fbpx
Wikipedia

Probability amplitude

In quantum mechanics, a probability amplitude is a complex number used for describing the behaviour of systems. The square of the modulus of this quantity represents a probability density.

A wave function for a single electron on 5d atomic orbital of a hydrogen atom. The solid body shows the places where the electron's probability density is above a certain value (here 0.02 nm−3): this is calculated from the probability amplitude. The hue on the colored surface shows the complex phase of the wave function.

Probability amplitudes provide a relationship between the quantum state vector of a system and the results of observations of that system, a link was first proposed by Max Born, in 1926. Interpretation of values of a wave function as the probability amplitude is a pillar of the Copenhagen interpretation of quantum mechanics. In fact, the properties of the space of wave functions were being used to make physical predictions (such as emissions from atoms being at certain discrete energies) before any physical interpretation of a particular function was offered. Born was awarded half of the 1954 Nobel Prize in Physics for this understanding, and the probability thus calculated is sometimes called the "Born probability". These probabilistic concepts, namely the probability density and quantum measurements, were vigorously contested at the time by the original physicists working on the theory, such as Schrödinger and Einstein. It is the source of the mysterious consequences and philosophical difficulties in the interpretations of quantum mechanics—topics that continue to be debated even today.

Physical overview edit

Neglecting some technical complexities, the problem of quantum measurement is the behaviour of a quantum state, for which the value of the observable Q to be measured is uncertain. Such a state is thought to be a coherent superposition of the observable's eigenstates, states on which the value of the observable is uniquely defined, for different possible values of the observable.

When a measurement of Q is made, the system (under the Copenhagen interpretation) jumps to one of the eigenstates, returning the eigenvalue belonging to that eigenstate. The system may always be described by a linear combination or superposition of these eigenstates with unequal "weights". Intuitively it is clear that eigenstates with heavier "weights" are more "likely" to be produced. Indeed, which of the above eigenstates the system jumps to is given by a probabilistic law: the probability of the system jumping to the state is proportional to the absolute value of the corresponding numerical weight squared. These numerical weights are called probability amplitudes, and this relationship used to calculate probabilities from given pure quantum states (such as wave functions) is called the Born rule.

Clearly, the sum of the probabilities, which equals the sum of the absolute squares of the probability amplitudes, must equal 1. This is the normalization requirement.

If the system is known to be in some eigenstate of Q (e.g. after an observation of the corresponding eigenvalue of Q) the probability of observing that eigenvalue becomes equal to 1 (certain) for all subsequent measurements of Q (so long as no other important forces act between the measurements). In other words, the probability amplitudes are zero for all the other eigenstates, and remain zero for the future measurements. If the set of eigenstates to which the system can jump upon measurement of Q is the same as the set of eigenstates for measurement of R, then subsequent measurements of either Q or R always produce the same values with probability of 1, no matter the order in which they are applied. The probability amplitudes are unaffected by either measurement, and the observables are said to commute.

By contrast, if the eigenstates of Q and R are different, then measurement of R produces a jump to a state that is not an eigenstate of Q. Therefore, if the system is known to be in some eigenstate of Q (all probability amplitudes zero except for one eigenstate), then when R is observed the probability amplitudes are changed. A second, subsequent observation of Q no longer certainly produces the eigenvalue corresponding to the starting state. In other words, the probability amplitudes for the second measurement of Q depend on whether it comes before or after a measurement of R, and the two observables do not commute.

Mathematical formulation edit

In a formal setup, the state of an isolated physical system in quantum mechanics is represented, at a fixed time  , by a state vector |Ψ⟩ belonging to a separable complex Hilbert space. Using bra–ket notation the relation between state vector and "position basis"   of the Hilbert space can be written as[1]

 .

Its relation with an observable can be elucidated by generalizing the quantum state   to a measurable function and its domain of definition to a given σ-finite measure space  . This allows for a refinement of Lebesgue's decomposition theorem, decomposing μ into three mutually singular parts

 

where μac is absolutely continuous with respect to the Lebesgue measure, μsc is singular with respect to the Lebesgue measure and atomless, and μpp is a pure point measure.[2][3]

Continuous amplitudes edit

A usual presentation of the probability amplitude is that of a wave function   belonging to the L2 space of (equivalence classes of) square integrable functions, i.e.,   belongs to L2(X) if and only if

 .

If the norm is equal to 1 and   such that

 ,

then   is the probability density function for a measurement of the particle's position at a given time, defined as the Radon–Nikodym derivative with respect to the Lebesgue measure (e.g. on the set R of all real numbers). As probability is a dimensionless quantity, |ψ(x)|2 must have the inverse dimension of the variable of integration x. For example, the above amplitude has dimension [L−1/2], where L represents length.

Whereas a Hilbert space is separable if and only if it admits a countable orthonormal basis, the range of a continuous random variable   is an uncountable set (i.e. the probability that the system is "at position  " will always be zero). As such, eigenstates of an observable need not necessarily be measurable functions belonging to L2(X) (see normalization condition below). A typical example is the position operator   defined as

 

whose eigenfunctions are Dirac delta functions

 

which clearly do not belong to L2(X). By replacing the state space by a suitable rigged Hilbert space, however, the rigorous notion of eigenstates from spectral theorem as well as spectral decomposition is preserved.[4]

Discrete amplitudes edit

Let   be atomic (i.e. the set   in   is an atom); specifying the measure of any discrete variable xA equal to 1. The amplitudes are composed of state vector |Ψ⟩ indexed by A; its components are denoted by ψ(x) for uniformity with the previous case. If the 2-norm of |Ψ⟩ is equal to 1, then |ψ(x)|2 is a probability mass function.

A convenient configuration space X is such that each point x produces some unique value of the observable Q. For discrete X it means that all elements of the standard basis are eigenvectors of Q. Then   is the probability amplitude for the eigenstate |x. If it corresponds to a non-degenerate eigenvalue of Q, then   gives the probability of the corresponding value of Q for the initial state |Ψ⟩.

|ψ(x)| = 1 if and only if |x is the same quantum state as |Ψ⟩. ψ(x) = 0 if and only if |x and |Ψ⟩ are orthogonal. Otherwise the modulus of ψ(x) is between 0 and 1.

A discrete probability amplitude may be considered as a fundamental frequency in the probability frequency domain (spherical harmonics) for the purposes of simplifying M-theory transformation calculations.[citation needed] Discrete dynamical variables are used in such problems as a particle in an idealized reflective box and quantum harmonic oscillator.[clarification needed]

Examples edit

An example of the discrete case is a quantum system that can be in two possible states, e.g. the polarization of a photon. When the polarization is measured, it could be the horizontal state   or the vertical state  . Until its polarization is measured the photon can be in a superposition of both these states, so its state   could be written as

 ,

with   and   the probability amplitudes for the states   and   respectively. When the photon's polarization is measured, the resulting state is either horizontal or vertical. But in a random experiment, the probability of being horizontally polarized is  , and the probability of being vertically polarized is  .

Hence, a photon in a state   would have a probability of   to come out horizontally polarized, and a probability of   to come out vertically polarized when an ensemble of measurements are made. The order of such results, is, however, completely random.

Another example is quantum spin. If a spin-measuring apparatus is pointing along the z-axis and is therefore able to measure the z-component of the spin ( ), the following must be true for the measurement of spin "up" and "down":

 
 

If one assumes that system is prepared, so that +1 is registered in   and then the apparatus is rotated to measure  , the following holds:

 

The probability amplitude of measuring spin up is given by  , since the system had the initial state  . The probability of measuring   is given by

 

Which agrees with experiment.

Normalization edit

In the example above, the measurement must give either | H ⟩ or | V ⟩, so the total probability of measuring | H ⟩ or | V ⟩ must be 1. This leads to a constraint that α2 + β2 = 1; more generally the sum of the squared moduli of the probability amplitudes of all the possible states is equal to one. If to understand "all the possible states" as an orthonormal basis, that makes sense in the discrete case, then this condition is the same as the norm-1 condition explained above.

One can always divide any non-zero element of a Hilbert space by its norm and obtain a normalized state vector. Not every wave function belongs to the Hilbert space L2(X), though. Wave functions that fulfill this constraint are called normalizable.

The Schrödinger equation, describing states of quantum particles, has solutions that describe a system and determine precisely how the state changes with time. Suppose a wave function ψ(x, t) gives a description of the particle (position x at a given time t). A wave function is square integrable if

 

After normalization the wave function still represents the same state and is therefore equal by definition to[5][6]

 

Under the standard Copenhagen interpretation, the normalized wavefunction gives probability amplitudes for the position of the particle. Hence, ρ(x) = |ψ(x, t)|2 is a probability density function and the probability that the particle is in the volume V at fixed time t is given by

 

The probability density function does not vary with time as the evolution of the wave function is dictated by the Schrödinger equation and is therefore entirely deterministic.[7] This is key to understanding the importance of this interpretation: for a given particle constant mass, initial ψ(x, t0) and potential, the Schrödinger equation fully determines subsequent wavefunctions. The above then gives probabilities of locations of the particle at all subsequent times.

In the context of the double-slit experiment edit

Probability amplitudes have special significance because they act in quantum mechanics as the equivalent of conventional probabilities, with many analogous laws, as described above. For example, in the classic double-slit experiment, electrons are fired randomly at two slits, and the probability distribution of detecting electrons at all parts on a large screen placed behind the slits, is questioned. An intuitive answer is that P(through either slit) = P(through first slit) + P(through second slit), where P(event) is the probability of that event. This is obvious if one assumes that an electron passes through either slit. When no measurement apparatus that determines through which slit the electrons travel is installed, the observed probability distribution on the screen reflects the interference pattern that is common with light waves. If one assumes the above law to be true, then this pattern cannot be explained. The particles cannot be said to go through either slit and the simple explanation does not work. The correct explanation is, however, by the association of probability amplitudes to each event. The complex amplitudes which represent the electron passing each slit (ψfirst and ψsecond) follow the law of precisely the form expected: ψtotal = ψfirst + ψsecond. This is the principle of quantum superposition. The probability, which is the modulus squared of the probability amplitude, then, follows the interference pattern under the requirement that amplitudes are complex:

 
Here,   and   are the arguments of ψfirst and ψsecond respectively. A purely real formulation has too few dimensions to describe the system's state when superposition is taken into account. That is, without the arguments of the amplitudes, we cannot describe the phase-dependent interference. The crucial term   is called the "interference term", and this would be missing if we had added the probabilities.

However, one may choose to devise an experiment in which the experimenter observes which slit each electron goes through. Then, due to wavefunction collapse, the interference pattern is not observed on the screen.

One may go further in devising an experiment in which the experimenter gets rid of this "which-path information" by a "quantum eraser". Then, according to the Copenhagen interpretation, the case A applies again and the interference pattern is restored.[8]

Conservation of probabilities and the continuity equation edit

Intuitively, since a normalised wave function stays normalised while evolving according to the wave equation, there will be a relationship between the change in the probability density of the particle's position and the change in the amplitude at these positions.

Define the probability current (or flux) j as

 

measured in units of (probability)/(area × time).

Then the current satisfies the equation

 

The probability density is  , this equation is exactly the continuity equation, appearing in many situations in physics where we need to describe the local conservation of quantities. The best example is in classical electrodynamics, where j corresponds to current density corresponding to electric charge, and the density is the charge-density. The corresponding continuity equation describes the local conservation of charges.

Composite systems edit

For two quantum systems with spaces L2(X1) and L2(X2) and given states 1 and 2 respectively, their combined state 12 can be expressed as ψ1(x1) ψ2(x2) a function on X1×X2, that gives the product of respective probability measures. In other words, amplitudes of a non-entangled composite state are products of original amplitudes, and respective observables on the systems 1 and 2 behave on these states as independent random variables. This strengthens the probabilistic interpretation explicated above .

Amplitudes in operators edit

The concept of amplitudes is also used in the context of scattering theory, notably in the form of S-matrices. Whereas moduli of vector components squared, for a given vector, give a fixed probability distribution, moduli of matrix elements squared are interpreted as transition probabilities just as in a random process. Like a finite-dimensional unit vector specifies a finite probability distribution, a finite-dimensional unitary matrix specifies transition probabilities between a finite number of states.

The "transitional" interpretation may be applied to L2s on non-discrete spaces as well.[clarification needed]

See also edit

Notes edit

  1. ^ The spanning set of a Hilbert space does not suffice for defining coordinates as wave functions form rays in a projective Hilbert space (rather than an ordinary Hilbert space). See: Projective frame
  2. ^ Simon 2005, p. 43.
  3. ^ Teschl 2014, p. 114-119.
  4. ^ de la Madrid Modino 2001, p. 97.
  5. ^ Bäuerle & de Kerf 1990, p. 330.
  6. ^ See also Wigner's theorem
  7. ^ Zwiebach 2022, p. 78.
  8. ^ A recent 2013 experiment gives insight regarding the correct physical interpretation of such phenomena. The information can actually be obtained, but then the electron seemingly went through all the possible paths simultaneously. (Certain ensemble-alike realistic interpretations of the wavefunction may presume such coexistence in all the points of an orbital.) Cf. Schmidt, L. Ph. H.; et al. (2013). (PDF). Physical Review Letters. 111 (10): 103201. Bibcode:2013PhRvL.111j3201S. doi:10.1103/PhysRevLett.111.103201. PMID 25166663. S2CID 2725093. Archived from the original (PDF) on 2019-03-07.

References edit

  • Bäuerle, Gerard G. A.; de Kerf, Eddy A. (1990). Lie Algebras, Part 1: Finite and Infinite Dimensional Lie Algebras and Applications in Physics. Studies in Mathematical Physics. Amsterdam: North Holland. ISBN 0-444-88776-8.
  • de la Madrid Modino, R. (2001). Quantum mechanics in rigged Hilbert space language (PhD thesis). Universidad de Valladolid.
  • Feynman, R. P.; Leighton, R. B.; Sands, M. (1989). "Probability Amplitudes". The Feynman Lectures on Physics. Vol. 3. Redwood City: Addison-Wesley. ISBN 0-201-51005-7.
  • Gudder, Stanley P. (1988). Quantum Probability. San Diego: Academic Press. ISBN 0-12-305340-4.
  • Simon, Barry (2005). Orthogonal polynomials on the unit circle. Part 1. Classical theory. American Mathematical Society Colloquium Publications. Vol. 54. Providence, R.I.: American Mathematical Society. ISBN 978-0-8218-3446-6. MR 2105088.
  • Teschl, G. (2014). Mathematical Methods in Quantum Mechanics. Providence (R.I): American Mathematical Soc. ISBN 978-1-4704-1704-8.
  • Zwiebach, Barton (2022). Mastering Quantum Mechanics. Cambridge, Mass: MIT Press. ISBN 978-0-262-04613-8.

probability, amplitude, this, article, about, probability, amplitude, quantum, mechanics, other, uses, amplitude, disambiguation, this, article, needs, additional, citations, verification, please, help, improve, this, article, adding, citations, reliable, sour. This article is about probability amplitude in quantum mechanics For other uses see Amplitude disambiguation This article needs additional citations for verification Please help improve this article by adding citations to reliable sources Unsourced material may be challenged and removed Find sources Probability amplitude news newspapers books scholar JSTOR January 2014 Learn how and when to remove this message In quantum mechanics a probability amplitude is a complex number used for describing the behaviour of systems The square of the modulus of this quantity represents a probability density A wave function for a single electron on 5d atomic orbital of a hydrogen atom The solid body shows the places where the electron s probability density is above a certain value here 0 02 nm 3 this is calculated from the probability amplitude The hue on the colored surface shows the complex phase of the wave function Probability amplitudes provide a relationship between the quantum state vector of a system and the results of observations of that system a link was first proposed by Max Born in 1926 Interpretation of values of a wave function as the probability amplitude is a pillar of the Copenhagen interpretation of quantum mechanics In fact the properties of the space of wave functions were being used to make physical predictions such as emissions from atoms being at certain discrete energies before any physical interpretation of a particular function was offered Born was awarded half of the 1954 Nobel Prize in Physics for this understanding and the probability thus calculated is sometimes called the Born probability These probabilistic concepts namely the probability density and quantum measurements were vigorously contested at the time by the original physicists working on the theory such as Schrodinger and Einstein It is the source of the mysterious consequences and philosophical difficulties in the interpretations of quantum mechanics topics that continue to be debated even today Contents 1 Physical overview 2 Mathematical formulation 2 1 Continuous amplitudes 2 2 Discrete amplitudes 2 3 Examples 3 Normalization 4 In the context of the double slit experiment 5 Conservation of probabilities and the continuity equation 6 Composite systems 7 Amplitudes in operators 8 See also 9 Notes 10 ReferencesPhysical overview editNeglecting some technical complexities the problem of quantum measurement is the behaviour of a quantum state for which the value of the observable Q to be measured is uncertain Such a state is thought to be a coherent superposition of the observable s eigenstates states on which the value of the observable is uniquely defined for different possible values of the observable When a measurement of Q is made the system under the Copenhagen interpretation jumps to one of the eigenstates returning the eigenvalue belonging to that eigenstate The system may always be described by a linear combination or superposition of these eigenstates with unequal weights Intuitively it is clear that eigenstates with heavier weights are more likely to be produced Indeed which of the above eigenstates the system jumps to is given by a probabilistic law the probability of the system jumping to the state is proportional to the absolute value of the corresponding numerical weight squared These numerical weights are called probability amplitudes and this relationship used to calculate probabilities from given pure quantum states such as wave functions is called the Born rule Clearly the sum of the probabilities which equals the sum of the absolute squares of the probability amplitudes must equal 1 This is the normalization requirement If the system is known to be in some eigenstate of Q e g after an observation of the corresponding eigenvalue of Q the probability of observing that eigenvalue becomes equal to 1 certain for all subsequent measurements of Q so long as no other important forces act between the measurements In other words the probability amplitudes are zero for all the other eigenstates and remain zero for the future measurements If the set of eigenstates to which the system can jump upon measurement of Q is the same as the set of eigenstates for measurement of R then subsequent measurements of either Q or R always produce the same values with probability of 1 no matter the order in which they are applied The probability amplitudes are unaffected by either measurement and the observables are said to commute By contrast if the eigenstates of Q and R are different then measurement of R produces a jump to a state that is not an eigenstate of Q Therefore if the system is known to be in some eigenstate of Q all probability amplitudes zero except for one eigenstate then when R is observed the probability amplitudes are changed A second subsequent observation of Q no longer certainly produces the eigenvalue corresponding to the starting state In other words the probability amplitudes for the second measurement of Q depend on whether it comes before or after a measurement of R and the two observables do not commute Mathematical formulation editSee also Bound state Definition In a formal setup the state of an isolated physical system in quantum mechanics is represented at a fixed time t displaystyle t nbsp by a state vector PS belonging to a separable complex Hilbert space Using bra ket notation the relation between state vector and position basis x displaystyle x rangle nbsp of the Hilbert space can be written as 1 ps x x PS displaystyle psi x langle x Psi rangle nbsp Its relation with an observable can be elucidated by generalizing the quantum state ps displaystyle psi nbsp to a measurable function and its domain of definition to a given s finite measure space X A m displaystyle X mathcal A mu nbsp This allows for a refinement of Lebesgue s decomposition theorem decomposing m into three mutually singular parts m m a c m s c m p p displaystyle mu mu mathrm ac mu mathrm sc mu mathrm pp nbsp where mac is absolutely continuous with respect to the Lebesgue measure msc is singular with respect to the Lebesgue measure and atomless and mpp is a pure point measure 2 3 Continuous amplitudes edit A usual presentation of the probability amplitude is that of a wave function ps displaystyle psi nbsp belonging to the L2 space of equivalence classes of square integrable functions i e ps displaystyle psi nbsp belongs to L2 X if and only if ps 2 X ps x 2 d x lt displaystyle psi 2 int X psi x 2 dx lt infty nbsp If the norm is equal to 1 and ps x 2 R 0 displaystyle psi x 2 in mathbb R geq 0 nbsp such that X ps x 2 d x X d m a c x 1 displaystyle int X psi x 2 dx equiv int X d mu ac x 1 nbsp then ps x 2 displaystyle psi x 2 nbsp is the probability density function for a measurement of the particle s position at a given time defined as the Radon Nikodym derivative with respect to the Lebesgue measure e g on the set R of all real numbers As probability is a dimensionless quantity ps x 2 must have the inverse dimension of the variable of integration x For example the above amplitude has dimension L 1 2 where L represents length Whereas a Hilbert space is separable if and only if it admits a countable orthonormal basis the range of a continuous random variable x displaystyle x nbsp is an uncountable set i e the probability that the system is at position x displaystyle x nbsp will always be zero As such eigenstates of an observable need not necessarily be measurable functions belonging to L2 X see normalization condition below A typical example is the position operator x displaystyle hat mathrm x nbsp defined as x x PS x x PS x 0 ps x x R displaystyle langle x hat mathrm x Psi rangle hat mathrm x langle x Psi rangle x 0 psi x quad x in mathbb R nbsp whose eigenfunctions are Dirac delta functions ps x d x x 0 displaystyle psi x delta x x 0 nbsp which clearly do not belong to L2 X By replacing the state space by a suitable rigged Hilbert space however the rigorous notion of eigenstates from spectral theorem as well as spectral decomposition is preserved 4 Discrete amplitudes edit Let m p p displaystyle mu pp nbsp be atomic i e the set A X displaystyle A subset X nbsp in A displaystyle mathcal A nbsp is an atom specifying the measure of any discrete variable x A equal to 1 The amplitudes are composed of state vector PS indexed by A its components are denoted by ps x for uniformity with the previous case If the ℓ2 norm of PS is equal to 1 then ps x 2 is a probability mass function A convenient configuration space X is such that each point x produces some unique value of the observable Q For discrete X it means that all elements of the standard basis are eigenvectors of Q Then ps x displaystyle psi x nbsp is the probability amplitude for the eigenstate x If it corresponds to a non degenerate eigenvalue of Q then ps x 2 displaystyle psi x 2 nbsp gives the probability of the corresponding value of Q for the initial state PS ps x 1 if and only if x is the same quantum state as PS ps x 0 if and only if x and PS are orthogonal Otherwise the modulus of ps x is between 0 and 1 A discrete probability amplitude may be considered as a fundamental frequency in the probability frequency domain spherical harmonics for the purposes of simplifying M theory transformation calculations citation needed Discrete dynamical variables are used in such problems as a particle in an idealized reflective box and quantum harmonic oscillator clarification needed Examples edit An example of the discrete case is a quantum system that can be in two possible states e g the polarization of a photon When the polarization is measured it could be the horizontal state H displaystyle H rangle nbsp or the vertical state V displaystyle V rangle nbsp Until its polarization is measured the photon can be in a superposition of both these states so its state ps displaystyle psi rangle nbsp could be written as ps a H b V displaystyle psi rangle alpha H rangle beta V rangle nbsp with a displaystyle alpha nbsp and b displaystyle beta nbsp the probability amplitudes for the states H displaystyle H rangle nbsp and V displaystyle V rangle nbsp respectively When the photon s polarization is measured the resulting state is either horizontal or vertical But in a random experiment the probability of being horizontally polarized is a 2 displaystyle alpha 2 nbsp and the probability of being vertically polarized is b 2 displaystyle beta 2 nbsp Hence a photon in a state ps 1 3 H i 2 3 V textstyle psi rangle sqrt frac 1 3 H rangle i sqrt frac 2 3 V rangle nbsp would have a probability of 1 3 textstyle frac 1 3 nbsp to come out horizontally polarized and a probability of 2 3 textstyle frac 2 3 nbsp to come out vertically polarized when an ensemble of measurements are made The order of such results is however completely random Another example is quantum spin If a spin measuring apparatus is pointing along the z axis and is therefore able to measure the z component of the spin s z textstyle sigma z nbsp the following must be true for the measurement of spin up and down s z u 1 u displaystyle sigma z u rangle 1 u rangle nbsp s z d 1 d displaystyle sigma z d rangle 1 d rangle nbsp If one assumes that system is prepared so that 1 is registered in s x textstyle sigma x nbsp and then the apparatus is rotated to measure s z textstyle sigma z nbsp the following holds r u 1 2 u 1 2 d u 1 2 1 0 1 2 0 1 1 0 1 2 displaystyle begin aligned langle r u rangle amp left frac 1 sqrt 2 langle u frac 1 sqrt 2 langle d right cdot u rangle amp left frac 1 sqrt 2 begin pmatrix 1 0 end pmatrix frac 1 sqrt 2 begin pmatrix 0 1 end pmatrix right cdot begin pmatrix 1 0 end pmatrix amp frac 1 sqrt 2 end aligned nbsp The probability amplitude of measuring spin up is given by r u textstyle langle r u rangle nbsp since the system had the initial state r textstyle r rangle nbsp The probability of measuring u textstyle u rangle nbsp is given by P u r u u r 1 2 2 1 2 displaystyle P u rangle langle r u rangle langle u r rangle left frac 1 sqrt 2 right 2 frac 1 2 nbsp Which agrees with experiment Normalization editIn the example above the measurement must give either H or V so the total probability of measuring H or V must be 1 This leads to a constraint that a2 b2 1 more generally the sum of the squared moduli of the probability amplitudes of all the possible states is equal to one If to understand all the possible states as an orthonormal basis that makes sense in the discrete case then this condition is the same as the norm 1 condition explained above One can always divide any non zero element of a Hilbert space by its norm and obtain a normalized state vector Not every wave function belongs to the Hilbert space L2 X though Wave functions that fulfill this constraint are called normalizable The Schrodinger equation describing states of quantum particles has solutions that describe a system and determine precisely how the state changes with time Suppose a wave function ps x t gives a description of the particle position x at a given time t A wave function is square integrable if ps x t 2 d x a 2 lt displaystyle int psi mathbf x t 2 mathrm d mathbf x a 2 lt infty nbsp After normalization the wave function still represents the same state and is therefore equal by definition to 5 6 ps x t ps x t a displaystyle psi mathbf x t frac psi mathbf x t a nbsp Under the standard Copenhagen interpretation the normalized wavefunction gives probability amplitudes for the position of the particle Hence r x ps x t 2 is a probability density function and the probability that the particle is in the volume V at fixed time t is given by P x V t V ps x t 2 d x V r x d x displaystyle P mathbf x in V t int V psi mathbf x t 2 mathrm d mathbf x int V rho mathbf x mathrm d mathbf x nbsp The probability density function does not vary with time as the evolution of the wave function is dictated by the Schrodinger equation and is therefore entirely deterministic 7 This is key to understanding the importance of this interpretation for a given particle constant mass initial ps x t0 and potential the Schrodinger equation fully determines subsequent wavefunctions The above then gives probabilities of locations of the particle at all subsequent times In the context of the double slit experiment editMain article Double slit experiment Probability amplitudes have special significance because they act in quantum mechanics as the equivalent of conventional probabilities with many analogous laws as described above For example in the classic double slit experiment electrons are fired randomly at two slits and the probability distribution of detecting electrons at all parts on a large screen placed behind the slits is questioned An intuitive answer is that P through either slit P through first slit P through second slit where P event is the probability of that event This is obvious if one assumes that an electron passes through either slit When no measurement apparatus that determines through which slit the electrons travel is installed the observed probability distribution on the screen reflects the interference pattern that is common with light waves If one assumes the above law to be true then this pattern cannot be explained The particles cannot be said to go through either slit and the simple explanation does not work The correct explanation is however by the association of probability amplitudes to each event The complex amplitudes which represent the electron passing each slit psfirst and pssecond follow the law of precisely the form expected pstotal psfirst pssecond This is the principle of quantum superposition The probability which is the modulus squared of the probability amplitude then follows the interference pattern under the requirement that amplitudes are complex P ps first ps second 2 ps first 2 ps second 2 2 ps first ps second cos f 1 f 2 displaystyle P left psi text first psi text second right 2 left psi text first right 2 left psi text second right 2 2 left psi text first right left psi text second right cos varphi 1 varphi 2 nbsp Here f 1 displaystyle varphi 1 nbsp and f 2 displaystyle varphi 2 nbsp are the arguments of psfirst and pssecond respectively A purely real formulation has too few dimensions to describe the system s state when superposition is taken into account That is without the arguments of the amplitudes we cannot describe the phase dependent interference The crucial term 2 ps first ps second cos f 1 f 2 textstyle 2 left psi text first right left psi text second right cos varphi 1 varphi 2 nbsp is called the interference term and this would be missing if we had added the probabilities However one may choose to devise an experiment in which the experimenter observes which slit each electron goes through Then due to wavefunction collapse the interference pattern is not observed on the screen One may go further in devising an experiment in which the experimenter gets rid of this which path information by a quantum eraser Then according to the Copenhagen interpretation the case A applies again and the interference pattern is restored 8 Conservation of probabilities and the continuity equation editMain article Probability current Intuitively since a normalised wave function stays normalised while evolving according to the wave equation there will be a relationship between the change in the probability density of the particle s position and the change in the amplitude at these positions Define the probability current or flux j as j ℏ m 1 2 i ps ps ps ps ℏ m Im ps ps displaystyle mathbf j hbar over m 1 over 2i left psi nabla psi psi nabla psi right hbar over m operatorname Im left psi nabla psi right nbsp measured in units of probability area time Then the current satisfies the equation j t ps 2 0 displaystyle nabla cdot mathbf j partial over partial t psi 2 0 nbsp The probability density is r ps 2 displaystyle rho psi 2 nbsp this equation is exactly the continuity equation appearing in many situations in physics where we need to describe the local conservation of quantities The best example is in classical electrodynamics where j corresponds to current density corresponding to electric charge and the density is the charge density The corresponding continuity equation describes the local conservation of charges Composite systems editFor two quantum systems with spaces L2 X1 and L2 X2 and given states PS1 and PS2 respectively their combined state PS1 PS2 can be expressed as ps1 x1 ps2 x2 a function on X1 X2 that gives the product of respective probability measures In other words amplitudes of a non entangled composite state are products of original amplitudes and respective observables on the systems 1 and 2 behave on these states as independent random variables This strengthens the probabilistic interpretation explicated above Amplitudes in operators editThe concept of amplitudes is also used in the context of scattering theory notably in the form of S matrices Whereas moduli of vector components squared for a given vector give a fixed probability distribution moduli of matrix elements squared are interpreted as transition probabilities just as in a random process Like a finite dimensional unit vector specifies a finite probability distribution a finite dimensional unitary matrix specifies transition probabilities between a finite number of states The transitional interpretation may be applied to L2 s on non discrete spaces as well clarification needed See also editExpectation value quantum mechanics Free particle Finite potential barrier Matter wave Phase space formulation Uncertainty principle Ward s probability amplitude Wave packetNotes edit The spanning set of a Hilbert space does not suffice for defining coordinates as wave functions form rays in a projective Hilbert space rather than an ordinary Hilbert space See Projective frame Simon 2005 p 43 Teschl 2014 p 114 119 de la Madrid Modino 2001 p 97 Bauerle amp de Kerf 1990 p 330 See also Wigner s theorem Zwiebach 2022 p 78 A recent 2013 experiment gives insight regarding the correct physical interpretation of such phenomena The information can actually be obtained but then the electron seemingly went through all the possible paths simultaneously Certain ensemble alike realistic interpretations of the wavefunction may presume such coexistence in all the points of an orbital Cf Schmidt L Ph H et al 2013 Momentum Transfer to a Free Floating Double Slit Realization of a Thought Experiment from the Einstein Bohr Debates PDF Physical Review Letters 111 10 103201 Bibcode 2013PhRvL 111j3201S doi 10 1103 PhysRevLett 111 103201 PMID 25166663 S2CID 2725093 Archived from the original PDF on 2019 03 07 References editBauerle Gerard G A de Kerf Eddy A 1990 Lie Algebras Part 1 Finite and Infinite Dimensional Lie Algebras and Applications in Physics Studies in Mathematical Physics Amsterdam North Holland ISBN 0 444 88776 8 de la Madrid Modino R 2001 Quantum mechanics in rigged Hilbert space language PhD thesis Universidad de Valladolid Feynman R P Leighton R B Sands M 1989 Probability Amplitudes The Feynman Lectures on Physics Vol 3 Redwood City Addison Wesley ISBN 0 201 51005 7 Gudder Stanley P 1988 Quantum Probability San Diego Academic Press ISBN 0 12 305340 4 Simon Barry 2005 Orthogonal polynomials on the unit circle Part 1 Classical theory American Mathematical Society Colloquium Publications Vol 54 Providence R I American Mathematical Society ISBN 978 0 8218 3446 6 MR 2105088 Teschl G 2014 Mathematical Methods in Quantum Mechanics Providence R I American Mathematical Soc ISBN 978 1 4704 1704 8 Zwiebach Barton 2022 Mastering Quantum Mechanics Cambridge Mass MIT Press ISBN 978 0 262 04613 8 Retrieved from https en wikipedia org w index php title Probability amplitude amp oldid 1211292813, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.