Fact-checked by Grok 2 weeks ago

Transverse wave

A transverse wave is a wave in which the oscillations occur to the direction of . In waves, this involves the of the medium's particles to the wave's , resulting in oscillations at right angles to the path. transverse waves are characterized by their ability to propagate through solids and on surfaces, but they cannot travel through the bulk of fluids like gases or liquids due to the absence of a restoring force for motion. Common examples include waves on a stretched , where particles move up and down as the wave travels along the ; surface ripples on ; seismic S-waves, which cause deformation in the during earthquakes and travel only through solids; and electromagnetic waves, such as and radio waves, where the oscillating electric and magnetic fields are to each other and to the direction of . In contrast to longitudinal waves, transverse waves exhibit properties like , where the plane of oscillation can be restricted, making them essential in fields such as and .

Fundamentals

Definition

A transverse wave is a type of wave in which the oscillations occur perpendicular to the direction of wave propagation, thereby transferring energy without net displacement in the propagation direction. This includes mechanical transverse waves, where the particles of the medium vibrate perpendicular to the wave's direction, and electromagnetic waves, where the electric and oscillate perpendicular to each other and to the propagation direction. In mechanical transverse waves, the disturbance causes individual particles to vibrate back and forth—such as or side to side—in a plane orthogonal to the forward progress of the wave itself, forming a repeating that advances through the medium over time. This perpendicular motion distinguishes transverse waves as a fundamental mode of propagation, enabling phenomena where the wave's energy travels independently of the medium's bulk movement. This perpendicularity of oscillation to propagation contrasts with longitudinal waves, in which particle motion aligns parallel to the wave's direction.

Key Characteristics

Transverse waves are characterized by several fundamental properties that describe their oscillatory behavior and propagation. The amplitude represents the maximum extent of , such as the displacement of the medium's particles from their position in waves or the peak in electromagnetic waves, determining the wave's intensity or carried. The wavelength (\lambda) is the spatial between consecutive crests or troughs of the wave, providing a measure of its spatial periodicity. The frequency (f) quantifies the number of oscillations or per unit time, typically measured in hertz (Hz), while the period (T) is the duration of one complete , inversely related to by T = 1/f. The phase indicates the specific of a point on the wave relative to a reference point in its , often expressed as the argument of the sinusoidal describing the wave, such as \phi = kx - \omega t, where it helps determine between different wave components. A defining feature of transverse waves is the directionality of , perpendicular to the direction of wave . This perpendicularity distinguishes them from other wave types and allows for motion confined to a perpendicular to (linear transverse motion) or more complex paths, such as elliptical or circular trajectories in three dimensions when combining components along the two independent perpendicular axes. The speed (v) of a transverse wave depends on the properties of the medium through which it travels (or the for electromagnetic waves) and relates the other characteristics via the relation v = f \lambda, where the wave speed remains constant for a given medium while and adjust inversely to maintain this balance. These properties collectively enable phenomena like , where the orientation of oscillations influences wave behavior.

Comparison with Longitudinal Waves

Transverse waves differ fundamentally from longitudinal waves in the direction of relative to the direction of wave propagation. In transverse waves, oscillations occur perpendicular to the propagation direction, resulting in crests and troughs, whereas in longitudinal waves, oscillations occur parallel to the propagation direction, producing compressions and rarefactions. This distinction arises because transverse waves require a medium with to support perpendicular motion, such as solids or taut strings, while longitudinal waves can propagate through media lacking such rigidity, including fluids like or . Electromagnetic transverse waves, however, do not require a medium. Despite these differences, both transverse and longitudinal waves share core similarities as disturbances that propagate without net transport of matter. They exhibit common properties, including , , , and speed, which determine their behavior in and . The following table summarizes key contrasts in and medium requirements (noting that electromagnetic transverse waves do not require a medium):
AspectTransverse Waves (Mechanical)Longitudinal Waves
Oscillation DirectionPerpendicular to Parallel to
Oscillation PathUp-and-down or side-to-side (circular or linear in )Back-and-forth along axis
Medium RequirementNeeds (e.g., solids, strings)No needed (e.g., fluids, gases)
Propagation CapabilityPossible in solids; limited in fluidsPossible in solids, fluids, and gases
Occurrence ExamplesVibrations on strings, seismic S-wavesSound waves in air, seismic P-waves

Examples in Nature and Technology

Mechanical Transverse Waves

Mechanical transverse waves are disturbances in a medium where particles oscillate to the of wave propagation, requiring the medium to provide restoring forces that act transversely to the . These waves necessitate a medium capable of sustaining , such as solids or tensed strings, where elasticity or tension supplies the restoring force; in contrast, fluids without boundaries cannot support pure transverse waves due to the lack of such resistance. A classic example is the wave on a tensed , as seen in the vibration of a guitar , where plucking causes transverse oscillations that propagate along the 's length while the itself moves up and down perpendicular to that direction. Another prominent instance occurs during earthquakes with seismic S-, which are transverse that cause ground particles to move horizontally or vertically relative to the 's propagation path through Earth's solid crust. Surface on , such as ripples, exhibit transverse components where particles primarily move in circular orbits with vertical displacements dominating near the surface, combining with minor longitudinal motion. A straightforward experimental involves securing one end of a long or string to a fixed point and shaking the free end up and down to generate traveling transverse waves that propagate along the , allowing observation of wave speed variations with tension or . This setup can also produce standing waves by adjusting the shaking to match the rope's natural resonances, illustrating and antinode patterns.

Electromagnetic Transverse Waves

Electromagnetic waves consist of oscillating electric and magnetic fields that are perpendicular to each other and to the direction of wave propagation, making them inherently transverse. These fields vary sinusoidally in phase, with the electric field \mathbf{E} inducing the magnetic field \mathbf{B} and vice versa through mutual interaction, enabling self-propagation without a medium. In vacuum, all electromagnetic waves travel at the constant speed of light c \approx 3.00 \times 10^8 m/s, determined by the fundamental constants of permittivity \epsilon_0 and permeability \mu_0 of free space. Prominent examples include visible light, which spans wavelengths from about 400 to 700 nm and is responsible for human vision as well as in plants; radio waves, with wavelengths ranging from millimeters to kilometers used in communication technologies; and X-rays, featuring short wavelengths on the order of 0.01 to 10 nm, employed in and material analysis. Each of these propagates as a transverse wave in vacuum at speed c, carrying across the without requiring a physical medium. The transverse exclusivity of electromagnetic waves arises from Maxwell's equations, particularly Gauss's laws, which in free space (with no charges or currents) require the divergence of both \mathbf{E} and \mathbf{B} to be zero (\nabla \cdot \mathbf{E} = 0, \nabla \cdot \mathbf{B} = 0). For plane waves, this implies no longitudinal components along the direction, as such components would produce nonzero , violating the source-free conditions; thus, only transverse oscillations can sustain . This fundamental constraint ensures that electromagnetic waves maintain their perpendicular field structure throughout .

Mathematical Description

Wave Equation

The wave equation provides the fundamental mathematical framework for describing the propagation of transverse waves in a one-dimensional medium, such as a taut . It is a that relates the transverse y(x, t) to its spatial and temporal variations: \frac{\partial^2 y}{\partial t^2} = v^2 \frac{\partial^2 y}{\partial x^2}, where x is the along the direction of , t is time, and v is the constant speed of the wave. This second-order arises from the physics of wave motion and applies to small-amplitude transverse disturbances where the displacement is to the propagation direction. The derivation begins by applying Newton's second law to an infinitesimal element of the medium, typically a with uniform linear density \mu ( per unit ) under constant T. For a small of \Delta x, the is \mu \Delta x, and the net transverse force is the difference in the vertical components of at the ends, approximated as T \frac{\partial^2 y}{\partial x^2} \Delta x for small slopes. This force equals times transverse acceleration \mu \Delta x \frac{\partial^2 y}{\partial t^2}, yielding the wave equation upon dividing by \Delta x and taking the limit \Delta x \to 0. The wave speed emerges as v = \sqrt{T / \mu}, reflecting how higher increases speed while greater density decreases it./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/16%3A_Waves/16.04%3A_Wave_Speed_on_a_Stretched_String) A common solution to this equation for monochromatic waves is the sinusoidal form y(x, t) = A \sin(kx - \omega t + \phi), where A is the , k = 2\pi / \lambda is the wave number (\lambda being the ), \omega = 2\pi f is the (f being the ), and \phi is the constant. This traveling wave solution satisfies the equation because the dispersion relation \omega = v k holds, ensuring consistency between spatial and temporal derivatives.

Polarization States

Polarization in transverse waves refers to the orientation of the oscillations to the direction of propagation. For waves in two or three dimensions, such as electromagnetic waves, the or vector can oscillate in various patterns, leading to different states. These states are determined by the relative amplitudes and differences of the components along perpendicular axes. Linear polarization occurs when the oscillation is confined to a fixed plane, such as vertical or relative to the propagation direction. In this state, the vector of an electromagnetic vibrates back and forth along a straight line, with no phase difference between any components. For example, a linearly polarized propagating in the z-direction can have its along the x-axis, described by \vec{E} = E_0 \hat{x} \cos(kz - \omega t). This is the simplest form of and is commonly produced using polarizers that transmit only one orientation of the field. Circular and elliptical polarization arise from the superposition of two perpendicular linear oscillations with a phase difference, typically \pi/2 for circular cases. In circular polarization, the amplitudes of the perpendicular components are equal, and the electric field vector rotates at a constant angular speed, tracing a circle in the plane perpendicular to propagation. The handedness—right-handed or left-handed—is defined by the rotation direction when looking toward the source: clockwise for right-circular and counterclockwise for left-circular. For a right-handed circularly polarized electromagnetic wave propagating in the +z direction, the electric field components are given by E_x = A \cos(kz - \omega t), \quad E_y = A \sin(kz - \omega t), where A is the amplitude. Left-handed polarization reverses the y-component sign: E_y = -A \sin(kz - \omega t). These can be conceptually represented using Jones vectors, such as \begin{pmatrix} 1 \\ i \end{pmatrix} for left-circular and \begin{pmatrix} 1 \\ -i \end{pmatrix} for right-circular (normalized). Elliptical polarization is a generalization where the amplitudes differ (A ≠ B) or the phase difference is not exactly \pi/2, resulting in the field tracing an ellipse; for instance, unequal A and B in the above equations produce an ellipse tilted along the major axis. In electromagnetic waves, particularly , polarization states are crucial because natural from most sources is unpolarized, meaning it consists of a random of all possible orientations with no fixed phase relationships. However, interactions with matter can select or alter these states; for example, polarizers transmit , while birefringent materials like quarter-wave plates convert linear to by introducing the necessary phase shift. These effects are fundamental in , enabling applications such as reducing glare in or analyzing molecular structures in .

Superposition Principle

The superposition principle governs the behavior of in linear , stating that the total at any point is the sum of the displacements from each individual wave, provided the medium responds linearly to the applied forces. For , where displacements occur to the of , this addition applies in the plane of , assuming the waves are polarized in the same as relevant to their states. This principle leads to interference patterns when transverse waves overlap. Constructive interference occurs when the waves are in phase, meaning their phase difference is an integer multiple of $2\pi, resulting in maxima where the amplitudes add to produce a displacement up to twice that of a single wave. Destructive interference happens when the waves are out of phase by an odd multiple of \pi, causing minima where the amplitudes cancel, potentially reducing the displacement to zero. A mathematical illustration of superposition for two identical transverse waves with A, wave number k, \omega, and phase difference \delta is given by: y_1 = A \sin(kx - \omega t) y_2 = A \sin(kx - \omega t + \delta) The displacement y is: y = y_1 + y_2 = 2A \cos\left(\frac{\delta}{2}\right) \sin\left(kx - \omega t + \frac{\delta}{2}\right) Here, the of the wave is $2A \cos(\delta/2), which reaches a maximum of $2A for \delta = 0 (constructive) and zero for \delta = \pi (destructive). In applications to transverse waves, such as electromagnetic waves, the produces observable fringes in , as seen in Young's double-slit experiment where coherent waves from two slits interfere to form alternating bright and dark bands on a screen due to path length differences causing phase shifts. Analogously, though sound waves are longitudinal, the principle similarly yields beats when frequencies differ slightly, a phenomenon that parallels intensity modulations in superposed transverse waves like .

Energy and Power Transmission

In transverse waves, energy is transported through the medium via oscillatory motion, divided equally between kinetic and potential forms on average. For a sinusoidal transverse wave propagating along a stretched string, the kinetic energy arises from the transverse velocity of string elements, while the potential energy stems from the stretching of the string beyond its equilibrium length. These energies fluctuate over each cycle, but their time-averaged total per unit length, known as the average energy density u, is given by u = \frac{1}{2} \mu \omega^2 A^2, where \mu is the linear mass density of the string, \omega is the angular frequency, and A is the amplitude. This equality between average kinetic and potential energies holds because, at any point, the maximum kinetic energy occurs when potential is zero, and vice versa, leading to equipartition over the wave period. The power transmitted by such a wave, representing the rate of energy flow past a point on the string, is the product of the and the wave speed v. The time-averaged power P for a sinusoidal wave is thus P = \frac{1}{2} \mu v \omega^2 A^2. This formula quantifies how energy propagates along the string at speed v = \sqrt{T/\mu}, where T is the , without any net of the medium's . Electromagnetic transverse waves, such as light, carry energy through oscillating electric and magnetic fields in vacuum or media. The instantaneous energy flux is described by the Poynting vector \mathbf{S} = \frac{1}{\mu_0} \mathbf{E} \times \mathbf{B}, where \mathbf{E} and \mathbf{B} are the electric and magnetic field vectors, and \mu_0 is the permeability of free space; its magnitude points in the direction of propagation and gives the power per unit area./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.04%3A_Energy_Carried_by_Electromagnetic_Waves) For a plane sinusoidal wave, the time-averaged intensity I, or power per unit area, simplifies to I = \frac{1}{2} c \epsilon_0 E_0^2, where c is the speed of light, \epsilon_0 is the permittivity of free space, and E_0 is the electric field amplitude (with B_0 = E_0 / c)./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.04%3A_Energy_Carried_by_Electromagnetic_Waves) This intensity measures the radiant energy flow, analogous to power in mechanical waves. In both and electromagnetic transverse waves, occurs without net transport of ; particles or disturbances oscillate locally around equilibrium positions while the wave pattern advances, transferring from one region to another./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.04%3A_Energy_Carried_by_Electromagnetic_Waves)

Applications and Phenomena

Wave Propagation in Media

When a transverse wave encounters an between two media, part of the wave is back into the first medium, while the remainder is transmitted into the second medium. The law of states that the angle of incidence equals the angle of , ensuring that the reflected wave's direction lies in the defined by the incident wave and the normal to the . This behavior arises from the conditions that maintain of the wave's and the transverse component of the wave's across the . For oblique incidence, the transmitted wave refracts according to , which relates the angles of incidence and to the wave speeds in the respective media: n_1 \sin \theta_1 = n_2 \sin \theta_2, where n_1 and n_2 are the refractive indices (inversely proportional to the velocities) of the first and second media, \theta_1 is the angle of incidence, and \theta_2 is the angle of . This law ensures matching along the , preventing discontinuities in the wave front. In transverse waves on strings with different linear densities but the same , an analogous occurs, bending the wave path toward the normal when entering a denser medium. At such boundaries, the amplitudes of the reflected and transmitted waves depend on the impedance mismatch between the media. For mechanical transverse waves on strings, the impedance is defined as Z = T / v, where T is and v is wave speed. For a wave incident from medium 1 to medium 2, the r (ratio of reflected to incident amplitude) is r = (Z_1 - Z_2)/(Z_1 + Z_2), while the t = 2 Z_1 / (Z_1 + Z_2). For electromagnetic waves, the intrinsic impedance is \eta = \sqrt{\mu / \varepsilon}, where \mu is permeability and \varepsilon is , and the coefficients for the amplitude are r = (\eta_2 - \eta_1)/(\eta_2 + \eta_1), t = 2 \eta_2 / (\eta_2 + \eta_1). These coefficients determine the energy partitioning, with no net energy loss at ideal boundaries. The boundary conditions for transverse waves on joined strings specifically require of transverse (to avoid breaking) and of the transverse , which equates to the times the (T \partial y / \partial x) being equal on both sides. If tensions differ, the slopes adjust accordingly to balance forces. In dispersive media, the propagation speed of transverse waves varies with , causing different components of a to travel at different velocities. This frequency-dependent phase velocity v_p(\omega) leads to , where a initially narrow spreads out over distance as its components separate. For electromagnetic transverse waves in dielectrics, arises from the frequency dependence of the n(\omega), often modeled by the based on atomic resonances. In non-ideal media, this effect limits in applications like optical fibers, where broadening can degrade information transmission. Attenuation in transverse wave propagation occurs through energy loss mechanisms such as absorption, where wave energy converts to heat via material damping, or scattering, where inhomogeneities redirect energy away from the primary propagation direction. In absorbing media, the wave amplitude decays exponentially as e^{-\alpha z}, with attenuation coefficient \alpha increasing with frequency due to resonant interactions. For electromagnetic transverse waves, intrinsic absorption follows the imaginary part of the dielectric constant, while scattering dominates in turbid media like biological tissues. In mechanical contexts, such as transverse waves in viscoelastic solids, attenuation results from internal friction, broadening wave pulses and reducing peak amplitudes over distance. These losses distinguish real media from ideal non-dispersive, lossless cases.

Interference and Diffraction

Interference in transverse waves arises from the superposition principle, where waves from multiple sources or paths combine to produce regions of constructive and destructive interference. In Young's double-slit experiment, monochromatic light passes through two closely spaced slits separated by distance d, creating two coherent sources that produce an interference pattern on a distant screen. The path difference \delta between waves from the two slits to a point on the screen at angle \theta is given by \delta = d \sin \theta. Constructive interference, resulting in bright fringes, occurs when this path difference is an integer multiple of the wavelength: m\lambda = \delta, where m = 0, 1, 2, \dots. Destructive interference, producing dark fringes, happens when \delta = (m + 1/2)\lambda. This pattern demonstrates the wave nature of light, a transverse electromagnetic wave, with fringe spacing depending on wavelength, slit separation, and distance to the screen. Diffraction refers to the bending and spreading of around obstacles or through apertures, also explained by the Huygens-Fresnel principle, which posits that every point on a acts as a source of secondary spherical wavelets that with each other. For a single slit of width a illuminated by monochromatic , the diffraction pattern consists of a central bright maximum flanked by minima. The condition for minima is derived from the path differences across the slit, yielding \sin \theta = m \lambda / a, where m = \pm 1, \pm 2, \dots, leading to destructive at those angles. The intensity distribution follows a sinc-squared , with the central maximum's width inversely proportional to a. In transverse waves like , plays a key role in ; for instance, the orientation relative to the slit or obstacle affects the amplitude and pattern, as the wave's transverse nature couples the field components to the geometry. Polarization effects are particularly evident in diffraction gratings used in spectrometers, where multiple slits enhance angular . The grating equation d (\sin \theta_i + \sin \theta_m) = m \lambda determines diffraction orders, but efficiency varies with : for grooves parallel to the (TE mode), diffraction is stronger than for perpendicular (TM mode), leading to polarization-dependent and blaze angles optimized for specific orientations. This is crucial in applications like , where from stars requires accounting for instrumental to avoid biases in measurements. A vivid example of interference in transverse waves is the colorful patterns on soap bubbles, resulting from . Light reflects from both the inner and outer surfaces of the soap film's water-air , with a path difference of $2nt \cos \theta (where n is the , t the thickness, and \theta the incidence ), plus a \pi shift at one interface. Constructive for reflected occurs for wavelengths satisfying $2 n t \cos \theta = (m + 1/2) \lambda, producing iridescent colors that shift as the film thins or drains. X-ray diffraction in crystals provides another key phenomenon, exploiting the transverse wave properties of s to probe structures. X-rays scatter off successive planes in a crystal , interfering constructively when the path difference satisfies : $2d \sin \theta = m \lambda, where d is the interplanar spacing. This produces discrete spots or rings, enabling determination of crystal symmetries and parameters, as in the analysis of minerals or proteins. The transverse of X-rays influences scattering intensities, with certain orientations enhancing or suppressing reflections based on electron distributions.

Historical Development

The concept of transverse waves emerged within the broader development of wave theories for and other phenomena, beginning with early attempts to explain without relying on particle models. In 1678, proposed a wave theory of in his unpublished manuscript Traité de la Lumière, later published in 1690, describing as longitudinal pressure propagating through an elastic similar to sound . This idea faced significant rejection in the late 17th and 18th centuries, primarily due to Isaac Newton's influential corpuscular theory, which better aligned with observed and at the time. However, the longitudinal nature of Huygens' model struggled to account for later discoveries like , as such in a solid-like could not easily support the observed transverse restrictions without invoking problematic properties in the medium. The revival of the wave theory in the early 19th century, driven by Thomas Young and , marked a pivotal shift toward recognizing as transverse . Young's double-slit interference experiments in 1801 demonstrated wave superposition, but it was Fresnel's work on and from 1815 to 1819 that provided compelling evidence for transverse vibrations. In a 1818 memoir to the , Fresnel explained polarization phenomena—such as the inability of certain orientations to pass through polarizing crystals—by modeling as transverse oscillations perpendicular to the direction of propagation, with the possessing the necessary elasticity for such motions. This transverse hypothesis resolved inconsistencies in longitudinal models and was experimentally confirmed through predictions like the diffraction patterns observed by in 1818. James Clerk Maxwell's unification of , , and in the 1860s solidified the transverse wave nature of within . In his 1865 paper "A Dynamical Theory of the Electromagnetic Field," Maxwell derived equations showing that electric and magnetic fields oscillate perpendicular to each other and to the propagation direction, forming self-sustaining transverse waves traveling at the through vacuum, without needing an for propagation. This theoretical framework predicted electromagnetic waves across the , later verified by Heinrich Hertz's experiments in 1887, establishing transverse electromagnetic waves as a cornerstone of . In the 20th century, the understanding of transverse waves extended to and . Albert Einstein's 1905 explanation of the introduced light quanta, or photons, as discrete packets of electromagnetic energy inheriting the transverse properties of classical waves. Concurrently, in , Richard Dixon Oldham's analysis of records in 1906 classified secondary () waves as transverse shear waves, distinct from primary () longitudinal waves, providing early evidence for Earth's layered interior as S-waves failed to traverse the liquid outer core. These advancements maintained a classical foundation while integrating transverse wave concepts into , where photons exhibit two transverse states.

References

  1. [1]
    Transverse and Longitudinal Waves - HyperPhysics Concepts
    Transverse waves cannot propagate in a gas or a liquid because there is no mechanism for driving motion perpendicular to the propagation of the wave.
  2. [2]
    16.1 What Are Waves? - Front Matter
    A transverse wave is a wave in which the displacement of particles in the medium is perpendicular to the direction that the disturbance travels. The figure ...
  3. [3]
    Seismic Waves - HyperPhysics
    S waves are transverse waves which involve movement of the ground perpendicular to the velocity of propagation. They travel only through solids, and the absence ...
  4. [4]
    Nats S04-13
    Electromagnetic radiation is a transverse wave, composed of two parts, an electric and a magnetic component. Both are oriented 90° from each other. Each wave ...
  5. [5]
    Transverse Waves - UW–Madison Physics
    When a moving wave consists of oscillations occurring perpendicularly to the direction of energy transfer, it is called a Transverse Wave.
  6. [6]
    Waves - Physics
    Mar 1, 1999 · In a transverse wave, the motion of the particles of the medium is at right angles (i.e., transverse) to the direction the wave moves. In a ...
  7. [7]
    What is a Wave? - Graduate Program in Acoustics
    Jul 11, 2025 · Definition of a Wave​​ Webster's dictionary defines a wave as: a disturbance or variation that transfers energy progressively from point to point ...
  8. [8]
    16.2 Mathematics of Waves – University Physics Volume 1
    We described periodic waves by their characteristics of wavelength, period, amplitude, and wave speed of the wave.Missing: key | Show results with:key
  9. [9]
    [PDF] CHAPTER 16 - Waves
    • Describe the basic characteristics of wave motion. • Define the terms wavelength, amplitude, period, frequency, and wave speed. • Explain the difference ...
  10. [10]
    Waves – ISP209: The Mystery of the Physical World
    Wave velocity and wavelength are related to the wave's frequency and period by v w = λ T or. A transverse wave has a disturbance perpendicular to its direction ...
  11. [11]
    [PDF] CHAPTER 16: TRAVELING WAVES
    For any sinusoidal wave (e.g. 1D):. The wave phase determines where you are on the wave, i.e. a peak or trough, or somewhere in between.
  12. [12]
    [PDF] Chapter 16 WAVES - UNIVERSITY PHYSICS
    Page 4. FIGURE 16.4. (a) In a transverse wave, the medium oscillates perpendicular to the wave velocity. Here, the spring moves vertically up and down, while ...
  13. [13]
    Types of Mechanical Waves
    Transverse: medium pieces move perpendicular to wave motion. Circular Transverse: medium pieces move in circles around wave motion direction. Circular ...Missing: particle | Show results with:particle
  14. [14]
    Wave Motion in Mechanical Medium - Graduate Program in Acoustics
    There are two basic types of wave motion for mechanical waves: longitudinal waves and transverse waves. The animations below demonstrate both types of wave.
  15. [15]
    [PDF] 1.2 Waves - Physics Courses
    The speed of the transverse and longitudinal waves are different. • Longitudinal waves but not transverse waves can propagate in a fluid. Examples. • Transverse ...
  16. [16]
    [PDF] Mechanical waves - Duke Physics
    For mechanical waves the formula for v has a generic form: v = Stiffness/Inertia .
  17. [17]
    16.3 Wave Speed on a Stretched String – University Physics Volume 1
    If you pluck a string under tension, a transverse wave moves in the positive x-direction, as shown in Figure. The mass element is small but is enlarged in the ...
  18. [18]
    S-wave Motion- Incorporated Research Institutions for Seismology
    Particle motion is perpendicular to the direction of propagation (transverse). Transverse particle motion shown here is vertical but can be in any direction.
  19. [19]
    16.1 Traveling Waves – University Physics Volume 1
    Examples of transverse waves are the waves on stringed instruments or surface waves on water, such as ripples moving on a pond. Sound waves in air and water are ...
  20. [20]
    Spring Waves | Swarthmore Physics Demonstrations
    Jun 16, 2015 · Experiment with transverse and longitudinal wave forms; tie rope to long spring to demonstrate the effect of a change in medium on wave speed.
  21. [21]
  22. [22]
    Electromagnetic Spectrum - Introduction - Imagine the Universe!
    The electromagnetic spectrum is the range of all EM radiation, including radio, microwaves, infrared, visible, ultraviolet, X-rays, and gamma-rays.
  23. [23]
    [PDF] Derivation of the Wave Equation
    In these notes we apply Newton's law to an elastic string, concluding that small amplitude transverse vibrations of the string obey the wave equation.
  24. [24]
    [PDF] Transverse waves on a string
    In section 4.1 we derive the wave equation for transverse waves on a string. This equation will take exactly the same form as the wave equation we derived for ...
  25. [25]
    Traveling waves
    The quantity kx - ωt + φ is called the phase. At a fixed position x the displacement y varies as a function of time as A sin(ωt) = A sin[(2π/T)t] with a ...
  26. [26]
    [PDF] Chapter 16 Waves in One Dimension
    Show that a sinusoidal traveling wave of the form f (x, t) = A sin(kx – ωt) satisfies the wave equation for any value of k and ω. © 2015 Pearson Education, Inc.
  27. [27]
    [PDF] Chapter 12 - Polarization - MIT OpenCourseWare
    We describe the varieties of possible polarization states of a plane wave: linear, circular and elliptical. iii. We describe “unpolarized light,” and ...
  28. [28]
    Classification of Polarization - HyperPhysics
    A plane electromagnetic wave is said to be linearly polarized. The transverse electric field wave is accompanied by a magnetic field wave as illustrated.Missing: states | Show results with:states
  29. [29]
    [PDF] Lecture 14: Polarization
    Thus circular and linear polarizations are not linearly independent. Indeed, any possible polarization can be written as a linearly combination of left-handed ...
  30. [30]
    16.10 Superposition and Interference - College Physics 2e | OpenStax
    Jul 13, 2022 · The superposition of most waves produces a combination of constructive and destructive interference and can vary from place to place and time ...
  31. [31]
  32. [32]
    16.4 Energy and Power of a Wave - University Physics Volume 1
    Sep 19, 2016 · Begin with the equation of the time-averaged power of a sinusoidal wave on a string: P = 1 2 μ A 2 ω 2 v . P = 1 2 μ A 2 ω 2 v . The amplitude ...
  33. [33]
    [PDF] Refraction and Snell's law - MIT OpenCourseWare
    Refraction involves a change in the direction of wave propagation due to a change in propagation speed. It involves the oblique incidence of waves on media ...
  34. [34]
    [PDF] Traveling Waves and Boundary Interactions - MIT OpenCourseWare
    1. Continuity: The string must be continuous across the interface (Eq. (21)). This means the displacement y(x, t) when approaching the interface from the left ...
  35. [35]
    [PDF] Lecture 9: Reflection, Transmission and Impedance
    Now we have the boundary conditions. What is the solution? 2 Reflection and transmission. Suppose we have some incoming traveling wave. Before it hits the ...
  36. [36]
    Waves, Superposition, Dispersion - UMD Physics
    May 7, 2024 · This is important to understand: the frequency (and period) are determined by the source of the wave, and the wavelength comes from properties ...Missing: key characteristics
  37. [37]
    [PDF] Chapter 7. Electromagnetic Wave Propagation
    This chapter covers electromagnetic waves, starting with plane waves, then nonuniform systems, guided waves, and finally resonant cavities and energy ...
  38. [38]
    27.3 Young's Double Slit Experiment – College Physics
    destructive interference for a double slit: the path length difference must be a half-integral multiple of the wavelength. incoherent: waves have random phase ...
  39. [39]
    [PDF] Lecture 36 – Diffraction - Purdue Physics
    Huygens-Fresnel Principle​​ Huygens: – Every point on a wave front acts as a point source of secondary spherical waves that have the same phase as the original ...
  40. [40]
    [PDF] Chapter 36 - NJIT
    single-slit diffraction pattern with slit width a. • The diffraction minima are labeled by the integer m d. = ±1, ±2, … (“d” for. “diffraction”). • Figure (b) ...<|control11|><|separator|>
  41. [41]
    Measurement of the instrumental polarization of a high resolution ...
    Diffraction grating spectrometers exhibit a complex dependance on the polarization state of incident light. We have characterized the polarization ...
  42. [42]
    Interference Phenomena in Soap Bubbles: Interactive Java Tutorial
    Jun 14, 2018 · A good example of interference is the thin film of a soap bubble, which reflects a spectrum of beautiful colors when illuminated by natural ...
  43. [43]
    [PDF] Chapter 7: Basics of X-ray Diffraction
    In X-ray diffraction work we normally distinguish between single crystal and polycrystalline or powder applications. The single crystal sample is a perfect ...
  44. [44]
    Light through the ages: Ancient Greece to Maxwell - MacTutor
    The wave theory by Huygens and Hooke was a development of Descartes' ideas where now they proposed that light be a wave through the plenum, while Newton ...
  45. [45]
    A Very Brief History of Light | SpringerLink
    Dec 14, 2016 · In 1821, Fresnel published a paper in which he claimed that light is a transverse wave. Young had independently reached the same conclusion. ...<|control11|><|separator|>
  46. [46]
    July 1816: Fresnel's Evidence for the Wave Theory of Light
    Jul 1, 2016 · The wave theory was making a comeback, thanks in part to the work of a French civil engineer named Augustin-Jean Fresnel.
  47. [47]
    '…a paper …I hold to be great guns': a commentary on Maxwell ...
    Apr 13, 2015 · Maxwell's great paper of 1865 established his dynamical theory of the electromagnetic field. The origins of the paper lay in his earlier papers of 1856.
  48. [48]
    The dual nature of light as reflected in the Nobel archives
    Dec 2, 1999 · Albert Einstein in 1905 drew the conclusion that light sometimes behaves as particles ... The photon is the field quantum (particle) acting ...
  49. [49]
    [PDF] History of Seismology - Institute of Geophysics and Planetary Physics
    In 1900 R.D. Oldham, using measurements of the. 1897 Assam earthquake, classified the preliminary tremors into two phases, which he identified with longitudinal ...