Heisenberg uncertainty derivation explanation

edit

Subtracting eqn.(32) from (31), we get

 

Therefore, Error in the measurement of the particle:

 

Now, based on deBroglie’s matter wave description, we can define a relatonship between the wavenumber [k] and the momentum [p] of the matter wave-particle:

 

Thus frome the above we have:

 

and because 'h' is a constant, we have:

 

So now we have:

 

Note that this is true only if the probability distribution is normal. If it isn't   will be greater, as the normal distribution turns out to have the minimum possible product. This means that we can write our more general physical law as:

 

References:

Proof of Determinant to Dot and Cross Product

edit

 
 
 

 

 

 


 


 


 


 


 

Special Functions

edit

Complete Fermi Dirac Integrals

edit

 

Incomplete Fermi Dirac Integrals

edit

 

Airy Functions and Derivatives

edit

 
 

Clausen Functions

edit

 

Normalized Hydrogenic Bound States

edit

 
Where: 

Legendre Forms

edit

 

 

 

Carlson Forms

edit

 

 

 

 

Gamma Functions

edit

 

Psi (Digamma) Function

edit

 

Riemann Zeta Function

edit

 

Hurwitz Zeta Function

edit

 

Eta Function

edit

 

Transport Functions

edit

 


Schwarzschild Radius Mass-Density relation

edit

 

MISC

edit

Lagrangian

edit

The equations of motion are obtained by means of an action principle, written as:

 

where the action,  , is a functional of the dependent variables   with their derivatives and s itself

 

and where   denotes the set of n independent variables of the system, indexed by  

The equations of motion obtained from this functional derivative are the Euler–Lagrange equations of this action. For example, in the classical mechanics of particles, the only independent variable is time, t. So the Euler-Lagrange equations are

 

Dynamical systems whose equations of motion are obtainable by means of an action principle on a suitably chosen Lagrangian are known as Lagrangian dynamical systems. Examples of Lagrangian dynamical systems range from the classical version of the Standard Model, to Newton's equations, to purely mathematical problems such as geodesic equations and Plateau's problem.

Deriving Hamilton's equations

edit

We can derive Hamilton's equations by looking at how the Lagrangian changes as you change the time and the positions and velocities of particles

 

Now the generalized momenta were defined as   and Lagrange's equations tell us that

 

We can rearrange this to get

 

and substitute the result into the variation of the Lagrangian

 

We can rewrite this as

 

and rearrange again to get

 

The term on the left-hand side is just the Hamiltonian that we have defined before, so we find that

 

where the second equality holds because of the definition of the partial derivatives. Associating terms from both sides of the equation above yields Hamilton's equations

 

Covariant Derivative

edit

A covariant derivative is a (Koszul) connection on the tangent bundle and other tensor bundles. Thus it has a certain behavior on functions, on vector fields, on the duals of vector fields (i.e., covector fields), and most generally of all, on arbitrary tensor fields.

Functions

edit

Given a function  , the covariant derivative   coincides with the normal differentiation of a real function in the direction of the vector v, usually denoted by   and by  .

Vector fields

edit

A covariant derivative   of a vector field   in the direction of the vector   denoted   is defined by the following properties for any vector v, vector fields u, w and scalar functions f and g:

  1.   is algebraically linear in   so  
  2.   is additive in   so  
  3.   obeys the product rule, i.e.   where   is defined above.

Note that   at point p depends on the value of v at p and on values of u in a neighbourhood of p because of the last property, the product rule.

Covector fields

edit

Given a field of covectors (or one-form)  , its covariant derivative   can be defined using the following identity which is satisfied for all vector fields u

 

The covariant derivative of a covector field along a vector field v is again a covector field.

Tensor fields

edit

Once the covariant derivative is defined for fields of vectors and covectors it can be defined for arbitrary tensor fields using the following identities where   and   are any two tensors:

 

and if   and   are tensor fields of the same tensor bundle then

 

The covariant derivative of a tensor field along a vector field v is again a tensor field of the same type.

Christoffel symbols

edit

The Christoffel symbols can be derived from the vanishing of the covariant derivative of the metric tensor  :

 

As a shorthand notation, the nabla symbol and the partial derivative symbols are frequently dropped, and instead a semi-colon and a comma are used to set off the index that is being used for the derivative. Thus, the above is sometimes written as

 

By permuting the indices, and resumming, one can solve explicitly for the Christoffel symbols as a function of the metric tensor:

 

where the matrix   is an inverse of the matrix  , defined as (using the Kronecker delta, and Einstein notation for summation)  . Although the Christoffel symbols are written in the same notation as tensors with index notation, they are not tensors. Indeed, they do not transform like tensors under a change of coordinates; see below.

NB. Note that most authors choose to define the Christoffel symbols in a holonomic basis, which is the convention followed here. In an anholonomic basis, the Christoffel symbols take the more complex form

 

where   are the commutation coefficients of the basis; that is,

 

where ek are the basis vectors and   is the Lie bracket. An example of an anholonomic basis with non-vanishing commutation coefficients are the unit vectors in spherical and cylindrical coordinates.

The expressions below are valid only in a holonomic basis, unless otherwise noted.


--  

-- ακαηκςh ναςhιςτh

Akanksh Vashisth

edit