Square root

The square root of a non-negative number {\displaystyle yis that (uniquely determined) non-negative number whose square is equal to the given number ysymbol for the square root is the root sign The square root {\sqrt {}}of the number yis thus {\sqrt {y}}represented by number or term yunder the root {\sqrt {y}}called a radicand. Less common is the more detailed notation {\sqrt[ {2}]{y}}.Furthermore, the square root can be expressed as a power: y^{{{\frac {1}{2}}}}is equivalent to {\sqrt {y}}.For example, because 3^{2}=3\cdot 3=9and 3\geq 0the square root of is equal to 93.

Since the equation x^{2}=yhas y>0solutions for , one usually defines the square root as the non-negative of the two solutions, i.e. it is always true that y ≥ {\sqrt {y}}\geq 0.This achieves that the notion of square root is unambiguous. The two solutions of the equation are thus x_{1}={\sqrt {y}}and x_{2}=-{\sqrt {y}}.

Zoom

Graph of the square root function y={\sqrt {x}}

In double logarithmic representation, the graph of the square root function becomes a straight line with slope 1⁄2 .   Zoom
In double logarithmic representation, the graph of the square root function becomes a straight line with slope 1⁄2 .  

Preliminary remark on the definitions

There are two problems to be considered in the formal definition of the square root:

  • If one restricts oneself to non-negative rational numbers, then the square root is not defined in many cases. Already in antiquity, it was found that the number {\sqrt {2}}for example, cannot be a rational number (see Euclid's proof of the irrationality of the root of 2).
  • In general, two different numbers exist whose squares agree with a given number. For example, because the number would (-3)^{2}=(-3)\cdot (-3)=9also be -3a possible candidate for the square root of 9.

The square root symbol was first used during the 16th century. It is assumed that the symbol is a modified form of the small r, which is an abbreviation for the Latin word "radix" (root). Originally, the symbol was placed in front of the radicand; the horizontal extension was missing. Carl Friedrich Gauss therefore still used brackets for more complicated root expressions and wrote, for example, instead of {\sqrt {}}(b^{2}-4ac){\sqrt {b^{2}-4ac}}.

In English, the square root is called "square root", which is why many programming languages use the term "sqrt" for the square root function.

Square roots of real numbers

Definition: The square root {\sqrt {y}}a non-negative real number yis that non-negative real number xwhose square is equal tox^{2}=x\cdot xy

Equivalently, the real square root can be defined as a function like this: Let

{\begin{aligned}q\colon [0;\infty {[}&\rightarrow [0;\infty {[}\\x&\mapsto y=x^{2}\end{aligned}}

the (bijective) restriction of the square function to the set of non-negative real numbers. The inverse function of this function qis called the square root function y\mapsto x={\sqrt {y}}.

Comments

  • Note that the square function explained by {\mathbb {R}}\rightarrow {\mathbb {R}};x\mapsto x^{2}is defined for all real numbers but is not invertible. It is neither injective nor surjective.
  • The constraint qof the square function is reversible and is inverted by the real root function. Since only non-negative real numbers qoccur as images of , the real root function is defined only for these numbers.
  • Due to the restriction of qto non-negative real numbers made before the inversion, the values of the square root function are non-negative numbers. The restriction of the square function to other subsets of \mathbb {R} in which different real numbers always have different squares would lead to other inverse functions, but these are not called real square root functions.

Examples

Square numbers and their square roots

Radikand

Square root

Radikand

Square root

1

1

121

11

4

2

144

12

9

3

169

13

16

4

196

14

25

5

225

15

36

6

256

16

49

7

289

17

64

8

324

18

81

9

361

19

100

10

400

20

The square root of a natural number is either an integer or irrational. The proof is analogous to Euclid's proof of the irrationality of the square root of 2.

Properties and calculation rules

The properties of the square root function result from the properties of the square function restricted to the set of non-negative real numbers:

  • {\displaystyle {\sqrt {a\cdot b}}={\sqrt {a}}\cdot {\sqrt {b}}\;}for {\displaystyle \;0\leq a,\,0\leq b}.
  • {\displaystyle {\sqrt {a\cdot b}}={\sqrt {-a}}\cdot {\sqrt {-b}}\;}for {\displaystyle \;a\leq 0,\,b\leq 0}.
  • 0\leq a<b\;\Longleftrightarrow \;0\leq {\sqrt {a}}<{\sqrt {b}}, i.e., the square root function is strictly monotonically increasing.
  • {\sqrt {a^{2}}}=|a|holds with the real magnitude for any real numbers a.
  • On the other hand, ({\sqrt {a}})^{2}=aonly valid for non-negative a.
  • The square root function is differentiable on \mathbb{R} _{+}differentiable, there {\frac {{\mathrm d}{\sqrt {x}}}{{\mathrm d}x}}={\frac {1}{2{\sqrt {x}}}}.
  • At the point 0 it is not differentiable, its graph has there a perpendicular tangent with the equation x=0.
  • It is Riemann integrable on any closed subinterval [a,b]its domain of definition, one of its primitive functions is F(x)={\tfrac {2}{3}}\cdot {\sqrt {x^{3}}}.
Diagram of the square function (red and blue). By mirroring only the blue half on the bisector of the 1st quadrant, the diagram of the square root function (green) is created.Zoom
Diagram of the square function (red and blue). By mirroring only the blue half on the bisector of the 1st quadrant, the diagram of the square root function (green) is created.

Calculation of square roots from real numbers

Rational approximate values of some
square roots

{\displaystyle {\begin{array}{ccccr}{\sqrt {2}}&\approx &{\sqrt {\frac {49}{25}}}&=&{\frac {7}{5}}\\{\sqrt {2}}&\approx &{\sqrt {\frac {289}{144}}}&=&{\frac {17}{12}}\\{\sqrt {2}}&\approx &{\sqrt {\frac {1681}{841}}}&=&{\frac {41}{29}}\\{\sqrt {3}}&\approx &{\sqrt {\frac {49}{16}}}&=&{\frac {7}{4}}\\{\sqrt {3}}&\approx &{\sqrt {\frac {361}{121}}}&=&{\frac {19}{11}}\\{\sqrt {5}}&\approx &{\sqrt {\frac {81}{16}}}&=&{\frac {9}{4}}\\{\sqrt {6}}&\approx &{\sqrt {\frac {2401}{400}}}&=&{\frac {49}{20}}\\{\sqrt {7}}&\approx &{\sqrt {\frac {64}{9}}}&=&{\frac {8}{3}}\\{\sqrt {8}}&\approx &{\sqrt {\frac {289}{36}}}&=&{\frac {17}{6}}\\{\sqrt {10}}&\approx &{\sqrt {\frac {361}{36}}}&=&{\frac {19}{6}}\\{\sqrt {11}}&\approx &{\sqrt {\frac {100}{9}}}&=&{\frac {10}{3}}\end{array}}}

Even then, if the square root is to be taken from a natural number, the result is often an irrational number, whose decimal fraction is thus a non-periodic, non-breaking decimal fraction (namely, precisely when the result is not natural). The calculation of a square root that is not a rational number thus consists in determining an approximate value of sufficient accuracy. There are a number of ways to do this:

Written root extraction

This is an algorithm similar to the common method of written division.

Interval nesting

This procedure is quite easy to understand, although very tedious in practical implementation.

Example (approximate value for {\sqrt {2}}):

From 1^{2}=1<2= 2^{2}=4>22 {\displaystyle lies {\sqrt {2}}and 2. Therefore, try 1{,}1^{2}, 1{,}2^{2}and so on. From 1{,}4^{2}=1{,}96<25 1{,}5^{2}=2{,}25>22 {\displaystyle must lie {\sqrt {2}}and 1,5. Continuing this procedure with more and more decimal places finally yields an approximate value with the desired accuracy:

1{,}41421356^{2}<2<1{,}41421357^{2}\;\Rightarrow \;{\sqrt {2}}\approx 1{,}41421356

Babylonian root extraction or Heron method

This iteration method is often used in programming the root calculation for calculators because it converges quickly. It is Newton's method for finding zeros applied to the function x\mapsto x^{2}-a.

Taylor series development

The Taylor series development of the root function {\displaystyle t\mapsto {\sqrt {t}}}with development place t=1can be written as Taylor development of {\displaystyle x\mapsto (1+x)^{1/2}}around the place x=0as binomial series

{\displaystyle \sum _{n=0}^{\infty }{\binom {1/2}{n}}\,x^{n}=\sum _{n=0}^{\infty }{\binom {2n}{n}}\,{\frac {(-1)^{n+1}}{(2n-1)\,4^{n}}}\,x^{n}=1+{\frac {1}{2}}x-{\frac {1}{8}}x^{2}+{\frac {1}{16}}x^{3}-{\frac {5}{128}}x^{4}\pm \dotsb }

can be found, because this series for converges |x|\leq 1point-wise towards {\displaystyle {\sqrt {1+x}}}With {\displaystyle x:=t-1}this results in

{\displaystyle {\sqrt {t}}=1+{\frac {1}{2}}(t-1)-{\frac {1}{8}}(t-1)^{2}+{\frac {1}{16}}(t-1)^{3}-{\frac {5}{128}}(t-1)^{4}\pm \dotsb }for {\displaystyle 0\leq t\leq 2.}

Calculation by means of CORDIC algorithm

This method is mainly used in arithmetic units, FPUs and microcontrollers.

Determining the square root graphically

One possibility is the cathetus theorem: The number nwhose square root is sought, is plotted on a number line from {\displaystyle 0}Over the distance between {\displaystyle 0}and na semicircle with radius r={\tfrac {n}{2}}drawn (Thales circle). At 1a perpendicular to the base line is established which intersects the semicircle (height of a right triangle). The distance of this intersection to the zero point is the square root of n(cathetus).

Square roots of complex numbers

If zis a non-zero complex number, the equation has

w^{2}=z

exactly two solutions for wwhich are also zcalled roots or square roots of These lie in the Gaussian number plane on the two intersections of the circle around 0 with the radius {\sqrt {|z|}}and the bisector of the angle between the {\displaystyle 0}rays starting from through 1respectivelyz. The one of the two roots that lies in the right half-plane is called the principal value of the root. For negative (real) zroot with a positive imaginary part is the principal value.

If one writes the complex number zin the form

{\displaystyle z=r\cdot {\rm {e^{\mathrm {i} \varphi },}}}

where φ \varphi and rare real with r>0- π -\pi <\varphi \leq \pi , then the following applies to the principal value of the root:

{\displaystyle w_{1}={\sqrt {r}}\cdot {\rm {e^{\mathrm {i} \varphi /2}}}}

The second root value (the secondary value) is obtained by point mirroring (180° rotation) at the zero point:

{\displaystyle w_{2}={\sqrt {r}}\cdot {\rm {e^{\mathrm {i} (\varphi /2+\pi )}}}}

Definition

The complex function "square z", q\colon {\mathbb {C}}\rightarrow {\mathbb {C}};z\mapsto z^{2}just like the real square function, has no inverse function because it is not injective, but unlike the real numbers it is surjective, that is, every complex number is the square of a complex number. One can therefore define complex square root functions analogously to the real (non-negative) square roots by restricting the domain of definition of qto a subset Dof the complex numbers on which qis injective and remains surjective. Depending on which subset one chooses for this, one obtains different branches of the square root function as an inverse.

The main branch of the complex square root function is obtained by taking as the domain of definition of q

{\displaystyle D_{H}:=\{x+\mathrm {i} \,y\in \mathbb {C} \mid x>0{\text{ oder }}(x=0{\text{ und }}y\geq 0)\}}

this is the right half-plane of the complex number plane, from whose edge only the numbers with non-negative imaginary part D_{H}belong to The restriction of qto D_{H}is a bijective mapping from D_{H}to the complex numbers, hence its inverse function, the main branch of the square root on all of {\displaystyle \mathbb {C} }defined. The value {\sqrt {z}}this inverse function is called the principal value of the square root of z. If by {\sqrt {z}}a certain complex number is meant, then it is this principal value.

If zgiven in Cartesian coordinates, i.e. {\displaystyle z=x+{\rm {iy}}}with real numbers xand ythen the result is

{\displaystyle {\sqrt {z}}={\sqrt {x+{\rm {iy}}}}={\sqrt {\tfrac {|z|+x}{2}}}+\mathrm {i} \cdot \operatorname {sgn^{+}} (y)\cdot {\sqrt {\tfrac {|z|-x}{2}}}}

for the main value of the square root, where the function the value -1 \operatorname {sgn^{+}}for negative yand otherwise (i.e. also for y=0and thus unlike the sign function \operatorname{sgn}) has the value 1:

{\text{sgn}}^{+}(y)={\begin{cases}+1&{\text{ für }}y\geq 0\\-1&{\text{ für }}y<0\end{cases}}

The only minor branch of qis -{\sqrt {z}}.

Given zin polar coordinates, z=|z|\cdot {\mathrm e}^{{{\mathrm i}\cdot \arg(z)}}with \arg(z)\in (-\pi ,\pi ], then the principal value of the square root is given by

{\sqrt {z}}={\sqrt {|z|}}{\mathrm e}^{{{\mathrm i}\cdot \arg(z)/2}}

where is {\sqrt {|z|}}the real (non-negative) square root of |z|The minor value is again obtained as -{\sqrt {z}}={\sqrt {|z|}}{\mathrm e}^{{{\mathrm i}\cdot (\arg(z)/2+\pi )}}.

The absolute value of the two roots is therefore the square root of the absolute value of the complex number. For the main value, the argument \arg(z)("the angle of z", see below) is halved. The other solution results geometrically from point mirroring of this principal value at the origin.

The argument of a complex number z=x+{\mathrm i}\,yis the oriented angle \angle (EOZ)in the complex number plane, the points are E(1|0),O(0|0)and Z(x|y)in real coordinates. In the picture of the following example, the argument of zand the argument of marked in colour. w_1

·         Complex square root

·        

A branch of the square root

·        

Second branch

·        

The Riemann area of the square root shows how the two branches merge.

Example: Calculation of a complex square root

We are looking for the square roots of z=-1+{\mathrm i}\,{\sqrt {3}}.First, the amount of the radicand is determined:

|z|=\left|-1+{\mathrm i}{\sqrt {3}}\right|={\sqrt {(-1)^{2}+({\sqrt {3}})^{2}}}={\sqrt {1+3}}={\sqrt {4}}=2

This gives the main value of the square root as

{\displaystyle {\begin{aligned}w_{1}&={\sqrt {\tfrac {2+(-1)}{2}}}+\mathrm {i} \cdot \operatorname {sgn^{+}} ({\sqrt {3}})\cdot {\sqrt {\tfrac {2-(-1)}{2}}}\\[0.3em]&={\sqrt {\tfrac {1}{2}}}+\mathrm {i} \cdot (+1)\cdot {\sqrt {\tfrac {3}{2}}}={\sqrt {2}}\cdot \left({\tfrac {1}{2}}+\mathrm {i} \cdot {\tfrac {1}{2}}{\sqrt {3}}\right)\end{aligned}}}

The other root is obtained by reversing the sign:

w_{2}=-w_{1}={\sqrt {2}}\cdot \left(-{\tfrac {1}{2}}-{\mathrm i}\cdot {\tfrac {1}{2}}{\sqrt {3}}\right)

Power law

The power law

{\displaystyle (a\cdot b)^{r}=a^{r}\cdot b^{r}\qquad \qquad \qquad \qquad \qquad \qquad \quad {\text{(P)}}}


does not hold for {\displaystyle r=1/2}for all {\displaystyle a,b\in \mathbb {C} }even for the principal values of the roots.
This can already be seen from the special case resulting from the
further specification {\displaystyle a=b=:z}

{\displaystyle {\sqrt {z^{2}}}=\left({\sqrt {z}}\right)^{2},}

which, because of the identity {\displaystyle \left({\sqrt {z}}\right)^{2}=z}becomes

{\displaystyle {\sqrt {z^{2}}}=z}

according to which every negative number obviously already provides a counterexample, for example {\displaystyle z=-1}:

Because {\displaystyle (-1)^{2}=1}and {\displaystyle \arg(1)=0}the principal value of {\displaystyle {\sqrt {(-1)^{2}}}}has the argument {\displaystyle \arg({\sqrt {1}})=0/2=0}while the principal value of -1{\displaystyle \arg(-1)=\pi }has the argument

Comments

  1. Since principal values of roots from positive radicands must be positive, the counterexample shows that there cannot be a square root function for which the power law {\displaystyle {\text{(P)}}}all {\displaystyle a,b\in \mathbb {C} }, cannot exist.
  2. For {\displaystyle r=1/2}and any one can{\displaystyle a,b\in \mathbb {C} } freely choose the "sign" of two of the three roots in {\displaystyle {\text{(P)}}}after which exactly one possibility remains for the "sign" of the last third.

Zoom

Root extraction corresponds to angle bisection in the complex plane. Example: {\displaystyle {\sqrt {\mathrm {i} }}}

Square roots modulo n

Also in the residue class ring {\mathbb {Z}}/n{\mathbb {Z}}can be defined as square roots. Quite analogously to the real and complex numbers, qis called a square root of xholds:

{\displaystyle q^{2}\equiv x{\bmod {n}}}

However, to calculate square roots modulo must nuse different methods than when calculating real or complex square roots. To determine the square roots of xmodulo n, one can proceed as follows:

First determine the prime factorisation

n=p_{1}^{{m_{1}}}\cdot p_{2}^{{m_{2}}}\dotsm p_{{k}}^{{m_{k}}}

of the modulus nand then the solutions modulo the individual prime powers p^{m}. Finally, these solutions are put together using the Chinese Remainder Theorem to find the solution.

Calculation of square roots modulo a prime number p

The case p=2is simple: Because {\displaystyle 0^{2}=0,\,1^{2}=1}and {\displaystyle 1\not \equiv 0{\bmod {2}}}modulo 2 every number has a uniquely determined square root, namely itself. For prime numbers not equal to p2, the calculation of the square roots of done xlike this:

To test whether xa square root in {\mathbb Z}/p{\mathbb Z}, calculate the value of the Legendre symbol

{\displaystyle \left({\frac {x}{p}}\right)\equiv x^{\frac {p-1}{2}}{\bmod {p}}},

because it is valid:

{\displaystyle \left({\frac {x}{p}}\right)={\begin{cases}-1,&{\text{wenn }}x{\text{ quadratischer Nichtrest modulo }}p{\text{ ist}}\\0,&{\text{wenn }}x{\text{ und }}p{\text{ nicht teilerfremd sind }}\\1,&{\text{wenn }}x{\text{ ein quadratischer Rest modulo }}p{\text{ ist}}\end{cases}}}

In the first case xhas no square root in {\mathbb Z}/p{\mathbb Z}and in the second case only the square root 0. So the interesting case is the third case and therefore we assume in the following that {\bigl (}{\tfrac {x}{p}}{\bigr )}=1holds.

Calculation for the case p mod 4 = 3

If the Legendre symbol {\bigl (}{\tfrac {x}{p}}{\bigr )}is equal to 1, then

{\displaystyle q\equiv \pm x^{\frac {p+1}{4}}{\bmod {p}}}

the two square roots of xmodulo p.

Calculation for the case p mod 4 = 1

If the Legendre symbol {\bigl (}{\tfrac {x}{p}}{\bigr )}is equal to 1, then

{\displaystyle q\equiv \pm {\frac {x}{2r}}\left(W_{\frac {p-1}{4}}+W_{\frac {p+3}{4}}\right){\bmod {p}}}

the two square roots of xmodulo p. Here one chooses rthat

{\displaystyle \left({\frac {r^{2}-4x}{p}}\right)=-1}

is valid. To do this, one can simply rtest different values of The sequence W_{n}is recursive through

W_{n}={\begin{cases}r^{2}/x-2,&{\text{ wenn }}n=1\\W_{{n/2}}^{2}-2,&{\text{ wenn }}n{\text{ gerade}}\\W_{{(n+1)/2}}W_{{(n-1)/2}}-W_{1},&{\text{ wenn }}n>1{\text{ ungerade}}\end{cases}}

defined.

Calculation example for x=3and p=37:

According to the above formula, the square roots of xare given by

{\displaystyle q\equiv \pm {\frac {x}{2r}}\left(W_{9}+W_{10}\right){\bmod {3}}7}

given. For rone finds by trial and error the value r = 2because it holds:

{\displaystyle \left({\frac {r^{2}-4x}{p}}\right)\equiv (r^{2}-4x)^{\frac {p-1}{2}}\equiv (-8)^{18}\equiv 36\equiv -1{\bmod {3}}7}

values for W_{9}and W_{{10}}result like this:

{\displaystyle {\begin{matrix}W_{1}&\equiv &r^{2}/x-2&\equiv &4/3-2&\equiv &24&{\bmod {3}}7\\W_{2}&\equiv &W_{1}^{2}-2&\equiv &24^{2}-2&\equiv &19&{\bmod {3}}7\\W_{3}&\equiv &W_{1}W_{2}-W_{1}&\equiv &24\cdot 19-24&\equiv &25&{\bmod {3}}7\\W_{4}&\equiv &W_{2}^{2}-2&\equiv &19^{2}-2&\equiv &26&{\bmod {3}}7\\W_{5}&\equiv &W_{2}W_{3}-W_{1}&\equiv &19\cdot 25-24&\equiv &7&{\bmod {3}}7\\W_{9}&\equiv &W_{4}W_{5}-W_{1}&\equiv &26\cdot 7-24&\equiv &10&{\bmod {3}}7\\W_{10}&\equiv &W_{5}^{2}-2&\equiv &7^{2}-2&\equiv &10&{\bmod {3}}7\\\end{matrix}}}

Inserting these values gives

{\displaystyle q\equiv \pm {\frac {x}{2r}}\left(W_{9}+W_{10}\right)\equiv \pm {\frac {3}{4}}(10+10)\equiv \pm 15{\bmod {3}}7.}

That is: 15 and 22 are the two square roots of 3 modulo 37.

Square roots of matrices

Main article: Square root of a matrix

The root of a square matrix Ais all matrices Bwhich, when multiplied by themselves, Agive

A=B\cdot B\Leftrightarrow B{\text{ ist Wurzel von }}A

As with the root of real or complex numbers, the root of matrices is not necessarily unique. However, if we consider only positive definite symmetrical matrices, the root formation is unambiguous: every positive definite symmetrical matrix Ahas a unambiguous positive definite symmetrical root A^{{\frac {1}{2}}}.It is obtained by diagonalising Ausing an orthogonal matrix (this is always possible according to the spectral theorem) and then replacing the diagonal elements with their roots; however, the positive root must always be chosen. See also Cholesky decomposition. The uniqueness follows from the fact that the exponential mapping is a diffeomorphism from the vector space of symmetric matrices onto the subset of positive definite symmetric matrices.

Square root of an approximated integral operator

One can take the definite integral function {\displaystyle G,\,g_{i}:=g(x_{i})}from 0 to x_{i}with x_{i}=i\Delta xand {\displaystyle i=0,1,\dotsc ,n-1}a given function which {\displaystyle F,\,f_{i}:=f(x_{i})}f_{i}assumes the values at the equidistant grid points x_{i}as a matrix multiplication G=FIas follows (for n=4):

{\displaystyle G=FI={\begin{pmatrix}g_{0}&g_{1}&g_{2}&g_{3}\\0&g_{0}&g_{1}&g_{2}\\0&0&g_{0}&g_{1}\\0&0&0&g_{0}\end{pmatrix}}={\begin{pmatrix}f_{0}&f_{1}&f_{2}&f_{3}\\0&f_{0}&f_{1}&f_{2}\\0&0&f_{0}&f_{1}\\0&0&0&f_{0}\end{pmatrix}}{\begin{pmatrix}\Delta x&\Delta x&\Delta x&\Delta x\\0&\Delta x&\Delta x&\Delta x\\0&0&\Delta x&\Delta x\\0&0&0&\Delta x\end{pmatrix}}}

It is vividly clear that one can repeat this operation and thus {\displaystyle H,\,h_{i}:=h(x_{i})}obtain the double integral

H=GI=FII=FI^{2}

Thus, the matrix Ican be understood as a numerically approximated integral operator.

The matrix Iis not diagonalisable and its Jordanian normal form is:

{\begin{pmatrix}\Delta x&1&0&0\\0&\Delta x&1&0\\0&0&\Delta x&1\\0&0&0&\Delta x\end{pmatrix}}

To draw a square root from this, one could proceed as described for the non-diagonalisable matrices. However, there is a more direct formal solution in this case as follows:

{\displaystyle I^{\beta }={\begin{pmatrix}\alpha _{0}&\alpha _{1}&\alpha _{2}&\alpha _{3}\\0&\alpha _{0}&\alpha _{1}&\alpha _{2}\\0&0&\alpha _{0}&\alpha _{1}\\0&0&0&\alpha _{0}\end{pmatrix}}}

with α \alpha _{0}=(\Delta x)^{\beta }, \alpha _{k}=\sum _{{j=1}}^{{k}}{\frac {\Gamma (\beta +1)(-1)^{{j+1}}\alpha _{{k-j}}}{\Gamma (j+1)\Gamma (\beta -j+1)}}and {\displaystyle k=1,2,\dotsc ,n-1}.

In it, the indices of α \alpha denote the subdiagonals (0 is the diagonal) and the exponent β \beta is equal to {\tfrac {1}{2}}. If Δ \Delta xassumed to be real and positive, then (\Delta x)^{{\frac {1}{2}}}real and positive by definition.

Thus, one can numerically approximate a "half" definite integral {\displaystyle L,\,l_{i}:=l(x_{i})}from 0 to x_{i}the function f(x)as follows:

{\displaystyle L=FI^{\beta }={\begin{pmatrix}l_{0}&l_{1}&l_{2}&l_{3}\\0&l_{0}&l_{1}&l_{2}\\0&0&l_{0}&l_{1}\\0&0&0&l_{0}\end{pmatrix}}={\begin{pmatrix}f_{0}&f_{1}&f_{2}&f_{3}\\0&f_{0}&f_{1}&f_{2}\\0&0&f_{0}&f_{1}\\0&0&0&f_{0}\end{pmatrix}}{\begin{pmatrix}\alpha _{0}&\alpha _{1}&\alpha _{2}&\alpha _{3}\\0&\alpha _{0}&\alpha _{1}&\alpha _{2}\\0&0&\alpha _{0}&\alpha _{1}\\0&0&0&\alpha _{0}\end{pmatrix}}}

If one looks for all operators which, multiplied by themselves, give the approximated integral operator I, one must also insert the negative sign, that is, there are two solutions \pm I^{{\frac {1}{2}}}.

To derive the formula, one can first invert I, \beta exponentiate the result with β and finally invert again.

See also

  • Root of 2, Euclid's proof of irrationality of root 2
  • Root from 3
  • Root (Mathematics)
  • Modulo, residual class ring
  • Penrose's square root law

AlegsaOnline.com - 2020 / 2023 - License CC3