Big Chemical Encyclopedia

Chemical substances, components, reactions, process design ...

Articles Figures Tables About

Expectation values of observables

Wavefunctions that are normalized and orthogonal are said to be orthonormal. Symbolically, [Pg.21]

A simple exercise shows that the eigenvalue o of an operator with respect to a certain eigenfunction ijj can be extracted from Eq. (2.1) by multiplying both members of the eigenvalue equation on the left by integrating and using Eq. (2.11)  [Pg.21]

Therefore, we can say that the eigenfunctions of an operator O diagonalize the matrix O, whose elements are [Pg.21]

However, the state of a system is not an eigenstate of all operators. When the system wavefunction ip is not an eigenfunction of the operator A representing a certain observable A, the outcomes of a series of identical experiments are not identical. The average of all outcomes (see Chapter 1) is the expected value or the expectation value of the operator with respect to the state under consideration  [Pg.22]

This value is equivalent to a classical average of the physical observable at a given time, where each contribution is weighted by the corresponding probability of occurrence given by the values of (it is assumed that tp is normalized). An example is the mean distance of the electron from the nucleus of atom H, for a given orbital  [Pg.22]


We will consider the expectation value of observables G (aj, i), depending on polynomials of the creation and annihilation operators aj, ai. From Eq. (42) we obtain... [Pg.143]

The Hohenberg-Kohn theorem can be used to redefine entanglement measures in terms of new physical quantities expectation values of observables, ai, instead of external control parameters, li. Consider an arbitrary entanglement measure M for the ground state of Hamiltonian (85). For a bipartite entanglement, one can prove a central lemma, which very generally connects M and energy derivatives. [Pg.529]

Spatial extension, as expressed by the expectation value (r), is roughly comparable for 4 f and 5 f wave functions (Figs. 7 and 8). However, the many-electron wave functions resulting from the solution of the relativistic Dirac equation may also be used to calculate a number of physically interesting quantities, i.e. expectation values of observable... [Pg.19]

Just as logarithms and exponentials are inverse operations, integration is the inverse of differentiation. The integral can be shown to be the area under the curve in the same sense that the derivative is the slope of the tangent to the curve. The most common applications of integrals in chemistry and physics are normalization (for example, adjusting a probability distribution so that the sum of all the probabilities is 1) and calculation of the expectation values of observable quantities. [Pg.27]

Equation 6.8 does not always have to be satisfied f(x) does not have to be a stationary state. However, if f(x) does not satisfy Equation 6.8, the probability distribution P(x) and the expectation values of observables will change with time. The stationary states of a system constitute a complete basis set—which just means that any wavefunction f can be written as a superposition of the stationary states ... [Pg.132]

With a slight modification, Eq. (10.13) provides a powerful way to deal with an ensemble of many systems. The expectation value of observable A for an ensemble is... [Pg.422]

It is now required for observable quantities that the expectation value of any operator O taken with respect to tl[Pg.616]

If the function greater than or equal to the lowest energy Eg. Combining the latter two observations allows the energy expectation value of to be used to produce a very important inequality ... [Pg.57]

The partition function Z is given in the large-P limit, Z = limp co Zp, and expectation values of an observable are given as averages of corresponding estimators with the canonical measure in Eq. (19). The variables and R ( ) can be used as classical variables and classical Monte Carlo simulation techniques can be applied for the computation of averages. Note that if we formally put P = 1 in Eq. (19) we recover classical statistical mechanics, of course. [Pg.93]

Tlie expected value of a random variable is the average value of the random variable. The expected value of a random variable X is denoted by E(X). The expected value of a random variable can be interpreted as the long-run average of observations on tlie random variable. The procedure for calculating the expected value of a random variable depends on whether the random variable is discrete or continuous. [Pg.558]

Knowing the energy function H o) allows us, at least in principle, to calculate the average (or expected) value of any observable O of the system ... [Pg.327]

Its usefulness results primarily from the fact that the usual postulate ( mean value postulate, 1T) for the expectation value of an observable whose operator is P, or matrix P, namely p = atPa, may be replaced... [Pg.419]

The expression in Eq. (8-183) is the one used in practical calculations, the form shown in Eq. (8-184) demonstrates that the result is the expectation value of the observable R in the state 0>. [Pg.464]

Note that < 2 > is the expectation value of R in the state > and that w n) is the probability of this state in the ensemble. Therefore the expression Is exactly the ensemble average of the observ-... [Pg.467]

Bhawe (14) has simulated the periodic operation of a photo-chemically induced free-radical polymerization which has both monomer and solvent transfer steps and a recombination termination reaction. An increase of 50% in the value of Dp was observed over and above the expected value of 2.0. An interesting feature of this work is that when very short period oscillations were employed, virtually time-invariant products were predicted. [Pg.256]

The theory begins by constructing a target operator A that specifies the desired outcome of the experiment. In general,. 4 is a projection operator onto a set of observables. The objective is to maximize the target yield, or the expectation value of. 4 at a chosen target time, tf [23,24],... [Pg.252]

Since all energy-resolved observables can be inferred from appropriate expectation values of an energy-resolved wavefunction, Eq. (21) shows that the RWP method can be used to infer observables. Specific formulas for S matrix elements or reaction probabilities are given in Refs. [1] and [13]. See also Section IIIC below. [Pg.7]

Table 32.5 presents the expected values of the elements in the contingency Table 32.4. Note that the marginal sums in the two tables are the same. There are, however, large discrepancies between the observed and the expected values. Small discrepancies between the tabulated values of our illustrations and their exact values may arise due to rounding of intermediate results. Table 32.5 presents the expected values of the elements in the contingency Table 32.4. Note that the marginal sums in the two tables are the same. There are, however, large discrepancies between the observed and the expected values. Small discrepancies between the tabulated values of our illustrations and their exact values may arise due to rounding of intermediate results.
Nevertheless, the situation is not completely hopeless. There is a recipe for systematically approaching the wave function of the ground state P0> i- c., the state which delivers the lowest energy E0. This is the variational principle, which holds a very prominent place in all quantum-chemical applications. We recall from standard quantum mechanics that the expectation value of a particular observable represented by the appropriate operator O using any, possibly complex, wave function Etrial that is normalized according to equation (1-10) is given by... [Pg.23]

Equations (5.2)—(5.4) and Figs. 5.1-5.3 illustrate the nature of the structural observables obtained from gas-electron diffraction the intensity data provide intemuclear distances which are weighted averages of the expectation values of the individual vibrational molecular states. This presentation clearly illustrates that the temperature-dependent observable distribution averages are conceptually quite different from the singular, nonobservable and temperature independent equilibrium distances, usually denoted r -type distances, obtained from ab initio geometry optimizations. [Pg.137]

The difference between the expected demands d used in the model and the realization of an actual demand d causes a model mismatch. The scheduler corrects this error after the observation of this demand in a reactive manner by the production decisions taken at the beginning of the next period. For example, the decisions taken in period i = 1 take the expected value of the demand di into account (di = 6) while the decisions taken in period i = 2 take the true value into account. When di = 0, the large storage causes a relatively low production in the next period [xiih) = 5) whereas in case of d = 12, a deficit results and the production of the next period is larger (x2(t2) = 12). The sequence of decisions is a function of the observed demands and thus the sequence varies over the scenarios. [Pg.190]


See other pages where Expectation values of observables is mentioned: [Pg.13]    [Pg.252]    [Pg.254]    [Pg.13]    [Pg.147]    [Pg.21]    [Pg.21]    [Pg.169]    [Pg.252]    [Pg.254]    [Pg.127]    [Pg.13]    [Pg.252]    [Pg.254]    [Pg.13]    [Pg.147]    [Pg.21]    [Pg.21]    [Pg.169]    [Pg.252]    [Pg.254]    [Pg.127]    [Pg.454]    [Pg.167]    [Pg.244]    [Pg.256]    [Pg.32]    [Pg.43]    [Pg.47]    [Pg.442]    [Pg.101]    [Pg.300]    [Pg.294]    [Pg.193]    [Pg.194]    [Pg.273]    [Pg.81]    [Pg.194]   


SEARCH



Expectancies

Expectation value

Expectations

Expected

Observables, expectation values

Observation of

Observed Values of

Values observed

© 2024 chempedia.info