Simple Harmonic Oscillator Pdf

Part of a series of articles about
Classical mechanics
F=ma{displaystyle {vec {F}}=m{vec {a}}}
  • Energy
  • Inertia / Moment of inertia

  • Mechanical power

  • Moment
  • Analytical mechanics
  • Damping (ratio)
  • Inertial / Non-inertial reference frame
  • Motion (linear)
  • Rigid body
  • Centrifugal force
  • Angular acceleration / displacement / frequency / velocity

Simple Harmonic Motion A system can oscillate in many ways, but we will be. Natural frequency of the oscillator. Solving the Harmonic Oscillator Equation Morgan Root NCSU Department of Math. Spring-Mass System. System”, or a Simple Harmonic Oscillator. We will solve this first. M&y&(t)+ky(t) =0. Simple Harmonic Oscillator y(t) ( Kt) y(t) ( Kt) y t Ky t K k m sin and cos this equation. Chapter 8 The Simple Harmonic Oscillator A winter rose. How can a rose bloom in December? Amazing but true, there it is, a yellow winter rose. The rain and the cold have worn at the petals but the beauty is eternal regardless.

In classical mechanics, a harmonic oscillator is a system that, when displaced from its equilibrium position, experiences a restoring forceFproportional to the displacement x:

F=kx,{displaystyle {vec {F}}=-k{vec {x}},}

where k is a positive constant.

If F is the only force acting on the system, the system is called a simple harmonic oscillator, and it undergoes simple harmonic motion: sinusoidaloscillations about the equilibrium point, with a constant amplitude and a constant frequency (which does not depend on the amplitude).

If a frictional force (damping) proportional to the velocity is also present, the harmonic oscillator is described as a damped oscillator. Depending on the friction coefficient, the system can:

  • Oscillate with a frequency lower than in the undamped case, and an amplitude decreasing with time (underdamped oscillator).
  • Decay to the equilibrium position, without oscillations (overdamped oscillator).

The boundary solution between an underdamped oscillator and an overdamped oscillator occurs at a particular value of the friction coefficient and is called critically damped.

If an external time-dependent force is present, the harmonic oscillator is described as a driven oscillator.

Mechanical examples include pendulums (with small angles of displacement), masses connected to springs, and acoustical systems. Other analogous systems include electrical harmonic oscillators such as RLC circuits. The harmonic oscillator model is very important in physics, because any mass subject to a force in stable equilibrium acts as a harmonic oscillator for small vibrations. Harmonic oscillators occur widely in nature and are exploited in many manmade devices, such as clocks and radio circuits. They are the source of virtually all sinusoidal vibrations and waves.

  • 3Driven harmonic oscillators
  • 5Universal oscillator equation
    • 5.2Steady-state solution
  • 8Examples
    • 8.2Spring/mass system

Simple harmonic oscillator[edit]

Mass-spring harmonic oscillator
Simple harmonic motion

A simple harmonic oscillator is an oscillator that is neither driven nor damped. It consists of a mass m, which experiences a single force F, which pulls the mass in the direction of the point x = 0 and depends only on the mass's position x and a constant k. Balance of forces (Newton's second law) for the system is

F=ma=md2xdt2=mx¨=kx.{displaystyle F=ma=m{frac {mathrm {d} ^{2}x}{mathrm {d} t^{2}}}=m{ddot {x}}=-kx.}

Solving this differential equation, we find that the motion is described by the function

x(t)=Acos(ωt+φ),{displaystyle x(t)=Acos(omega t+varphi ),}

where

ω=km.{displaystyle omega ={sqrt {frac {k}{m}}}.}

The motion is periodic, repeating itself in a sinusoidal fashion with constant amplitude A. In addition to its amplitude, the motion of a simple harmonic oscillator is characterized by its periodT=2π/ω{displaystyle T=2pi /omega }, the time for a single oscillation or its frequency f=1/T{displaystyle f=1/T}, the number of cycles per unit time. The position at a given time t also depends on the phaseφ, which determines the starting point on the sine wave. The period and frequency are determined by the size of the mass m and the force constant k, while the amplitude and phase are determined by the starting position and velocity.

Simple Harmonic Oscillation Pdf

The velocity and acceleration of a simple harmonic oscillator oscillate with the same frequency as the position, but with shifted phases. The velocity is maximal for zero displacement, while the acceleration is in the direction opposite to the displacement.

The potential energy stored in a simple harmonic oscillator at position x is

U=12kx2.{displaystyle U={frac {1}{2}}kx^{2}.}

Damped harmonic oscillator[edit]

Dependence of the system behavior on the value of the damping ratio ζ
Video clip demonstrating a damped harmonic oscillator consisting of a dynamics cart between two springs. An accelerometer on top of the cart shows the magnitude and direction of the acceleration.

In real oscillators, friction, or damping, slows the motion of the system. Due to frictional force, the velocity decreases in proportion to the acting frictional force. While in a simple undriven harmonic oscillator the only force acting on the mass is the restoring force, in a damped harmonic oscillator there is in addition a frictional force which is always in a direction to oppose the motion. In many vibrating systems the frictional force Ff can be modeled as being proportional to the velocity v of the object: Ff = −cv, where c is called the viscous damping coefficient.

The balance of forces (Newton's second law) for damped harmonic oscillators is then

F=kxcdxdt=md2xdt2,{displaystyle F=-kx-c{frac {mathrm {d} x}{mathrm {d} t}}=m{frac {mathrm {d} ^{2}x}{mathrm {d} t^{2}}},}[1][2][3]

which can be rewritten into the form

d2xdt2+2ζω0dxdt+ω02x=0,{displaystyle {frac {mathrm {d} ^{2}x}{mathrm {d} t^{2}}}+2zeta omega _{0}{frac {mathrm {d} x}{mathrm {d} t}}+omega _{0}^{2}x=0,}

where

ω0=km{displaystyle omega _{0}={sqrt {frac {k}{m}}}} is called the 'undamped angular frequency of the oscillator',
ζ=c2mk{displaystyle zeta ={frac {c}{2{sqrt {mk}}}}} is called the 'damping ratio'.
Step response of a damped harmonic oscillator; curves are plotted for three values of μ = ω1 = ω01 − ζ2. Time is in units of the decay time τ = 1/(ζω0).

The value of the damping ratio ζ critically determines the behavior of the system. A damped harmonic oscillator can be:

  • Overdamped (ζ > 1): The system returns (exponentially decays) to steady state without oscillating. Larger values of the damping ratio ζ return to equilibrium more slowly.
  • Critically damped (ζ = 1): The system returns to steady state as quickly as possible without oscillating (although overshoot can occur). This is often desired for the damping of systems such as doors.
  • Underdamped (ζ < 1): The system oscillates (with a slightly different frequency than the undamped case) with the amplitude gradually decreasing to zero. The angular frequency of the underdamped harmonic oscillator is given by ω1=ω01ζ2,{displaystyle omega _{1}=omega _{0}{sqrt {1-zeta ^{2}}},} the exponential decay of the underdamped harmonic oscillator is given by λ=ω0ζ.{displaystyle lambda =omega _{0}zeta .}

The Q factor of a damped oscillator is defined as

Q=2π×energy storedenergy lost per cycle.{displaystyle Q=2pi times {frac {text{energy stored}}{text{energy lost per cycle}}}.}

Q is related to the damping ratio by the equation Q=12ζ.{displaystyle Q={frac {1}{2zeta }}.}

Driven harmonic oscillators[edit]

Driven harmonic oscillators are damped oscillators further affected by an externally applied force F(t).

Newton's second law takes the form

F(t)kxcdxdt=md2xdt2.{displaystyle F(t)-kx-c{frac {mathrm {d} x}{mathrm {d} t}}=m{frac {mathrm {d} ^{2}x}{mathrm {d} t^{2}}}.}

It is usually rewritten into the form

d2xdt2+2ζω0dxdt+ω02x=F(t)m.{displaystyle {frac {mathrm {d} ^{2}x}{mathrm {d} t^{2}}}+2zeta omega _{0}{frac {mathrm {d} x}{mathrm {d} t}}+omega _{0}^{2}x={frac {F(t)}{m}}.}

This equation can be solved exactly for any driving force, using the solutions z(t) that satisfy the unforced equation

d2zdt2+2ζω0dzdt+ω02z=0,{displaystyle {frac {mathrm {d} ^{2}z}{mathrm {d} t^{2}}}+2zeta omega _{0}{frac {mathrm {d} z}{mathrm {d} t}}+omega _{0}^{2}z=0,}

and which can be expressed as damped sinusoidal oscillations:

z(t)=Aeζω0tsin(1ζ2ω0t+φ),{displaystyle z(t)=Amathrm {e} ^{-zeta omega _{0}t}sin left({sqrt {1-zeta ^{2}}}omega _{0}t+varphi right),}

in the case where ζ ≤ 1. The amplitude A and phase φ determine the behavior needed to match the initial conditions.

Step input[edit]

In the case ζ < 1 and a unit step input with x(0) = 0:

F(t)m={ω02t00t<0{displaystyle {frac {F(t)}{m}}={begin{cases}omega _{0}^{2}&tgeq 00&t<0end{cases}}}

the solution is

x(t)=1eζω0tsin(1ζ2ω0t+φ)sin(φ),{displaystyle x(t)=1-mathrm {e} ^{-zeta omega _{0}t}{frac {sin left({sqrt {1-zeta ^{2}}}omega _{0}t+varphi right)}{sin(varphi )}},}

with phase φ given by

cosφ=ζ.{displaystyle cos varphi =zeta .}

The time an oscillator needs to adapt to changed external conditions is of the order τ = 1/(ζω0). In physics, the adaptation is called relaxation, and τ is called the relaxation time.

In electrical engineering, a multiple of τ is called the settling time, i.e. the time necessary to ensure the signal is within a fixed departure from final value, typically within 10%. The term overshoot refers to the extent the response maximum exceeds final value, and undershoot refers to the extent the response falls below final value for times following the response maximum.

Sinusoidal driving force[edit]

Simple
Steady-state variation of amplitude with relative frequency ω/ω0{displaystyle omega /omega _{0}} and dampingζ{displaystyle zeta } of a driven simple harmonic oscillator

In the case of a sinusoidal driving force:

d2xdt2+2ζω0dxdt+ω02x=1mF0sin(ωt),{displaystyle {frac {mathrm {d} ^{2}x}{mathrm {d} t^{2}}}+2zeta omega _{0}{frac {mathrm {d} x}{mathrm {d} t}}+omega _{0}^{2}x={frac {1}{m}}F_{0}sin(omega t),}

where F0{displaystyle F_{0}} is the driving amplitude, and ω{displaystyle omega } is the driving frequency for a sinusoidal driving mechanism. This type of system appears in AC-driven RLC circuits (resistor–inductor–capacitor) and driven spring systems having internal mechanical resistance or external air resistance.

The general solution is a sum of a transient solution that depends on initial conditions, and a steady state that is independent of initial conditions and depends only on the driving amplitude F0{displaystyle F_{0}}, driving frequency ω{displaystyle omega }, undamped angular frequency ω0{displaystyle omega _{0}}, and the damping ratio ζ{displaystyle zeta }.

The steady-state solution is proportional to the driving force with an induced phase change φ{displaystyle varphi }:

x(t)=F0mZmωsin(ωt+φ),{displaystyle x(t)={frac {F_{0}}{mZ_{m}omega }}sin(omega t+varphi ),}

where

Zm=(2ω0ζ)2+1ω2(ω02ω2)2{displaystyle Z_{m}={sqrt {left(2omega _{0}zeta right)^{2}+{frac {1}{omega ^{2}}}(omega _{0}^{2}-omega ^{2})^{2}}}}

is the absolute value of the impedance or linear response function, and

φ=arctan(2ωω0ζω2ω02)+nπ{displaystyle varphi =arctan left({frac {2omega omega _{0}zeta }{omega ^{2}-omega _{0}^{2}}}right)+npi }

is the phase of the oscillation relative to the driving force. The phase value is usually taken to be between −180° and 0 (that is, it represents a phase lag, for both positive and negative values of the arctan argument).

For a particular driving frequency called the resonance, or resonant frequency ωr=ω012ζ2{displaystyle omega _{r}=omega _{0}{sqrt {1-2zeta ^{2}}}}, the amplitude (for a given F0{displaystyle F_{0}}) is maximal. This resonance effect only occurs when ζ<1/2{displaystyle zeta <1/{sqrt {2}}}, i.e. for significantly underdamped systems. For strongly underdamped systems the value of the amplitude can become quite large near the resonant frequency.

The transient solutions are the same as the unforced (F0=0{displaystyle F_{0}=0}) damped harmonic oscillator and represent the systems response to other events that occurred previously. The transient solutions typically die out rapidly enough that they can be ignored.

Parametric oscillators[edit]

A parametric oscillator is a driven harmonic oscillator in which the drive energy is provided by varying the parameters of the oscillator, such as the damping or restoring force.A familiar example of parametric oscillation is 'pumping' on a playground swing.[4][5][6]A person on a moving swing can increase the amplitude of the swing's oscillations without any external drive force (pushes) being applied, by changing the moment of inertia of the swing by rocking back and forth ('pumping') or alternately standing and squatting, in rhythm with the swing's oscillations. The varying of the parameters drives the system. Examples of parameters that may be varied are its resonance frequency ω{displaystyle omega } and damping β{displaystyle beta }.

Parametric oscillators are used in many applications. The classical varactor parametric oscillator oscillates when the diode's capacitance is varied periodically. The circuit that varies the diode's capacitance is called the 'pump' or 'driver'. In microwave electronics, waveguide/YAG based parametric oscillators operate in the same fashion. The designer varies a parameter periodically to induce oscillations.

Parametric oscillators have been developed as low-noise amplifiers, especially in the radio and microwave frequency range. Thermal noise is minimal, since a reactance (not a resistance) is varied. Another common use is frequency conversion, e.g., conversion from audio to radio frequencies. For example, the Optical parametric oscillator converts an input laser wave into two output waves of lower frequency (ωs,ωi{displaystyle omega _{s},omega _{i}}).

Parametric resonance occurs in a mechanical system when a system is parametrically excited and oscillates at one of its resonant frequencies. Parametric excitation differs from forcing, since the action appears as a time varying modification on a system parameter. This effect is different from regular resonance because it exhibits the instability phenomenon.

Universal oscillator equation[edit]

The equation

d2qdτ2+2ζdqdτ+q=0{displaystyle {frac {mathrm {d} ^{2}q}{mathrm {d} tau ^{2}}}+2zeta {frac {mathrm {d} q}{mathrm {d} tau }}+q=0}

is known as the universal oscillator equation, since all second-order linear oscillatory systems can be reduced to this form.[citation needed] This is done through nondimensionalization.

If the forcing function is f(t) = cos(ωt) = cos(ωtcτ) = cos(ωτ), where ω = ωtc, the equation becomes

d2qdτ2+2ζdqdτ+q=cos(ωτ).{displaystyle {frac {mathrm {d} ^{2}q}{mathrm {d} tau ^{2}}}+2zeta {frac {mathrm {d} q}{mathrm {d} tau }}+q=cos(omega tau ).}

The solution to this differential equation contains two parts: the 'transient' and the 'steady-state'.

Transient solution[edit]

The solution based on solving the ordinary differential equation is for arbitrary constants c1 and c2

qt(τ)={eζτ(c1eτζ21+c2eτζ21)ζ>1 (overdamping)eζτ(c1+c2τ)=eτ(c1+c2τ)ζ=1 (critical damping)eζτ[c1cos(1ζ2τ)+c2sin(1ζ2τ)]ζ<1 (underdamping){displaystyle q_{t}(tau )={begin{cases}mathrm {e} ^{-zeta tau }left(c_{1}mathrm {e} ^{tau {sqrt {zeta ^{2}-1}}}+c_{2}mathrm {e} ^{-tau {sqrt {zeta ^{2}-1}}}right)&zeta >1{text{ (overdamping)}}mathrm {e} ^{-zeta tau }(c_{1}+c_{2}tau )=mathrm {e} ^{-tau }(c_{1}+c_{2}tau )&zeta =1{text{ (critical damping)}}mathrm {e} ^{-zeta tau }left[c_{1}cos left({sqrt {1-zeta ^{2}}}tau right)+c_{2}sin left({sqrt {1-zeta ^{2}}}tau right)right]&zeta <1{text{ (underdamping)}}end{cases}}}

The transient solution is independent of the forcing function.

Steady-state solution[edit]

Apply the 'complex variables method' by solving the auxiliary equation below and then finding the real part of its solution:

d2qdτ2+2ζdqdτ+q=cos(ωτ)+isin(ωτ)=eiωτ.{displaystyle {frac {mathrm {d} ^{2}q}{mathrm {d} tau ^{2}}}+2zeta {frac {mathrm {d} q}{mathrm {d} tau }}+q=cos(omega tau )+mathrm {i} sin(omega tau )=mathrm {e} ^{mathrm {i} omega tau }.}

Supposing the solution is of the form

qs(τ)=Aei(ωτ+φ).{displaystyle q_{s}(tau )=Amathrm {e} ^{mathrm {i} (omega tau +varphi )}.}

Its derivatives from zeroth to second order are

qs=Aei(ωτ+φ),dqsdτ=iωAei(ωτ+φ),d2qsdτ2=ω2Aei(ωτ+φ).{displaystyle q_{s}=Amathrm {e} ^{mathrm {i} (omega tau +varphi )},quad {frac {mathrm {d} q_{s}}{mathrm {d} tau }}=mathrm {i} omega Amathrm {e} ^{mathrm {i} (omega tau +varphi )},quad {frac {mathrm {d} ^{2}q_{s}}{mathrm {d} tau ^{2}}}=-omega ^{2}Amathrm {e} ^{mathrm {i} (omega tau +varphi )}.}

Substituting these quantities into the differential equation gives

Simple Harmonic Oscillator Quantum Mechanics Pdf

ω2Aei(ωτ+φ)+2ζiωAei(ωτ+φ)+Aei(ωτ+φ)=(ω2A+2ζiωA+A)ei(ωτ+φ)=eiωτ.{displaystyle -omega ^{2}Amathrm {e} ^{mathrm {i} (omega tau +varphi )}+2zeta mathrm {i} omega Amathrm {e} ^{mathrm {i} (omega tau +varphi )}+Amathrm {e} ^{mathrm {i} (omega tau +varphi )}=(-omega ^{2}A+2zeta mathrm {i} omega A+A)mathrm {e} ^{mathrm {i} (omega tau +varphi )}=mathrm {e} ^{mathrm {i} omega tau }.}

Dividing by the exponential term on the left results in

ω2A+2ζiωA+A=eiφ=cosφisinφ.{displaystyle -omega ^{2}A+2zeta mathrm {i} omega A+A=mathrm {e} ^{-mathrm {i} varphi }=cos varphi -mathrm {i} sin varphi .}

Equating the real and imaginary parts results in two independent equations

A(1ω2)=cosφ,2ζωA=sinφ.{displaystyle A(1-omega ^{2})=cos varphi ,quad 2zeta omega A=-sin varphi .}

Amplitude part[edit]

Bode plot of the frequency response of an ideal harmonic oscillator

Squaring both equations and adding them together gives

A2(1ω2)2=cos2φ(2ζωA)2=sin2φ}A2[(1ω2)2+(2ζω)2]=1.{displaystyle left.{begin{aligned}A^{2}(1-omega ^{2})^{2}&=cos ^{2}varphi (2zeta omega A)^{2}&=sin ^{2}varphi end{aligned}}right}Rightarrow A^{2}[(1-omega ^{2})^{2}+(2zeta omega )^{2}]=1.}

Therefore,

A=A(ζ,ω)=sign(sinφ2ζω)1(1ω2)2+(2ζω)2.{displaystyle A=A(zeta ,omega )=operatorname {sign} left({frac {-sin varphi }{2zeta omega }}right){frac {1}{sqrt {(1-omega ^{2})^{2}+(2zeta omega )^{2}}}}.}

Compare this result with the theory section on resonance, as well as the 'magnitude part' of the RLC circuit. This amplitude function is particularly important in the analysis and understanding of the frequency response of second-order systems.

Phase part[edit]

To solve for φ, divide both equations to get

tanφ=2ζω1ω2=2ζωω21φφ(ζ,ω)=arctan(2ζωω21)+nπ.{displaystyle tan varphi =-{frac {2zeta omega }{1-omega ^{2}}}={frac {2zeta omega }{omega ^{2}-1}}Rightarrow varphi equiv varphi (zeta ,omega )=arctan left({frac {2zeta omega }{omega ^{2}-1}}right)+npi .}

This phase function is particularly important in the analysis and understanding of the frequency response of second-order systems.

Full solution[edit]

Combining the amplitude and phase portions results in the steady-state solution

qs(τ)=A(ζ,ω)cos(ωτ+φ(ζ,ω))=Acos(ωτ+φ).{displaystyle q_{s}(tau )=A(zeta ,omega )cos(omega tau +varphi (zeta ,omega ))=Acos(omega tau +varphi ).}

The solution of original universal oscillator equation is a superposition (sum) of the transient and steady-state solutions:

q(τ)=qt(τ)+qs(τ).{displaystyle q(tau )=q_{t}(tau )+q_{s}(tau ).}

For a more complete description of how to solve the above equation, see linear ODEs with constant coefficients.

Equivalent systems[edit]

Harmonic oscillators occurring in a number of areas of engineering are equivalent in the sense that their mathematical models are identical (see universal oscillator equation above). Below is a table showing analogous quantities in four harmonic oscillator systems in mechanics and electronics. If analogous parameters on the same line in the table are given numerically equal values, the behavior of the oscillators – their output waveform, resonant frequency, damping factor, etc. – are the same.

Translational mechanicalRotational mechanicalSeries RLC circuitParallel RLC circuit
Position x{displaystyle x}Angle θ{displaystyle theta }Chargeq{displaystyle q}Flux linkageφ{displaystyle varphi }
Velocitydxdt{displaystyle {frac {mathrm {d} x}{mathrm {d} t}}}Angular velocitydθdt{displaystyle {frac {mathrm {d} theta }{mathrm {d} t}}}Currentdqdt{displaystyle {frac {mathrm {d} q}{mathrm {d} t}}}Voltagedφdt{displaystyle {frac {mathrm {d} varphi }{mathrm {d} t}}}
Massm{displaystyle m}Moment of inertiaI{displaystyle I}InductanceL{displaystyle L}CapacitanceC{displaystyle C}
Spring constantk{displaystyle k}Torsion constantμ{displaystyle mu }Elastance1/C{displaystyle 1/C}Magnetic reluctance1/L{displaystyle 1/L}
Dampingc{displaystyle c}Rotational frictionΓ{displaystyle Gamma }ResistanceR{displaystyle R}ConductanceG=1/R{displaystyle G=1/R}
Drive forceF(t){displaystyle F(t)}Drive torqueτ(t){displaystyle tau (t)}Voltagee{displaystyle e}Currenti{displaystyle i}
Undamped resonant frequencyfn{displaystyle f_{n}}:
12πkm{displaystyle {frac {1}{2pi }}{sqrt {frac {k}{m}}}}12πμI{displaystyle {frac {1}{2pi }}{sqrt {frac {mu }{I}}}}12π1LC{displaystyle {frac {1}{2pi }}{sqrt {frac {1}{LC}}}}12π1LC{displaystyle {frac {1}{2pi }}{sqrt {frac {1}{LC}}}}
Damping ratioζ{displaystyle zeta }:
c21km{displaystyle {frac {c}{2}}{sqrt {frac {1}{km}}}}Γ21Iμ{displaystyle {frac {Gamma }{2}}{sqrt {frac {1}{Imu }}}}R2CL{displaystyle {frac {R}{2}}{sqrt {frac {C}{L}}}}G2LC{displaystyle {frac {G}{2}}{sqrt {frac {L}{C}}}}
Differential equation:
mx¨+cx˙+kx=F{displaystyle m{ddot {x}}+c{dot {x}}+kx=F}Iθ¨+Γθ˙+μθ=τ{displaystyle I{ddot {theta }}+Gamma {dot {theta }}+mu theta =tau }Lq¨+Rq˙+q/C=e{displaystyle L{ddot {q}}+R{dot {q}}+q/C=e}Cφ¨+Gφ˙+φ/L=i{displaystyle C{ddot {varphi }}+G{dot {varphi }}+varphi /L=i}

Application to a conservative force[edit]

The problem of the simple harmonic oscillator occurs frequently in physics, because a mass at equilibrium under the influence of any conservative force, in the limit of small motions, behaves as a simple harmonic oscillator.

A conservative force is one that is associated with a potential energy. The potential-energy function of a harmonic oscillator is

V(x)=12kx2.{displaystyle V(x)={frac {1}{2}}kx^{2}.}

Given an arbitrary potential-energy function V(x){displaystyle V(x)}, one can do a Taylor expansion in terms of x{displaystyle x} around an energy minimum (x=x0{displaystyle x=x_{0}}) to model the behavior of small perturbations from equilibrium.

V(x)=V(x0)+V(x0)(xx0)+12V(2)(x0)(xx0)2+O(xx0)3.{displaystyle V(x)=V(x_{0})+V'(x_{0})cdot (x-x_{0})+{frac {1}{2}}V^{(2)}(x_{0})cdot (x-x_{0})^{2}+O(x-x_{0})^{3}.}

Because V(x0){displaystyle V(x_{0})} is a minimum, the first derivative evaluated at x0{displaystyle x_{0}} must be zero, so the linear term drops out:

V(x)=V(x0)+12V(2)(x0)(xx0)2+O(xx0)3.{displaystyle V(x)=V(x_{0})+{frac {1}{2}}V^{(2)}(x_{0})cdot (x-x_{0})^{2}+O(x-x_{0})^{3}.}

The constant termV(x0) is arbitrary and thus may be dropped, and a coordinate transformation allows the form of the simple harmonic oscillator to be retrieved:

V(x)12V(2)(0)x2=12kx2.{displaystyle V(x)approx {frac {1}{2}}V^{(2)}(0)cdot x^{2}={frac {1}{2}}kx^{2}.}

Thus, given an arbitrary potential-energy function V(x){displaystyle V(x)} with a non-vanishing second derivative, one can use the solution to the simple harmonic oscillator to provide an approximate solution for small perturbations around the equilibrium point.

Examples[edit]

Simple pendulum[edit]

A simple pendulum exhibits approximately simple harmonic motion under the conditions of no damping and small amplitude.

Assuming no damping, the differential equation governing a simple pendulum of length l{displaystyle l}, where g{displaystyle g} is the local acceleration of gravity, is

d2θdt2+glsinθ=0.{displaystyle {frac {d^{2}theta }{dt^{2}}}+{frac {g}{l}}sin theta =0.}

If the maximal displacement of the pendulum is small, we can use the approximation sinθθ{displaystyle sin theta approx theta } and instead consider the equation

d2θdt2+glθ=0.{displaystyle {frac {d^{2}theta }{dt^{2}}}+{frac {g}{l}}theta =0.}

The general solution to this differential equation is

θ(t)=Acos(glt+φ),{displaystyle theta (t)=Acos left({sqrt {frac {g}{l}}}t+varphi right),}

where A{displaystyle A} and φ{displaystyle varphi } are constants that depends on the initial conditions.Using as initial conditions θ(0)=θ0{displaystyle theta (0)=theta _{0}} and θ˙(0)=0{displaystyle {dot {theta }}(0)=0}, the solution is given by

θ(t)=θ0cos(glt),{displaystyle theta (t)=theta _{0}cos left({sqrt {frac {g}{l}}}tright),}

where θ0{displaystyle theta _{0}} is the largest angle attained by the pendulum (that is, θ0{displaystyle theta _{0}} is the amplitude of the pendulum). The period, the time for one complete oscillation, is given by the expression

τ=2πlg=2πω,{displaystyle tau =2pi {sqrt {frac {l}{g}}}={frac {2pi }{omega }},}

which is a good approximation of the actual period when θ0{displaystyle theta _{0}} is small. Notice that in this approximation the period τ{displaystyle tau } is independent of the amplitude θ0{displaystyle theta _{0}}. In the above equation, ω{displaystyle omega } represents the angular frequency.

Spring/mass system[edit]

Spring–mass system in equilibrium (A), compressed (B) and stretched (C) states

When a spring is stretched or compressed by a mass, the spring develops a restoring force. Hooke's law gives the relationship of the force exerted by the spring when the spring is compressed or stretched a certain length:

F(t)=kx(t),{displaystyle F(t)=-kx(t),}

where F is the force, k is the spring constant, and x is the displacement of the mass with respect to the equilibrium position. The minus sign in the equation indicates that the force exerted by the spring always acts in a direction that is opposite to the displacement (i.e. the force always acts towards the zero position), and so prevents the mass from flying off to infinity.

By using either force balance or an energy method, it can be readily shown that the motion of this system is given by the following differential equation:

F(t)=kx(t)=md2dt2x(t)=ma,{displaystyle F(t)=-kx(t)=m{frac {mathrm {d} ^{2}}{mathrm {d} t^{2}}}x(t)=ma,}

the latter being Newton's second law of motion.

If the initial displacement is A, and there is no initial velocity, the solution of this equation is given by

x(t)=Acos(kmt).{displaystyle x(t)=Acos left({sqrt {frac {k}{m}}}tright).}

Given an ideal massless spring, m{displaystyle m} is the mass on the end of the spring. If the spring itself has mass, its effective mass must be included in m{displaystyle m}.

Energy variation in the spring–damping system[edit]

In terms of energy, all systems have two types of energy: potential energy and kinetic energy. When a spring is stretched or compressed, it stores elastic potential energy, which then is transferred into kinetic energy. The potential energy within a spring is determined by the equation U=kx2/2.{displaystyle U=kx^{2}/2.}

When the spring is stretched or compressed, kinetic energy of the mass gets converted into potential energy of the spring. By conservation of energy, assuming the datum is defined at the equilibrium position, when the spring reaches its maximal potential energy, the kinetic energy of the mass is zero. When the spring is released, it tries to return to equilibrium, and all its potential energy converts to kinetic energy of the mass.

Definition of terms[edit]

SymbolDefinitionDimensionsSI units
a{displaystyle a}Acceleration of massLT2{displaystyle mathbf {LT^{-2}} }m/s2
A{displaystyle A}Peak amplitude of oscillationL{displaystyle mathbf {L} }m
c{displaystyle c}Viscous damping coefficientMT1{displaystyle mathbf {MT^{-1}} }N·s/m
f{displaystyle f}FrequencyT1{displaystyle mathbf {T^{-1}} }Hz
F{displaystyle F}Drive forceMLT2{displaystyle mathbf {MLT^{-2}} }N
g{displaystyle g}Acceleration of gravity at the Earth's surfaceLT2{displaystyle mathbf {LT^{-2}} }m/s2
i{displaystyle i}Imaginary unit, 1{displaystyle {sqrt {-1}}}
k{displaystyle k}Spring constantMT2{displaystyle mathbf {MT^{-2}} }N/m
m,M{displaystyle m,M}MassM{displaystyle mathbf {M} }kg
Q{displaystyle Q}Quality factor
T{displaystyle T}Period of oscillationT{displaystyle mathbf {T} }s
t{displaystyle t}TimeT{displaystyle mathbf {T} }s
U{displaystyle U}Potential energy stored in oscillatorML2T2{displaystyle mathbf {ML^{2}T^{-2}} }J
x{displaystyle x}Position of massL{displaystyle mathbf {L} }m
ζ{displaystyle zeta }Damping ratio
φ{displaystyle varphi }Phase shiftrad
ω{displaystyle omega }Angular frequencyT1{displaystyle mathbf {T^{-1}} }rad/s
ω0{displaystyle omega _{0}}Natural resonant angular frequencyT1{displaystyle mathbf {T^{-1}} }rad/s

See also[edit]

Simple Harmonic Oscillator Equation

Notes[edit]

  1. ^Fowles & Cassiday (1986, p. 86)
  2. ^Kreyszig (1972, p. 65)
  3. ^Tipler (1998, pp. 369,389)
  4. ^Case, William. 'Two ways of driving a child's swing'. Archived from the original on 9 December 2011. Retrieved 27 November 2011.
  5. ^Case, W. B. (1996). 'The pumping of a swing from the standing position'. American Journal of Physics. 64 (3): 215–220. Bibcode:1996AmJPh..64..215C. doi:10.1119/1.18209.
  6. ^Roura, P.; Gonzalez, J.A. (2010). 'Towards a more realistic description of swing pumping due to the exchange of angular momentum'. European Journal of Physics. 31 (5): 1195–1207. Bibcode:2010EJPh...31.1195R. doi:10.1088/0143-0807/31/5/020.

References[edit]

  • Fowles, Grant R.; Cassiday, George L. (1986), Analytic Mechanics (5th ed.), Fort Worth: Saunders College Publishing, ISBN0-03-96746-5, LCCN93085193CS1 maint: ignored ISBN errors (link)
  • Hayek, Sabih I. (15 Apr 2003). 'Mechanical Vibration and Damping'. Encyclopedia of Applied Physics. WILEY-VCH Verlag GmbH & Co KGaA. doi:10.1002/3527600434.eap231. ISBN9783527600434.
  • Kreyszig, Erwin (1972), Advanced Engineering Mathematics (3rd ed.), New York: Wiley, ISBN0-471-50728-8
  • Serway, Raymond A.; Jewett, John W. (2003). Physics for Scientists and Engineers. Brooks/Cole. ISBN0-534-40842-7.
  • Tipler, Paul (1998). Physics for Scientists and Engineers: Vol. 1 (4th ed.). W. H. Freeman. ISBN1-57259-492-6.
  • Wylie, C. R. (1975). Advanced Engineering Mathematics (4th ed.). McGraw-Hill. ISBN0-07-072180-7.

External links[edit]

Wikimedia Commons has media related to Harmonic oscillation.
  • The Harmonic Oscillator from The Feynman Lectures on Physics
  • Hazewinkel, Michiel, ed. (2001) [1994], 'Oscillator, harmonic', Encyclopedia of Mathematics, Springer Science+Business Media B.V. / Kluwer Academic Publishers, ISBN978-1-55608-010-4
  • Harmonic Oscillator from The Chaos Hypertextbook
Retrieved from 'https://en.wikipedia.org/w/index.php?title=Harmonic_oscillator&oldid=916699414'