Fact-checked by Grok 2 weeks ago

Simple harmonic motion

Simple harmonic motion (SHM) is a type of periodic oscillatory motion in which a body moves back and forth about an such that its is always directed toward the equilibrium and is directly proportional to the from it. This motion arises when the restoring force follows , F = -kx, where k is the force and x is the . Mathematically, the as a function of time is described by x(t) = A \cos(\omega t + \phi), where A is the , \omega = 2\pi / T is the , T is the , and \phi is the . The period of SHM for a -spring system is given by T = 2\pi \sqrt{m/k}, independent of , which is a key characteristic distinguishing it from other oscillations. Common examples include a attached to a oscillating horizontally or vertically, and a simple for small angular displacements less than about 15 degrees, where the motion approximates SHM. SHM serves as the foundational model for understanding more complex phenomena, such as vibrations in mechanical systems, sound waves, and oscillations in molecules. The study of SHM originated with Galileo's observations of pendulum motion c. 1583, followed by developments from and in the late , who analyzed the isochronous properties and mathematical solutions of harmonic oscillators. These insights laid the groundwork for and continue to underpin fields like , acoustics, and quantum physics, where the models energy levels in atoms.

Definition and Characteristics

Definition

Simple harmonic motion (SHM) is a type of periodic motion in which an object oscillates about an equilibrium such that the restoring force acting on it is directly proportional to its from that and directed opposite to the . This proportionality results in the object's varying sinusoidally with time, producing a smooth, repetitive back-and-forth trajectory. SHM represents an idealized form of commonly observed in systems where the net force follows , F = -kx, with k as the constant of proportionality and x as the . The ideal assumptions underlying SHM include a frictionless environment to eliminate energy dissipation through damping, ensuring perpetual oscillation without amplitude decay. Additionally, the restoring force must be linear, adhering strictly to Hooke's law for springs or equivalent relations in other systems, and small-angle approximations are applied where necessary to maintain the proportionality, such as in pendulum motions./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/15%3A_Oscillations/15.02%3A_Simple_Harmonic_Motion) The foundational concepts of SHM emerged in the 17th century through the studies of Christiaan Huygens on pendulum oscillations and Robert Hooke on spring elasticity, laying the groundwork for understanding periodic restoring forces. In this motion, the period T is the time for one complete oscillation, during which the object returns to its initial position with the same direction of velocity, while the frequency f is the number of oscillations per unit time, given by f = 1/T./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/15%3A_Oscillations/15.02%3A_Simple_Harmonic_Motion)

Key Characteristics

Simple harmonic motion (SHM) exhibits about the position, where the trajectory in forms an symmetric with respect to both the and axes, ensuring that the motion mirrors itself across the point. This implies that the time spent traversing from one extreme to the opposite extreme equals the time to return to the starting extreme, with each segment taking half the period T/2, reflecting the balanced nature of the oscillatory path. A defining feature of ideal SHM is isochronism, where the of remains constant regardless of the , provided displacements are small enough to satisfy the . This property, first noted by Galileo for , arises from the quadratic potential that governs the motion and holds approximately for small angles or extensions. The acceleration in SHM is directly proportional to the from but directed oppositely, expressed as a = -\omega^2 x, where \omega is the angular frequency and x is the displacement. In SHM, velocity reaches its maximum value at the equilibrium position, where displacement is zero, and drops to zero at the extreme positions of maximum displacement. Conversely, displacement is maximum at the extremes and zero at equilibrium. The state of motion evolves continuously through a phase angle, which parameterizes the sinusoidal functions describing position, velocity, and acceleration, capturing the progression through the cycle. This phase provides a unified way to describe the timing and orientation of the oscillation relative to a reference.

Mathematical Formulation

Differential Equation

The differential equation governing simple harmonic motion (SHM) arises from applying Newton's second law to a system where the restoring force is directly proportional to the displacement from equilibrium, such as a mass-spring system. The restoring force is given by Hooke's law, F = -kx, where k > 0 is the force constant and x is the displacement. Newton's second law states F = ma, with acceleration a = \frac{d^2x}{dt^2}, so for a mass m, m \frac{d^2x}{dt^2} = -kx. Dividing through by m yields the standard form of the differential equation: \frac{d^2x}{dt^2} + \frac{k}{m} x = 0. This equation is often expressed in a more general form as \frac{d^2x}{dt^2} + \omega^2 x = 0, where \omega = \sqrt{k/m} is the angular frequency of the motion, with units of radians per second. As a second-order linear homogeneous ordinary differential equation with constant coefficients, its solutions are inherently oscillatory and take the form of sinusoidal functions, reflecting the periodic nature of SHM. To confirm, substitute the trial solution x(t) = A \cos(\omega t + \phi), where A is the amplitude and \phi is the phase constant. The second derivative is \frac{d^2x}{dt^2} = -\omega^2 A \cos(\omega t + \phi) = -\omega^2 x(t). Plugging into the differential equation gives -\omega^2 x + \omega^2 x = 0, which is satisfied for \omega = \sqrt{k/m}.

General Solution

The general solution to the differential equation governing simple harmonic motion, \frac{d^2x}{dt^2} + \omega^2 x = 0, is given by x(t) = A \cos(\omega t + \phi), where A is the , representing the maximum from the position; \omega is the , which determines the rate of ; and \phi is the phase constant, which specifies the initial phase of the motion. This form arises from the r^2 + \omega^2 = 0, yielding r = \pm i \omega, and the real part of the exponential provides the cosine dependence. Equivalent alternative forms include x(t) = A \sin(\omega t + \psi), where \psi = \phi + \pi/2, or the linear combination x(t) = C \cos(\omega t) + D \sin(\omega t), with A = \sqrt{C^2 + D^2} and \tan \phi = -D/C (adjusting for the appropriate quadrant). These representations are interchangeable and chosen based on initial conditions for convenience. The constants A and \phi are determined from initial conditions, such as the position x(0) and velocity v(0) = \frac{dx}{dt}(0). Specifically, x(0) = A \cos \phi and v(0) = -A \omega \sin \phi, allowing solution for A = \sqrt{x(0)^2 + \left( \frac{v(0)}{\omega} \right)^2} and \phi = \tan^{-1} \left( -\frac{v(0)}{\omega x(0)} \right). The period of the motion is T = \frac{2\pi}{\omega}, the time for one complete cycle, and the frequency is f = \frac{\omega}{2\pi} = \frac{1}{T}, measured in hertz. In more advanced contexts, such as linear systems or wave mechanics, simple harmonic motion is often represented using exponentials for analytical convenience, where x(t) = \mathrm{Re} \left[ \tilde{A} e^{i \omega t} \right] and \tilde{A} is a amplitude incorporating the . This form facilitates superposition and without altering the physical real-valued displacement.

Dynamics and Energy

Restoring Force and Acceleration

In simple harmonic motion (SHM), the restoring force acting on the oscillating object is directly proportional to its from the position and directed opposite to that , given by as F = -kx, where k is the force constant and x is the . This negative ensures the force always pulls or pushes the object back toward , preventing unbounded motion and enabling periodic . The acceleration of the object follows from Newton's second law, a = F/m = -(k/m)x, where m is the , which simplifies to a = -\omega^2 x with \omega = \sqrt{k/m} as the . Thus, is also proportional to displacement but opposite in direction, always directed toward the point; for positive x, both and are negative, causing the to decrease and eventually reverse at the turning points where displacement is maximum. This opposition between displacement and —where peaks when displacement is zero and vice versa—characterizes the sinusoidal nature of SHM. Graphically, plotting against yields a straight line passing through the origin with -\omega^2, confirming the linear relationship inherent to SHM. In real physical systems, this behavior requires the restoring force to be linearly dependent on , a condition approximated in systems like springs or pendulums for small amplitudes but deviating at larger displacements due to nonlinear effects.

Kinetic and Potential Energy

In simple harmonic motion (SHM), the total E remains constant throughout the , assuming no dissipative forces are present, and is expressed as E = \frac{1}{2} k A^2, where k is the effective spring constant and A is the of the motion. This constancy arises from the conservative nature of the restoring , ensuring that the system's is conserved as it oscillates. The total energy E is the sum of the kinetic energy K and the potential energy U, such that K + U = E at every point in the motion. The potential energy U, which represents the energy stored due to displacement from the equilibrium position, is given by U = \frac{1}{2} k x^2, where x is the instantaneous displacement. This potential energy reaches its minimum value of zero at the equilibrium position (x = 0) and its maximum value of E at the extremes of the motion (x = \pm A). The K, associated with the of the oscillating , is expressed as K = \frac{1}{2} m v^2, where m is the and v is the instantaneous . In the context of SHM, this can be rewritten using the relation for derived from the general , yielding K = \frac{1}{2} m \omega^2 (A^2 - x^2), where \omega = \sqrt{k/m} is the . Consequently, K achieves its maximum value of E at the equilibrium position (x = 0) and drops to zero at the displacement extremes (x = \pm A). As the system oscillates, continuously interconverts between its kinetic and potential forms while maintaining the total E invariant. At the points of maximum , all is potential (K = 0, U = E); conversely, at , all is kinetic (U = 0, K = E). This interconversion is driven by the restoring force, which performs negative work on the , thereby transferring from kinetic to potential (and vice versa) without net loss.

Examples

Mass-Spring System

The -spring system serves as the canonical example of simple harmonic motion (SHM), where a m is attached to one end of a with spring constant k, and the other end is fixed. In the horizontal configuration, the system oscillates along a frictionless surface, with measured from the spring's unstretched position. The restoring force follows , F = -kx, leading to the m \frac{d^2x}{dt^2} = -kx, which yields an \omega = \sqrt{\frac{k}{m}} and T = 2\pi \sqrt{\frac{m}{k}}. For the vertical configuration, shifts the position downward by \Delta x = \frac{mg}{k}, where the spring is stretched to balance the weight, but oscillations about this new still exhibit SHM with the same \omega and T, independent of . is measured from this stretched , and the A represents the maximum deviation from it, ensuring the motion remains linear for small A. The effective restoring force remains F = -k(x - \Delta x), confirming the system's SHM behavior. This system is readily verified experimentally by first confirming through static measurements of force versus displacement to determine k, then observing the period's dependence on m and independence from A using timed oscillations. Such labs demonstrate the predicted T with high accuracy for moderate masses and small amplitudes. The analysis assumes an ideal spring: massless, perfectly elastic, and obeying linearly without damping or friction. In reality, the spring has negligible but non-zero , and for large amplitudes, deviations from linearity occur due to material nonlinearity, altering \omega and causing anharmonic motion.

Simple Pendulum

A simple pendulum consists of a m attached to a massless or of length L, suspended from a fixed and free to swing in a vertical plane under the influence of . When displaced from its equilibrium position and released, the pendulum bob oscillates about the vertical. The restoring torque arises from the gravitational force component tangential to the arc of motion, given by \tau = -mg \sin \theta \, L, where \theta is the from the vertical. For small angular displacements, the \sin \theta \approx \theta (with \theta in radians) simplifies this to \tau \approx -mg \theta \, L. Applying Newton's second law for rotation, \tau = I \alpha with I = m L^2 and \alpha = d^2 \theta / dt^2, yields the \frac{d^2 \theta}{dt^2} + \frac{g}{L} \theta = 0. This is the standard form for simple harmonic motion, with \omega = \sqrt{g/L} and T = 2\pi \sqrt{L/g}. The period depends only on the length L and g, independent of m or amplitude under this approximation, illustrating the isochronous nature of small oscillations. The holds well for initial displacements less than 15°, where \sin \theta and \theta differ by less than 1%, ensuring the period remains nearly constant. Beyond this range, the actual period increases slightly due to the nonlinearity of \sin \theta. Physically, the can be viewed as analogous to a mass- system, where provides an effective spring constant k = mg / L, with the arc displacement s = L \theta serving as the linear extension. This gravitational restoring force mimics the linear proportionality of for small angles, enabling the harmonic behavior.

Projection of Uniform Circular Motion

Simple harmonic motion can be visualized as the projection of uniform onto a straight line. Imagine a particle moving at constant angular speed \omega around a circle of radius A, which represents the of the oscillation. If the particle's position in the plane is described by Cartesian coordinates, the x-component of its displacement from the center is x(t) = A \cos(\omega t + \phi), where \phi is the initial phase angle. This expression directly matches the general solution for the displacement in simple harmonic motion. The and of this projected motion further confirm the harmonic nature. Differentiating the gives the x-component of : v_x(t) = -A \omega \sin(\omega t + \phi). The is then a_x(t) = -\omega^2 A \cos(\omega t + \phi) = -\omega^2 x(t), which satisfies the for simple harmonic motion, \frac{d^2 x}{dt^2} + \omega^2 x = 0. This kinematic derivation highlights how the centripetal in projects to a restoring proportional to . The circular path, often called the reference circle, provides a geometric tool for representing the of the , with the angular position \theta = \omega t + \phi corresponding to the . The y-component of the motion, y(t) = A \sin(\omega t + \phi), executes simple harmonic motion that is shifted by 90 degrees relative to the x-component, a relationship known as . This difference arises naturally from the orthogonal projections and underscores the vectorial composition in two dimensions. This projection model finds applications in understanding more complex oscillatory systems. For instance, it facilitates the analysis of coupled oscillators by visualizing normal modes as synchronized circular projections, and it explains the formation of Lissajous figures, which emerge when two perpendicular simple harmonic motions with commensurate frequencies are combined. Notably, this analogy is purely mathematical and kinematic, without implying any physical force driving the circular motion; it serves solely to illustrate the sinusoidal time dependence and phase characteristics of simple harmonic motion.

Mechanical Linkages

Mechanical linkages are engineered mechanisms that generate or approximate simple harmonic motion (SHM) through the conversion of rotary to linear motion or vice versa, commonly employed in machinery to achieve precise oscillatory paths. The mechanism exemplifies an ideal realization of SHM in a , where a rotating pin slides within a slotted , driving the yoke in pure linear . In this setup, the of the yoke from its position is described by the equation x = r \cos(\theta), with \theta = \omega t, where r is the crank radius, \omega is the , and t is time; this directly mirrors the projection of uniform onto a . Unlike more complex linkages, the eliminates side loads on the sliding element in ideal frictionless conditions, ensuring the output follows exact sinusoidal motion. The piston-crank mechanism, also known as the , approximates SHM but deviates for larger angles due to the connecting rod's influence. For small crank angles relative to the connecting rod length, the piston's motion closely follows SHM, with displacement roughly x \approx r (1 - \cos(\theta)) near the extremes, making it suitable for applications like internal combustion engines where exact harmonicity is not critical. This approximation holds because the geometry reduces to a near-projection of when the rod is much longer than the , minimizing nonlinear effects. Other linkages, such as the double slider crank or , generate SHM along specific paths by constraining two sliders in perpendicular grooves connected by a link with a marked point. In the , the point traces an , but its projections onto the groove axes exhibit pure SHM with equal to the groove spacing, providing a practical for demonstrating orthogonal components. These offer key advantages in converting rotary to exactly in ideal cases like the , avoiding the approximations and secondary forces inherent in slider-crank designs, which enhances efficiency in low-friction environments. Historically, such linkages have been integral to engines and devices since the , with the notably applied in early pumps to produce reliable reciprocating action for drives.

Extensions and Applications

Phasor Representation

In simple harmonic motion (SHM), the representation models the oscillatory as the projection of a rotating , offering a compact vector-based approach to analyze , , and . A is a of fixed equal to the A, rotating counterclockwise in the at constant angular speed \omega. The x-component (real part) of this vector at any time t yields the x(t) = A \cos(\omega t + \phi), where \phi is the phase constant determining the initial . This geometric interpretation links the linear SHM to uniform , with the phasor's tip tracing a circle of radius A. To represent associated quantities, additional phasors are introduced with specific phase shifts relative to the displacement phasor. The velocity phasor, which has magnitude A\omega, leads the phasor by $90^\circ (or \pi/2 radians), corresponding to v(t) = -A\omega \sin(\omega t + \phi). The acceleration phasor, with magnitude A\omega^2, leads by $180^\circ (or \pi radians), aligning opposite to the and matching a(t) = -A\omega^2 \cos(\omega t + \phi). Phasor diagrams visualize these relationships by drawing the vectors tail-to-origin: the along the horizontal axis, velocity vertical and ahead, and acceleration horizontal but reversed, all rotating synchronously at \omega. The complex exponential notation formalizes this representation, expressing displacement as x(t) = \operatorname{Re} \left[ A e^{i(\omega t + \phi)} \right], where the complex amplitude A e^{i\phi} encodes and initial phase. Velocity follows as i\omega times the displacement , yielding v(t) = \operatorname{Re} \left[ i\omega A e^{i(\omega t + \phi)} \right], while acceleration is -\omega^2 times it, simplifying time derivatives to algebraic multiplications by i\omega or -\omega^2. This method excels in handling superposition, where multiple SHMs of identical add vectorially via summation to find the resultant and directly. It also underpins by decomposing complex periodic signals into harmonic , enabling efficient computation of oscillatory systems.

Relation to

Simple harmonic motion (SHM) forms the foundational mechanism underlying harmonic , which can be either transverse or longitudinal. In a harmonic wave, each particle of the medium executes SHM about its position, with the disturbance propagating through the medium while the particles themselves do not travel net distances. This oscillatory behavior of individual particles gives rise to the wave's periodic nature, where the displacement varies sinusoidally with time at any fixed point./12%3A_Waves_in_One_Dimension/12.01%3A_Traveling_Waves) The mathematical description of a traveling harmonic wave is captured by the one-dimensional wave equation, \frac{\partial^2 y}{\partial t^2} = v^2 \frac{\partial^2 y}{\partial x^2}, where y(x, t) is the , v is the constant wave speed, x is , and t is time. A general solution for a wave propagating in the positive x-direction is y(x, t) = A \cos(kx - \omega t + \phi), with A as the , k as the wave number, \omega as the , and \phi as the phase constant. At a fixed position x, the equation simplifies to \frac{\partial^2 y}{\partial t^2} + \omega^2 y = 0, which is the standard for SHM, confirming the oscillation of each particle with \omega / 2\pi. The wavelength \lambda relates to the wave number by \lambda = 2\pi / k, and the wave speed is given by v = \omega / k = f \lambda, where f is the ./12%3A_Waves_in_One_Dimension/12.01%3A_Traveling_Waves) Standing waves emerge from the superposition of two identical harmonic waves traveling in opposite directions, such as along a fixed at both ends. The resulting pattern features stationary nodes, where displacement is always zero, and antinodes, where displacement reaches maximum , with the wave profile oscillating in place rather than propagating. addition can represent this constructively at antinodes and destructively at nodes. These principles apply broadly in natural phenomena, including waves, where air molecules undergo longitudinal SHM to propagate variations, and , modeled as electromagnetic waves in which the electric and oscillate harmonically, effectively behaving as coupled SHM systems across vast collections of oscillators.

References

  1. [1]
    [PDF] SIMPLE HARMONIC MOTION
    When the restoring force is directly proportional to displacement from equilibrium, the oscillation is called simple harmonic motion (SHM). • Important ...
  2. [2]
    15.1 Simple Harmonic Motion – University Physics Volume 1
    A very common type of periodic motion is called simple harmonic motion (SHM). A system that oscillates with SHM is called a simple harmonic oscillator.Missing: key | Show results with:key
  3. [3]
    Simple Harmonic Motion
    An interesting and important characteristic of all simple harmonic motion is that the period remains the same even for motions with quite different amplitudes.Missing: key | Show results with:key
  4. [4]
    5.5 Simple Harmonic Motion - Physics | OpenStax
    Mar 26, 2020 · For small displacements of less than 15 degrees, a pendulum experiences simple harmonic oscillation, meaning that its restoring force is ...<|control11|><|separator|>
  5. [5]
    21 The Harmonic Oscillator - Feynman Lectures - Caltech
    The simplest mechanical system whose motion follows a linear differential equation with constant coefficients is a mass on a spring.
  6. [6]
    [PPT] Simple harmonic motion (PPT, 460KB)
    The study of SHM started with Galileo's pendulum experiments (1638, Two new sciences). Huygens, Newton & others developed further analyses.
  7. [7]
    Simple Harmonic Motion and Resonance | Middle Tennessee State ...
    SHM occurs around an equilibrium position when a mass is subject to a linear restoring force. A linear restoring force is one that gets progressively larger ...Missing: characteristics | Show results with:characteristics
  8. [8]
    Simple harmonic motion - Physics
    Nov 17, 1999 · An object experiencing simple harmonic motion is traveling in one dimension, and its one-dimensional motion is given by an equation of the form
  9. [9]
    118. 16.3 Simple Harmonic Motion: A Special Periodic Motion
    Simple Harmonic Motion (SHM) is the name given to oscillatory motion for a system where the net force can be described by Hooke's law.
  10. [10]
    June 16, 1657: Christiaan Huygens Patents the First Pendulum Clock
    Jun 16, 2017 · He became the first to report the phenomenon of coupled oscillation in two pendulum clocks (which he invented) in his bedroom while recovering ...Missing: harmonic | Show results with:harmonic
  11. [11]
    [PDF] Chapter 23 Simple Harmonic Motion - MIT OpenCourseWare
    Jul 23, 2022 · The accuracy of clocks was increased and the size reduced by the discovery of the oscillatory properties of springs by Robert Hooke.
  12. [12]
    [PDF] SYMMETRY
    Such a figure is symmetric about both the symmetry line x=0 and also the symmetry line dx/dt=0. It represents simple harmonic motion as viewed in the phase ...
  13. [13]
    MATHEMATICA TUTORIAL, Part 1.4: Pendulum
    Its approximate isochronism, was first discovered by Galileo, who is ... Beyond this limit, the equation of motion is nonlinear: the simple harmonic motion ...
  14. [14]
    15.1 Simple Harmonic Motion - University Physics Volume 1
    Sep 19, 2016 · In simple harmonic motion, the acceleration of the system, and therefore the net force, is proportional to the displacement and acts in the ...Missing: symmetry, isochronism,
  15. [15]
    Simple Harmonic Motion Frequency - HyperPhysics
    Using Hooke's law and neglecting damping and the mass of the spring, Newton's second law gives the equation of motion: The solution to this differential ...
  16. [16]
    Differential Equations - Mechanical Vibrations
    Nov 16, 2022 · Mechanical vibrations involve a mass attached to a spring, using second-order differential equations to model the displacement of the mass.<|control11|><|separator|>
  17. [17]
    15.1 Simple Harmonic Motion – General Physics Using Calculus I
    The angular frequency ω , period T, and frequency f of a simple harmonic oscillator are given by ω = k m , T = 2 π m k , and f = 1 2 π k m , where m is the mass ...
  18. [18]
    [PDF] Lecture 1: Simple Harmonic Oscillators
    As we will see in this course, oscillators are the building blocks of a tremendous diversity of physical phenomena and technologies, including musical ...
  19. [19]
    Simple Harmonic Motion - HyperPhysics
    Simple harmonic motion is typified by the motion of a mass on a spring when it is subject to the linear elastic restoring force given by Hooke's Law.Missing: definition | Show results with:definition
  20. [20]
    [PDF] SIMPLE HARMONIC MOTION - Rutgers Physics
    When the restoring force is opposite and directly proportional to the displacement, the oscillating object will exhibit Simple Harmonic Motion, SHM. The object ...
  21. [21]
    Harmonic motion
    Simple harmonic motion is accelerated motion. If an object exhibits simple harmonic motion, a force must be acting on the object. The force is. F = ma = -mω2 ...
  22. [22]
    [PDF] Chapter 15
    An object moves with simple harmonic motion whenever its acceleration is proportional to its position and is oppositely directed to the displacement from.
  23. [23]
    15.6 Energy Conservation in SHM
    Kinetic energy is an energy of movement. Kinetic energy can be transformed into potential energy as an object changes position. Potential energy is an energy ...
  24. [24]
    [PDF] Chapter 23 Simple Harmonic Motion
    Jul 23, 2013 · There are two basic ways to measure time: by duration or periodic motion. Early clocks measured duration by calibrating the burning of incense ...
  25. [25]
    [PDF] Experiment 5 - College of San Mateo
    The purposes of this experiment are to measure the force constant of a spring, to verify the theoretical equation for the period of a mass oscillating in simple ...
  26. [26]
    [PDF] Experiment IX: Simple Harmonic Motion - FSU Hadronic Physics
    In this lab we will verify that the force from a spring obeys Hooke's law and the period of the motion of a mass on a spring is indeed given by Equation 9-3.
  27. [27]
    [PDF] Springs: Part I
    Because γ = 0 when there is no resistive force to dampen the motion, we say the mass/spring system is undamped when γ = 0. We will see that the motion of the ...Missing: limitations | Show results with:limitations
  28. [28]
    Oscillation of a Simple Pendulum - Graduate Program in Acoustics
    By applying Newton's secont law for rotational systems, the equation of motion for the pendulum may be obtained equation of simple harmonic motion τ = I α ⇒ − ...
  29. [29]
    15.4 Pendulums – General Physics Using Calculus I
    Using the small angle approximation gives an approximate solution for small angles, d 2 θ d t 2 = − g L θ . T = 2 π L g . The period of a simple pendulum ...
  30. [30]
    The Simple Pendulum – Introductory Physics for the Health and Life ...
    with effective spring constant k = m g L and displacement . Therefore, for small displacements, the simple pendulum behaves as a simple harmonic oscillator. ...
  31. [31]
    16.6 Uniform Circular Motion and Simple Harmonic Motion
    There is an easy way to produce simple harmonic motion by using uniform circular motion. Figure 2 shows one way of using this method. A ball is attached to a ...
  32. [32]
    121 Uniform Circular Motion and Simple Harmonic Motion
    The projection of an object in uniform circular motion follows simple harmonic motion. The mathematical descriptions for SHM can be derived from the geometry of ...
  33. [33]
  34. [34]
    [PDF] Physics, Chapter 12: Periodic Motion - UNL Digital Commons
    When an object moves in uniform circular motion, its projection onto the x or y axis moves in simple harmonic motion. The projected position corresponds to the ...
  35. [35]
    "Analysis and Design of Basic Linkages for Harmonic Motion ...
    ... simple harmonic motion at the output members. A ... slider-crank linkages. The method involves the ... simple harmonic motion at the output members.
  36. [36]
    [PDF] Chapter 15 - UNIVERSITY PHYSICS
    An object attached to a spring sliding on a frictionless surface is an uncomplicated simple harmonic oscillator. ... This design is known as a scotch yoke.
  37. [37]
    (PDF) An Automated Scotch Yoke Mechanism - Academia.edu
    The motion of the Scotch Yoke Mechanism is such that pure simple harmonic motion is produced by the mechanism when driven by an eccentric or crank. Because ...
  38. [38]
    Scotch yoke – Knowledge and References - Taylor & Francis
    As a result the piston acquires a simple, as opposed to non-simple, harmonic motion. That is, it now oscillates about its equilibrium or mid-stroke position ...
  39. [39]
    [PDF] Kinematics of Machines - Cornell eCommons
    Composition of Simple Harmonic Motion not along Same Line.. . . . . ... . 67. V. Page 10. VI. CONTENTS ... SLIDER-CRANK CHAINS. 34. Slider-crank Chains ...
  40. [40]
    (PDF) Mechanics of Machinery I - Academia.edu
    B.M. El-Souhily 7- Elliptic Trammel : An elliptic trammel is a four-bar linkage with RPPR joints, an inversion of Scotch Yoke. It can be used to trace ...
  41. [41]
    Scotch Yoke Actuators: Types, Applications, and Maintenance Guide
    The origins of the scotch yoke mechanism date back to the 18th century. It was initially used in steam engines and other machinery. Over time, its applications ...
  42. [42]
    16.2 Mathematics of Waves - University Physics Volume 1 - OpenStax
    Sep 19, 2016 · The magnitude of the maximum acceleration is |aymax|=Aω2. The particles of the medium, or the mass elements, oscillate in simple harmonic ...
  43. [43]
    [PDF] • Doppler effect • Standing waves from superposition - Physics - UMD
    Standing Waves: graphical (I). • Two sine waves (same f, A, ) traveling in opposite directions: superposition is wave, but not moving λ. Page 10. • Nodes ...
  44. [44]
    17.1 Sound Waves - University Physics Volume 1 | OpenStax
    Sep 19, 2016 · The air molecules oscillate in simple harmonic motion about their equilibrium positions, as shown in part (b). Note that sound waves in air are ...<|control11|><|separator|>