Heisenberg uncertainty derivation explanation
edit
Subtracting eqn.(32) from (31), we get
(
δ
k
2
(
x
2
−
x
1
)
=
π
)
⇒
(
x
2
−
x
1
=
2
π
δ
k
)
{\displaystyle \left({\frac {\delta k}{2}}(x_{2}-x_{1})=\pi \right)\Rightarrow \left(x_{2}-x_{1}={\frac {2\pi }{\delta k}}\right)}
Therefore, Error in the measurement of the particle:
Δ
x
=
2
π
Δ
k
=
2
π
Δ
(
2
π
λ
)
{\displaystyle \Delta x={\frac {2\pi }{\Delta k}}={\frac {2\pi }{\Delta \left({\frac {2\pi }{\lambda }}\right)}}}
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:
p
=
ℏ
Δ
k
=
h
2
π
2
π
λ
⇒
2
π
λ
=
2
π
p
h
{\displaystyle p=\hbar \Delta k={\frac {h}{2\pi }}{\frac {2\pi }{\lambda }}\Rightarrow {\frac {2\pi }{\lambda }}={\frac {2\pi p}{h}}}
Thus frome the above we have:
2
π
Δ
(
2
π
λ
)
=
2
π
Δ
(
2
π
p
h
)
=
1
Δ
(
p
h
)
{\displaystyle {\frac {2\pi }{\Delta \left({\frac {2\pi }{\lambda }}\right)}}={\frac {2\pi }{\Delta \left({\frac {2\pi p}{h}}\right)}}={\frac {1}{\Delta \left({\frac {p}{h}}\right)}}}
and because 'h' is a constant, we have:
1
Δ
(
p
h
)
=
Δ
h
Δ
p
=
h
Δ
p
{\displaystyle {\frac {1}{\Delta \left({\frac {p}{h}}\right)}}={\frac {\Delta h}{\Delta p}}={\frac {h}{\Delta p}}}
So now we have:
Δ
x
.
Δ
p
=
h
{\displaystyle \Delta x.\Delta p=h}
Note that this is true only if the probability distribution is normal. If it isn't
Δ
x
.
Δ
p
{\displaystyle \Delta x.\Delta p}
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:
Δ
x
.
Δ
p
≥
ℏ
[
=
h
2
π
]
{\displaystyle \Delta x.\Delta p\geq \hbar \left[={\frac {h}{2\pi }}\right]}
References:
Proof of Determinant to Dot and Cross Product
edit
M
=
(
a
1
a
2
a
3
b
1
b
2
b
3
c
1
c
2
c
3
)
=
(
A
→
B
→
C
→
)
{\displaystyle M={\begin{pmatrix}a_{1}&a_{2}&a_{3}\\b_{1}&b_{2}&b_{3}\\c_{1}&c_{2}&c_{3}\end{pmatrix}}={\begin{pmatrix}{\overrightarrow {A}}\\{\overrightarrow {B}}\\{\overrightarrow {C}}\end{pmatrix}}}
∥
M
∥=
|
a
1
a
2
a
3
b
1
b
2
b
3
c
1
c
2
c
3
|
{\displaystyle \parallel M\parallel ={\begin{vmatrix}a_{1}&a_{2}&a_{3}\\b_{1}&b_{2}&b_{3}\\c_{1}&c_{2}&c_{3}\end{vmatrix}}}
=
a
1
|
b
2
b
3
c
2
c
3
|
−
a
2
|
b
1
b
3
c
1
c
3
|
+
a
3
|
b
1
b
2
c
1
c
2
|
{\displaystyle =a_{1}{\begin{vmatrix}b_{2}&b_{3}\\c_{2}&c_{3}\end{vmatrix}}-a_{2}{\begin{vmatrix}b_{1}&b_{3}\\c_{1}&c_{3}\end{vmatrix}}+a_{3}{\begin{vmatrix}b_{1}&b_{2}\\c_{1}&c_{2}\end{vmatrix}}}
=
a
1
×
(
b
2
c
3
−
b
3
c
2
)
−
a
2
×
(
b
1
c
3
−
b
3
c
1
)
+
a
3
×
(
b
1
c
2
−
b
2
c
1
)
{\displaystyle =a_{1}\times (b_{2}c_{3}-b_{3}c_{2})-a_{2}\times (b_{1}c_{3}-b_{3}c_{1})+a_{3}\times (b_{1}c_{2}-b_{2}c_{1})}
=
a
1
×
(
b
2
c
3
−
b
3
c
2
)
+
a
2
×
(
b
3
c
1
−
b
1
c
3
)
+
a
3
×
(
b
1
c
2
−
b
2
c
1
)
{\displaystyle =a_{1}\times (b_{2}c_{3}-b_{3}c_{2})+a_{2}\times (b_{3}c_{1}-b_{1}c_{3})+a_{3}\times (b_{1}c_{2}-b_{2}c_{1})}
=
A
→
⋅
⟨
b
2
c
3
−
b
3
c
2
,
b
3
c
1
−
b
1
c
3
,
b
1
c
2
−
b
2
c
1
⟩
{\displaystyle ={\overrightarrow {A}}\cdot \langle b_{2}c_{3}-b_{3}c_{2},b_{3}c_{1}-b_{1}c_{3},b_{1}c_{2}-b_{2}c_{1}\rangle }
⟨
b
2
c
3
−
b
3
c
2
,
b
3
c
1
−
b
1
c
3
,
b
1
c
2
−
b
2
c
1
⟩
=
⟨
M
1
⟩
{\displaystyle \langle b_{2}c_{3}-b_{3}c_{2},b_{3}c_{1}-b_{1}c_{3},b_{1}c_{2}-b_{2}c_{1}\rangle =\langle M_{1}\rangle }
⇒∥
M
∥=
A
→
⋅
⟨
M
1
⟩
{\displaystyle \Rightarrow \parallel M\parallel ={\overrightarrow {A}}\cdot \langle M_{1}\rangle }
B
→
×
C
→
=
|
i
^
j
^
k
^
b
1
b
2
b
3
c
1
c
2
c
3
|
=
(
b
2
c
3
−
b
3
c
2
)
×
i
^
+
(
b
3
c
1
−
b
1
c
3
)
×
j
^
+
(
b
1
c
2
−
b
2
c
1
)
×
k
^
=
⟨
b
2
c
3
−
b
3
c
2
,
b
3
c
1
−
b
1
c
3
,
b
1
c
2
−
b
2
c
1
⟩
=
⟨
M
2
⟩
{\displaystyle {\overrightarrow {B}}\times {\overrightarrow {C}}={\begin{vmatrix}{\widehat {i}}&{\widehat {j}}&{\widehat {k}}\\b_{1}&b_{2}&b_{3}\\c_{1}&c_{2}&c_{3}\end{vmatrix}}=(b_{2}c_{3}-b_{3}c_{2})\times {\widehat {i}}+(b_{3}c_{1}-b_{1}c_{3})\times {\widehat {j}}+(b_{1}c_{2}-b_{2}c_{1})\times {\widehat {k}}=\langle b_{2}c_{3}-b_{3}c_{2},b_{3}c_{1}-b_{1}c_{3},b_{1}c_{2}-b_{2}c_{1}\rangle =\langle M_{2}\rangle }
⟨
M
1
⟩
=
⟨
M
2
⟩
⇒∥
M
∥=
A
→
⋅
⟨
M
2
⟩
{\displaystyle \langle M_{1}\rangle =\langle M_{2}\rangle \Rightarrow \parallel M\parallel ={\overrightarrow {A}}\cdot \langle M_{2}\rangle }
⇒∥
M
∥=
A
→
⋅
(
B
→
×
C
→
)
{\displaystyle \Rightarrow \parallel M\parallel ={\overrightarrow {A}}\cdot ({{\overrightarrow {B}}\times {\overrightarrow {C}}})}
Complete Fermi Dirac Integrals
edit
F
j
(
x
)
:=
(
1
/
Γ
(
j
+
1
)
)
∫
0
∞
d
t
(
t
j
/
(
exp
(
t
−
x
)
+
1
)
)
{\displaystyle F_{j}(x):=(1/\Gamma (j+1))\int _{0}^{\infty }dt(t^{j}/(\exp(t-x)+1))}
Incomplete Fermi Dirac Integrals
edit
F
j
(
x
,
b
)
:=
(
1
/
Γ
(
j
+
1
)
)
∫
b
∞
d
t
(
t
j
/
(
exp
(
t
−
x
)
+
1
)
)
{\displaystyle F_{j}(x,b):=(1/\Gamma (j+1))\int _{b}^{\infty }dt(t^{j}/(\exp(t-x)+1))}
Airy Functions and Derivatives
edit
A
i
(
x
)
=
(
1
/
π
)
∫
0
∞
cos
(
(
1
/
3
)
t
3
+
x
t
)
d
t
{\displaystyle Ai(x)=(1/\pi )\int _{0}^{\infty }\cos((1/3)t^{3}+xt)dt}
B
i
(
x
)
=
(
1
/
π
)
∫
0
∞
(
e
(
−
(
1
/
3
)
t
3
)
+
sin
(
(
1
/
3
)
t
3
+
x
t
)
)
d
t
{\displaystyle Bi(x)=(1/\pi )\int _{0}^{\infty }(e^{(}-(1/3)t^{3})+\sin((1/3)t^{3}+xt))dt}
C
l
2
(
x
)
=
−
∫
0
x
d
t
log
(
2
sin
(
t
/
2
)
)
{\displaystyle Cl_{2}(x)=-\int _{0}^{x}dt\log(2\sin(t/2))}
Normalized Hydrogenic Bound States
edit
R
n
:=
2
(
Z
3
/
2
/
n
2
)
(
n
−
l
−
1
)
!
/
(
n
+
l
)
!
exp
(
−
Z
r
/
n
)
(
2
Z
r
/
n
)
l
L
n
−
l
−
1
2
l
+
1
(
2
Z
r
/
n
)
{\displaystyle R_{n}:=2(Z^{3/2}/n^{2}){\sqrt {(n-l-1)!/(n+l)!}}\exp(-Zr/n)(2Zr/n)^{l}L_{n-l-1}^{2l+1}(2Zr/n)}
Where:
L
n
k
(
x
)
=
(
−
1
)
k
(
d
k
/
d
x
k
)
L
(
n
+
k
)
(
x
)
{\displaystyle L_{n}^{k}(x)=(-1)^{k}(d^{k}/dx^{k})L_{(}n+k)(x)}
F
(
ϕ
,
k
)
=
∫
0
ϕ
d
t
1
/
(
(
1
−
k
2
sin
2
(
t
)
)
)
{\displaystyle F(\phi ,k)=\int _{0}^{\phi }dt1/{\sqrt {(}}(1-k^{2}\sin ^{2}(t)))}
E
(
ϕ
,
k
)
=
∫
0
ϕ
d
t
(
(
1
−
k
2
sin
2
(
t
)
)
)
{\displaystyle E(\phi ,k)=\int _{0}^{\phi }dt{\sqrt {(}}(1-k^{2}\sin ^{2}(t)))}
Π
(
ϕ
,
k
,
n
)
=
∫
0
ϕ
d
t
1
/
(
(
1
+
n
sin
2
(
t
)
)
(
1
−
k
2
sin
2
(
t
)
)
{\displaystyle \Pi (\phi ,k,n)=\int _{0}^{\phi }dt1/((1+n\sin ^{2}(t)){\sqrt {(}}1-k^{2}\sin ^{2}(t))}
R
C
(
x
,
y
)
=
1
/
2
∫
0
∞
d
t
(
t
+
x
)
(
−
1
/
2
)
(
t
+
y
)
(
−
1
)
{\displaystyle RC(x,y)=1/2\int _{0}^{\infty }dt(t+x)^{(}-1/2)(t+y)^{(}-1)}
R
D
(
x
,
y
,
z
)
=
3
/
2
∫
0
∞
d
t
(
t
+
x
)
(
−
1
/
2
)
(
t
+
y
)
(
−
1
/
2
)
(
t
+
z
)
(
−
3
/
2
)
{\displaystyle RD(x,y,z)=3/2\int _{0}^{\infty }dt(t+x)^{(}-1/2)(t+y)^{(}-1/2)(t+z)^{(}-3/2)}
R
F
(
x
,
y
,
z
)
=
1
/
2
∫
0
∞
d
t
(
t
+
x
)
(
−
1
/
2
)
(
t
+
y
)
(
−
1
/
2
)
(
t
+
z
)
(
−
1
/
2
)
{\displaystyle RF(x,y,z)=1/2\int _{0}^{\infty }dt(t+x)^{(}-1/2)(t+y)^{(}-1/2)(t+z)^{(}-1/2)}
R
J
(
x
,
y
,
z
,
p
)
=
3
/
2
∫
0
∞
d
t
(
t
+
x
)
(
−
1
/
2
)
(
t
+
y
)
(
−
1
/
2
)
(
t
+
z
)
(
−
1
/
2
)
(
t
+
p
)
(
−
1
)
{\displaystyle RJ(x,y,z,p)=3/2\int _{0}^{\infty }dt(t+x)^{(}-1/2)(t+y)^{(}-1/2)(t+z)^{(}-1/2)(t+p)^{(}-1)}
Γ
(
x
)
=
∫
0
∞
d
t
t
x
−
1
exp
(
−
t
)
{\displaystyle \Gamma (x)=\int _{0}^{\infty }dtt^{x-1}\exp(-t)}
Psi (Digamma) Function
edit
ψ
(
n
)
(
x
)
=
(
d
/
d
x
)
n
ψ
(
x
)
=
(
d
/
d
x
)
n
+
1
log
(
Γ
(
x
)
)
⋅
{\displaystyle \psi ^{(n)}(x)=(d/dx)^{n}\psi (x)=(d/dx)^{n+1}\log(\Gamma (x))\cdot }
Riemann Zeta Function
edit
ζ
(
s
)
=
∑
k
=
1
∞
k
−
s
{\displaystyle \zeta (s)=\sum _{k=1}^{\infty }k^{-s}}
Hurwitz Zeta Function
edit
ζ
(
s
,
q
)
=
∑
0
∞
(
k
+
q
)
−
s
{\displaystyle \zeta (s,q)=\sum _{0}^{\infty }(k+q)^{-s}}
η
(
s
)
=
(
1
−
2
1
−
s
)
ζ
(
s
)
⋅
{\displaystyle \eta (s)=(1-2^{1-s})\zeta (s)\cdot }
J
(
n
,
x
)
:=
∫
0
x
d
t
t
n
e
t
/
(
e
t
−
1
)
2
{\displaystyle J(n,x):=\int _{0}^{x}dtt^{n}e^{t}/(e^{t}-1)^{2}}
Schwarzschild Radius Mass-Density relation
edit
ρ
=
3
c
6
32
π
G
3
m
2
⋅
{\displaystyle \rho ={\frac {3c^{6}}{32\pi G^{3}m^{2}}}\cdot \,\!}
The equations of motion are obtained by means of an action principle, written as:
δ
S
δ
φ
i
=
0
.
{\displaystyle {\frac {\delta {\mathcal {S}}}{\delta \varphi _{i}}}=0\,.}
where the action ,
S
{\displaystyle {\mathcal {S}}}
, is a functional of the dependent variables
φ
i
(
s
)
{\displaystyle \varphi _{i}(s)}
with their derivatives and s itself
S
[
φ
i
,
∂
φ
i
∂
s
]
=
∫
L
[
φ
i
[
s
]
,
∂
φ
i
[
s
]
∂
s
α
,
s
α
]
d
n
s
{\displaystyle {\mathcal {S}}\left[\varphi _{i},{\frac {\partial \varphi _{i}}{\partial s}}\right]=\int {{\mathcal {L}}\left[\varphi _{i}[s],{\frac {\partial \varphi _{i}[s]}{\partial s^{\alpha }}},s^{\alpha }\right]\,\mathrm {d} ^{n}s}}
and where
s
=
{
s
α
}
{\displaystyle s=\{s^{\alpha }\}\!}
denotes the set of n independent variables of the system, indexed by
α
=
1
,
2
,
3
,
…
,
n
.
{\displaystyle \alpha =1,2,3,\ldots ,n.}
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
d
d
t
∂
L
∂
φ
˙
i
=
∂
L
∂
φ
i
.
{\displaystyle {\frac {d}{dt}}{\frac {\partial {\mathcal {L}}}{\partial {\dot {\varphi }}_{i}}}={\frac {\partial {\mathcal {L}}}{\partial \varphi _{i}}}.}
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
d
L
=
∑
i
(
∂
L
∂
q
i
d
q
i
+
∂
L
∂
q
˙
i
d
q
˙
i
)
+
∂
L
∂
t
d
t
.
{\displaystyle \mathrm {d} {\mathcal {L}}=\sum _{i}\left({\frac {\partial {\mathcal {L}}}{\partial q_{i}}}\mathrm {d} q_{i}+{\frac {\partial {\mathcal {L}}}{\partial {{\dot {q}}_{i}}}}\mathrm {d} {{\dot {q}}_{i}}\right)+{\frac {\partial {\mathcal {L}}}{\partial t}}\mathrm {d} t\,.}
Now the generalized momenta were defined as
p
i
=
∂
L
∂
q
˙
i
{\displaystyle p_{i}={\frac {\partial {\mathcal {L}}}{\partial {{\dot {q}}_{i}}}}}
and Lagrange's equations tell us that
d
d
t
∂
L
∂
q
˙
i
−
∂
L
∂
q
i
=
0
{\displaystyle {\frac {\mathrm {d} }{\mathrm {d} t}}{\frac {\partial {\mathcal {L}}}{\partial {{\dot {q}}_{i}}}}-{\frac {\partial {\mathcal {L}}}{\partial q_{i}}}=0\,}
We can rearrange this to get
∂
L
∂
q
i
=
p
˙
i
{\displaystyle {\frac {\partial {\mathcal {L}}}{\partial q_{i}}}={\dot {p}}_{i}\,}
and substitute the result into the variation of the Lagrangian
d
L
=
∑
i
[
p
˙
i
d
q
i
+
p
i
d
q
˙
i
]
+
∂
L
∂
t
d
t
.
{\displaystyle \mathrm {d} {\mathcal {L}}=\sum _{i}\left[{\dot {p}}_{i}\mathrm {d} q_{i}+p_{i}\mathrm {d} {{\dot {q}}_{i}}\right]+{\frac {\partial {\mathcal {L}}}{\partial t}}\mathrm {d} t\,.}
We can rewrite this as
d
L
=
∑
i
[
p
˙
i
d
q
i
+
d
(
p
i
q
˙
i
)
−
q
˙
i
d
p
i
]
+
∂
L
∂
t
d
t
{\displaystyle \mathrm {d} {\mathcal {L}}=\sum _{i}\left[{\dot {p}}_{i}\mathrm {d} q_{i}+\mathrm {d} \left(p_{i}{{\dot {q}}_{i}}\right)-{{\dot {q}}_{i}}\mathrm {d} p_{i}\right]+{\frac {\partial {\mathcal {L}}}{\partial t}}\mathrm {d} t\,}
and rearrange again to get
d
(
∑
i
p
i
q
˙
i
−
L
)
=
∑
i
[
−
p
˙
i
d
q
i
+
q
˙
i
d
p
i
]
−
∂
L
∂
t
d
t
.
{\displaystyle \mathrm {d} \left(\sum _{i}p_{i}{{\dot {q}}_{i}}-{\mathcal {L}}\right)=\sum _{i}\left[-{\dot {p}}_{i}\mathrm {d} q_{i}+{{\dot {q}}_{i}}\mathrm {d} p_{i}\right]-{\frac {\partial {\mathcal {L}}}{\partial t}}\mathrm {d} t\,.}
The term on the left-hand side is just the Hamiltonian that we have defined before, so we find that
d
H
=
∑
i
[
−
p
˙
i
d
q
i
+
q
˙
i
d
p
i
]
−
∂
L
∂
t
d
t
=
∑
i
[
∂
H
∂
q
i
d
q
i
+
∂
H
∂
p
i
d
p
i
]
+
∂
H
∂
t
d
t
{\displaystyle \mathrm {d} {\mathcal {H}}=\sum _{i}\left[-{\dot {p}}_{i}\mathrm {d} q_{i}+{{\dot {q}}_{i}}\mathrm {d} p_{i}\right]-{\frac {\partial {\mathcal {L}}}{\partial t}}\mathrm {d} t=\sum _{i}\left[{\frac {\partial {\mathcal {H}}}{\partial q_{i}}}\mathrm {d} q_{i}+{\frac {\partial {\mathcal {H}}}{\partial p_{i}}}\mathrm {d} p_{i}\right]+{\frac {\partial {\mathcal {H}}}{\partial t}}\mathrm {d} t\,}
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
∂
H
∂
q
j
=
−
p
˙
j
,
∂
H
∂
p
j
=
q
˙
j
,
∂
H
∂
t
=
−
∂
L
∂
t
.
{\displaystyle {\frac {\partial {\mathcal {H}}}{\partial q_{j}}}=-{\dot {p}}_{j}\,,\quad {\frac {\partial {\mathcal {H}}}{\partial p_{j}}}={\dot {q}}_{j}\,,\quad {\frac {\partial {\mathcal {H}}}{\partial t}}=-{\partial {\mathcal {L}} \over \partial t}\,.}
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 .
Given a function
f
{\displaystyle f\,}
, the covariant derivative
∇
v
f
{\displaystyle \nabla _{\mathbf {v} }f}
coincides with the normal differentiation of a real function in the direction of the vector v , usually denoted by
v
f
{\displaystyle {\mathbf {v} }f}
and by
d
f
(
v
)
{\displaystyle df({\mathbf {v} })}
.
A covariant derivative
∇
{\displaystyle \nabla }
of a vector field
u
{\displaystyle {\mathbf {u} }}
in the direction of the vector
v
{\displaystyle {\mathbf {v} }}
denoted
∇
v
u
{\displaystyle \nabla _{\mathbf {v} }{\mathbf {u} }}
is defined by the following properties for any vector v , vector fields u, w and scalar functions f and g :
∇
v
u
{\displaystyle \nabla _{\mathbf {v} }{\mathbf {u} }}
is algebraically linear in
v
{\displaystyle {\mathbf {v} }}
so
∇
f
v
+
g
w
u
=
f
∇
v
u
+
g
∇
w
u
{\displaystyle \nabla _{f{\mathbf {v} }+g{\mathbf {w} }}{\mathbf {u} }=f\nabla _{\mathbf {v} }{\mathbf {u} }+g\nabla _{\mathbf {w} }{\mathbf {u} }}
∇
v
u
{\displaystyle \nabla _{\mathbf {v} }{\mathbf {u} }}
is additive in
u
{\displaystyle {\mathbf {u} }}
so
∇
v
(
u
+
w
)
=
∇
v
u
+
∇
v
w
{\displaystyle \nabla _{\mathbf {v} }({\mathbf {u} }+{\mathbf {w} })=\nabla _{\mathbf {v} }{\mathbf {u} }+\nabla _{\mathbf {v} }{\mathbf {w} }}
∇
v
u
{\displaystyle \nabla _{\mathbf {v} }{\mathbf {u} }}
obeys the product rule , i.e.
∇
v
f
u
=
f
∇
v
u
+
u
∇
v
f
{\displaystyle \nabla _{\mathbf {v} }f{\mathbf {u} }=f\nabla _{\mathbf {v} }{\mathbf {u} }+{\mathbf {u} }\nabla _{\mathbf {v} }f}
where
∇
v
f
{\displaystyle \nabla _{\mathbf {v} }f}
is defined above.
Note that
∇
v
u
{\displaystyle \nabla _{\mathbf {v} }{\mathbf {u} }}
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.
Given a field of covectors (or one-form )
α
{\displaystyle \alpha }
, its covariant derivative
∇
v
α
{\displaystyle \nabla _{\mathbf {v} }\alpha }
can be defined using the following identity which is satisfied for all vector fields u
∇
v
(
α
(
u
)
)
=
(
∇
v
α
)
(
u
)
+
α
(
∇
v
u
)
.
{\displaystyle \nabla _{\mathbf {v} }(\alpha ({\mathbf {u} }))=(\nabla _{\mathbf {v} }\alpha )({\mathbf {u} })+\alpha (\nabla _{\mathbf {v} }{\mathbf {u} }).}
The covariant derivative of a covector field along a vector field v is again a covector field.
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
φ
{\displaystyle \varphi }
and
ψ
{\displaystyle \psi \,}
are any two tensors:
∇
v
(
φ
⊗
ψ
)
=
(
∇
v
φ
)
⊗
ψ
+
φ
⊗
(
∇
v
ψ
)
,
{\displaystyle \nabla _{\mathbf {v} }(\varphi \otimes \psi )=(\nabla _{\mathbf {v} }\varphi )\otimes \psi +\varphi \otimes (\nabla _{\mathbf {v} }\psi ),}
and if
φ
{\displaystyle \varphi }
and
ψ
{\displaystyle \psi }
are tensor fields of the same tensor bundle then
∇
v
(
φ
+
ψ
)
=
∇
v
φ
+
∇
v
ψ
.
{\displaystyle \nabla _{\mathbf {v} }(\varphi +\psi )=\nabla _{\mathbf {v} }\varphi +\nabla _{\mathbf {v} }\psi .}
The covariant derivative of a tensor field along a vector field v is again a tensor field of the same type.
The Christoffel symbols can be derived from the vanishing of the covariant derivative of the metric tensor
g
i
k
{\displaystyle g_{ik}\ }
:
0
=
∇
ℓ
g
i
k
=
∂
g
i
k
∂
x
ℓ
−
g
m
k
Γ
m
i
ℓ
−
g
i
m
Γ
m
k
ℓ
.
{\displaystyle 0=\nabla _{\ell }g_{ik}={\frac {\partial g_{ik}}{\partial x^{\ell }}}-g_{mk}\Gamma ^{m}{}_{i\ell }-g_{im}\Gamma ^{m}{}_{k\ell }.\ }
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
0
=
g
i
k
;
ℓ
=
g
i
k
,
ℓ
−
g
m
k
Γ
m
i
ℓ
−
g
i
m
Γ
m
k
ℓ
.
{\displaystyle 0=\,g_{ik;\ell }=g_{ik,\ell }-g_{mk}\Gamma ^{m}{}_{i\ell }-g_{im}\Gamma ^{m}{}_{k\ell }.\ }
By permuting the indices, and resumming, one can solve explicitly for the Christoffel symbols as a function of the metric tensor:
Γ
i
k
ℓ
=
1
2
g
i
m
(
∂
g
m
k
∂
x
ℓ
+
∂
g
m
ℓ
∂
x
k
−
∂
g
k
ℓ
∂
x
m
)
=
1
2
g
i
m
(
g
m
k
,
ℓ
+
g
m
ℓ
,
k
−
g
k
ℓ
,
m
)
,
{\displaystyle \Gamma ^{i}{}_{k\ell }={\frac {1}{2}}g^{im}\left({\frac {\partial g_{mk}}{\partial x^{\ell }}}+{\frac {\partial g_{m\ell }}{\partial x^{k}}}-{\frac {\partial g_{k\ell }}{\partial x^{m}}}\right)={1 \over 2}g^{im}(g_{mk,\ell }+g_{m\ell ,k}-g_{k\ell ,m}),\ }
where the matrix
(
g
j
k
)
{\displaystyle (g^{jk}\ )}
is an inverse of the matrix
(
g
j
k
)
{\displaystyle (g_{jk}\ )}
, defined as (using the Kronecker delta , and Einstein notation for summation)
g
j
i
g
i
k
=
δ
j
k
{\displaystyle g^{ji}g_{ik}=\delta ^{j}{}_{k}\ }
.
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
Γ
i
k
ℓ
=
1
2
g
i
m
(
g
m
k
,
ℓ
+
g
m
ℓ
,
k
−
g
k
ℓ
,
m
+
c
m
k
ℓ
+
c
m
ℓ
k
+
c
k
ℓ
m
)
{\displaystyle \Gamma ^{i}{}_{k\ell }={\frac {1}{2}}g^{im}\left(g_{mk,\ell }+g_{m\ell ,k}-g_{k\ell ,m}+c_{mk\ell }+c_{m\ell k}+c_{k\ell m}\right)\ }
where
c
k
ℓ
m
=
g
m
p
c
k
ℓ
p
{\displaystyle c_{k\ell m}=g_{mp}{c_{k\ell }}^{p}\ }
are the commutation coefficients of the basis; that is,
[
e
k
,
e
ℓ
]
=
c
k
ℓ
m
e
m
{\displaystyle [e_{k},e_{\ell }]=c_{k\ell }{}^{m}e_{m}\,\ }
where e k are the basis vectors and
[
,
]
{\displaystyle [,]\ }
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.
--
α
κ
α
η
κ
ς
h
ν
α
ς
h
ι
ς
τ
h
{\displaystyle \alpha \kappa \alpha \eta \kappa \varsigma h\ \nu \alpha \varsigma h\iota \varsigma \tau h\,\!}
-- ακαηκςh ναςhιςτh