap and waveguide using a two-color evanescent light field ...
    Dec 7, 2004 · We suggest using a two-color evanescent light field around a subwavelength-diameter fiber to trap and guide atoms. The optical fiber carries ...
  41. [41]
    Atom-light interactions using optical nanofibres—a perspective
    Apr 18, 2024 · To control atom–photon coupling, two-colour evanescent fields are used for atom trapping near the nanofibre surface [51–53]. This ...
  42. [42]
    Calibrating Evanescent-Wave Penetration Depths for Biological ...
    Sep 3, 2019 · Direct measurement of the evanescent field profile produced by objective-based total internal reflection fluorescence. J. Biomed. Opt, 11 ...
  43. [43]
    Direct measurement of the evanescent field profile produced by ...
    We present an experimental method, where the depth and axial profile of the evanescent field can be measured directly by microscopic observation.
  44. [44]
    [PDF] Coupled mode theory and coupled mode photonic devices: A Review
    Coupling is broadly of two types, the evanescent field coupling between modes of two adjacent waveguides or fibres and coupling between modes of the same fibre ...
  45. [45]
  46. [46]
    Highly selective evanescent modal filter for two-mode optical fibers
    Efficient and selective coupling between a single-mode fiber and the LP11 mode of a double-mode fiber is demonstrated in an evanescent directional coupler.
  47. [47]
    Characterization of the evanescent field profile and bound mass ...
    ... exponential decay constant. As described previously, the shift in resonance wavelength is proportional to the intensity of the evanescent field extending ...
  48. [48]
    Sensitivity Comparison of Surface Plasmon Resonance and ... - NIH
    For silver-based SPR sensors (Ag-SiO2-6nm), a change of 0.01 refractive index units (RIU) induced a shift in the resonance angle by 0.68 degrees, which is 1.8 ...
  49. [49]
    Enhanced sensitivity of a surface plasmon resonance biosensor ...
    With the light incident angle of 75° and 66°, we experimentally demonstrated that the proposed SPR sensor could achieve the average sensitivity of 3341 nm/RIU ...
  50. [50]
    Etched optical fiber for measuring concentration and refractive index ...
    Sep 13, 2018 · The study of the reduction of an optical fiber by chemical etching has been suggested to determine the concentrations of sucrose in water and their refractive ...
  51. [51]
    [PDF] Etched optical fiber for measuring concentration and refractive index ...
    In the present work, an optical fiber sensor reduced by chemi- cal etching was developed to measure different concentrations of sucrose by evanescent waves. The ...
  52. [52]
    Distributed Refractive Index Sensing Based on Etched Ge-Doped ...
    Apr 28, 2023 · A distributed optical fiber refractive index sensor based on etched Ge-doped SMF in optical frequency domain reflection (OFDR) was proposed and demonstrated.
  53. [53]
    Study on the limit of detection in MZI-based biosensor systems
    Apr 8, 2019 · Mach-Zehnder interferometers are integrated photonic sensors that have yielded excellent detection limits down to 10 −7 RIU.
  54. [54]
    Highly Sensitive Protein Detection by Asymmetric Mach–Zehnder ...
    Jun 11, 2020 · We demonstrate a highly sensitive label-free Mach-Zehnder interferometer (MZI) biosensor based on silicon nitride slot waveguide. Unlike the ...
  55. [55]
    Label-free real-time optical monitoring of DNA hybridization using ...
    Dec 21, 2018 · Label-free real-time optical monitoring of DNA hybridization using SiN Mach–Zehnder interferometer-based integrated biosensing platform.<|separator|>
  56. [56]
    Evanescent field Sensors Based on Tantalum Pentoxide Waveguides
    The evanescent field with a penetration depth of about 200 nm into the adjacent medium ensures that only fluorophores close to the surface are excited, which ...
  57. [57]
    Intrinsic structure of biological layers: Vertical inhomogeneity profiles ...
    ... penetration depth of the so-called evanescent field (typically up to 100–200 nm above the surface). Usually, the effective refractive index of the ...
  58. [58]
    Evanescent Wave Fluorescence Optical Biosensors
    The evanescent field typically extends only 100-200 nm from the waveguide surface, limiting the sensing volume and potentially reducing signal strength for ...
  59. [59]
    Evanescent wave fluorescence biosensors: Advances of the ... - NIH
    The evanescent wave, first described by (Hirschfeld 1966) arises from the manner in which light behaves when confined in an optical waveguide. Guided light is ...
  60. [60]
    [PDF] Interferometric Evanescent Wave Biosensor Principles and ...
    Biosensors based on evanescent field detection have shown to be very good candidates for point of care devices due to their extreme sensitivity for label-free ...
  61. [61]
    Evanescent wave-based optical biosensors for innovations, medical ...
    Jul 7, 2025 · EW biosensors exhibit outstanding sensing performance for both bulk sensing and surface sensing [8,9].
  62. [62]
    Recent advancements in microfluidic-based biosensors for detection ...
    Recent advancements include improved biomarker detection, enhanced POC testing speed, sophisticated fabrication, exosome use, and hybrid integration for  ...
  63. [63]
    Recent Advances in Microfluidics-Integrated Optical Biosensors
    Label-free detection; · All optical biosensors; · Integrated optical sensory devices; · Evanescent field-based biosensors (ATR); · Surface plasmon resonance sensors ...Missing: post- | Show results with:post-
  64. [64]
    Highly efficient and selective integrated directional couplers ... - Nature
    Dec 20, 2023 · The length of evanescent mode coupling can be estimated as Lc = (π/(2К))·(2n + 1), where Lc is the coupling length, at a distance the power ...
  65. [65]
  66. [66]
  67. [67]
  68. [68]
    Mid-infrared silicon photonic waveguides and devices [Invited]
    It is noted that the propagation loss of strip waveguide is significantly larger than that of a rib waveguide due to more interactions with etch-induced side- ...
  69. [69]
    Roadmapping the next generation of silicon photonics - Nature
    Jan 25, 2024 · We chart the generational trends in silicon photonics technology, drawing parallels from the generational definitions of CMOS technology.
  70. [70]
    Hybrid integration methods for on-chip quantum photonics
    Aers, and R. L. Williams, “On-chip integration of single photon sources via evanescent coupling of tapered nanowires to SiN waveguides,” Adv. Quantum Technol.
  71. [71]
    Total Internal Reflection Fluorescence (TIRF) Microscopy
    Mar 11, 2012 · TIRF is a special technique in fluorescence microscopy developed by Daniel Axelrod at the University of Michigan, Ann Arbor in the early 1980s.
  72. [72]
    Total internal reflection fluorescence microscopy in cell biology
    This review describes a microscopy technique based on total internal reflection fluorescence which is well suited for optical sectioning at cell-substrate ...Missing: 1979 paper
  73. [73]
  74. [74]
    Evanescent wave imaging in optical lithography - SPIE Digital Library
    The use of evanescent wave lithography (EWL) has been employed for 26nm resolution at 1.85NA using a 193nm ArF excimer laser wavelength to record images in a ...<|control11|><|separator|>
  75. [75]
    Chapter 7: Total internal reflection fluorescence microscopy - PubMed
    Total internal reflection fluorescence microscopy (TIRFM), also known as evanescent wave microscopy, is used in a wide range of applications.Missing: 1979 | Show results with:1979