Numeric integration with julia

Let \(f(x)\) be some non-negative, continuous function over the interval \([a,b]\). The area under the graph of \(f(x)\) is given by the definite integral:

\[ \text{Area under f} = \int_a^b f(x) dx \]

Computing this area is often made easier with the Fundamental Theorem of Calculus which states in one form that one can compute a definite integral through knowledge of an antiderivative. In particular, if \(F(x)\) is an antiderivative for \(f(x)\), a continuous function, then

\[ \int_a^b f(x) dx = F(b) - F(a). \]

This is great as long as some antiderivative is known. However, it is a fact of life that not all nice functions will have an antiderivative in a convenient form. In these cases, the above approach is of no help. Rather, to find the area, one can turn to approximations that progressively get better as more approximations are taken.

One such approximation is given by the familiar Riemann sums, which we will look at here. However, the problem of trying to find the area of geometric figures did not start with Riemann some 150 years ago, indeed it has a much longer history. In the time of Pythagorus the idea of calculating area was one of being able to construct a square of equal area to a figure. This was known as quadrature. Finding such answers for figures bounded by curves was difficult, though Archimedes effectively computed this area under \(f(x) = x^2\) about 2,000 years before Riemann sums using triangles, not rectangles to approximate the area. By medieval Europe, the term quadrature evolved to be the computation of an area by any means. For example, Galileo and Roberval found the area bounded by a cycloid arch. Along the way, other approximations were used. We will see those due to Simpson and Gauss, both predating Riemann.

Riemann sums

In 1854 Riemann was the first to give a rigorous definition of the integral of a continuous function on a closed interval, the problem we wish to solve here, using the concept of a Riemann sum. A Riemann sum is one of the simplest to understand approximations to the area under a curve. The basic idea is that the interval \([a,b]\) is partitioned through points \(a = x_0 < x_1 < \cdots x_n = b\) and the area under \(f(x)\) between \(x_i\) and \(x_{i+1}\) is approximated by a rectangle with the base \(x_{i+1} - x_i\) and height given by \(f(x_i^*)\), where \(x_i^*\) is some point in the interval \([x_i, x_{i+1}]\). Typical choices are the left point or the right point of the interval, or the \(x\) value which minizes or maximizes \(f\) over the interval. The figure shows these four choices for some sample function. For a Riemann integrable function, such as a continuous function on \([a,b]\), any of the choices will yield the same value as the partition's mesh shrinks to \(0\).

Embedded Image

As with other limits, we can numerically approximate the limit by computing the Riemann sum for some partition. The steps for this include:

  • creating a partition of \([a,b]\). We will use evenly spaced points for convenience.

  • Selecting the \(x_i^*\) within the partition

  • Computing the values \(f(x_i^*)(x_{i+1} - x_i)\) for each \(i\)

  • adding these values up

If we partition \([a,b]\) into \(n\) same sized intervals, then each has length \(\delta = (b-a)/n\) and so the points are separated by this amount. As such, we can choose our \(a = x_0 < x_1 < \dots < x_n = b\) with commands like:

a, b, n = 1, 3, 5;
delta = (b-a)/n;
x = a + (0:n) * delta           # alternatively linspace(a, b, n+1)
## 1.0:0.4:3.0

To apply a function to a range of values, we may use a map, a comprehension, a for loop or the "dot" notation. We will use map here. Recall, the syntax:

f(x) = x^2
fx = map(f, x)
## [1.0,1.9599999999999997,3.24,4.840000000000001,6.760000000000001,9.0]

Now to add the numbers up. For this task, the sum function is available

sum(fx)
## 26.8

Okay, just one subtlety, we really only want the points

[ a + (0:n-1) * delta ]'
## 1x5 Array{Float64,2}:
##  1.0  1.4  1.8  2.2  2.6

for the left Riemann sum and the points

[ a + (1:n) * delta ]
## [1.4,1.7999999999999998,2.2,2.6,3.0]

for the right.

Putting this together, here are commands to approximate the area under the curve \(f(x)=x^2\) using 10 left Riemann sums:

f(x) = x^2
a, b, n = 1, 3, 5;         ## note n=5
delta = (b - a)/n;         ## nothing to change below here
xs = a + (0:n-1) * delta;          ## n, right is 1:n * delta 
fx = map(f, xs);               
sum(fx) * delta
## 7.120000000000001

We compare this value to the known value from the Fundamental Theorem of Calculus, as \(F(x) = x^3/3\) is an antiderivative:

F(x) = x^3/3
F(b) - F(a)
## 8.666666666666666

Boy, not too close. We need a better approximation of course, which means simply that we need n to be bigger.

Practice

Question

Repeat with n=100

For the same problem, let \(n=100\). What do you get?

Question

Repeat with n=1,000

For the same problem, let \(n=1000\). What do you get?

Question

Repeat with n=10,000

For the same problem, let \(n=10,000\). is the difference between the answer and the actual answer within \(0.001\)?

Question

How big should n be? (Strang)

Let \(f(x) = (10 + \cos(2\pi x))^{-1}\). For the integral over \([0,1]\), the known answer is \(1/\sqrt{99}\). How big must \(n\) be so that the error in the Riemann sum is less than \(10^{-8}\)?

Integrate function

Here we write a function to do the integration. This needs the basic inputs of

  • a function,
  • the interval's start and end value, and
  • the number of equal-sized subintervals.

In addition, we allow for the possibility of passing in a function to compute the approximate area for a given subinterval. We give a default value where the left-hand endpoint is chosen.

You can type or copy and paste these two function definitions in:

function integrate(f::Function, a::Real, b::Real, n::Integer, approx_area::Function)
   delta = (b-a)/n
   x = a + (0:n) * delta
   sum( [ approx_area(f, x[i], x[i+1]) for i in 1:n ] )
end

We will use the left endpoint for the default choice of point in each subinterval:

function integrate(f::Function, a::Real, b::Real, n::Integer) 
     left_riemann(f::Function, a, b) = f(a)*(b-a)
     integrate(f, a, b, n, left_riemann)
end

The integrate function in action

The basic usage of the integrate function is straightforward. Here we approximate the integral of \(e^{-x^2}\) from \(0\) to \(3\) using \(10,000\) subintervals:

f(x) =  exp(-x^2)       
integrate(f, 0, 3, 10_000)
## 0.8863573297424971

How big should the number of intervals be? More intervals will give better answers, but unlike Newton's method we have no stopping criteria. For this problem, we look at various values based on n:

[integrate(f, 0, 3, n) for n in [100, 1000, 10000, 100000]]   ## or use 10.^(2:5)
## {0.9012054416030275,0.8877071625894734,0.8863573297424971,0.8862223464083187}

We see a value around \(0.886\) as the answer.

Continuity

For some integrals, you may need to make a minor adjustment for lack of continuity. The function \(f(x) = \sin(x)/x\) over the interval \([0, \pi]\) has to be defined to be \(1\) at \(0\) to be continuous. We do so here:

f(x) = x > 0 ? sin(x)/x : 1

Then integrate may be used as before, this time with \(50,000\) subintervals:

integrate(f, 0, pi, 50_000)
## 1.8519684678042823

Had we simply specified f(x) = sin(x)/x, then julia would have returned NaN for x=0 which have led to the entire integral being computed as NaN:

f(x) = sin(x)/x
integrate(f, 0, pi, 50_000)
## NaN

left versus right

If we define right_riemann as

right_riemann(f, a, b) = f(b) * (b-a)

Then we can compare the right and left Riemann sums. Let's do so for the monotonic function \(e^x\) over the interval \([0,2]\).

f(x) = exp(x)
ns = 10.^(2:5);
ys = [integrate(f, 0, 2, n, right_riemann) - integrate(f, 0, 2, n) for n in ns];
[ns ys]
## 4x2 Array{Any,2}:
##     100  0.127781   
##    1000  0.0127781  
##   10000  0.00127781 
##  100000  0.000127781

Since these are also the minimum and maximum Riemann sums, the above gives a bound on the error in the approximations. We can see it converges quite slowly, in that there are quite a few computations needed to get even a modest bound. (\(100,000\) for \(0.00013\)).

Practice

Question

Define the midpoint area by

midpoint(f, a, b) = f((a+b)/2) * (b-a)

With \(n=10,000\) what does integrate return when \(f(x) = \sin^2(x)\) between \(0\) and \(\pi\)?

Question

Repeat the above analysis comparing the right and left Riemann sums, but this time multiply by \(n\), as follows:

f(x) = exp(x)
[n * (integrate(f, 0, 2, n, right_riemann) - integrate(f, 0, 2, n)) for n in 10.^(2:5)]
## {12.778112197861269,12.778112197860736,12.77811219787317,12.778112197864289}

This shows what?

Improvements to the rectangle method

The basic left or right Riemann sum will converge, but the convergence is really slow. The value of using rectangles over a grid to approximate area is for theoretical computations, for numeric computations better approximations were known well before Riemann. We mention a few:

  • The trapezoid rule and Simpson's rule approximate the area under the curve better, as instead of a rectangle they use a trapezoid (linear fit between two points) or a quadratic fit between the two points.)

  • Gauss quadrature uses non-evenly selected points within the range and a weighting which is exact for polynomials of a given degree.

  • Adaptive methods pick a non-uniform set of points to use based on where a function is less well behaved.

Trapezoid and Simpson's rule

The trapezoid rule simply replaces the approximation of the area in a subinterval by a trapezoid, as opposed to a rectangle. Easy enough, digging up the formula from geometry for the area of a trapezoid, we can write our approximation function with:

trapezoid(f, a, b) = (1/2) * (f(a) + f(b)) * (b-a)

We can use this as follows. Let's approximate the area under \(5x^4\) curve between \(0\) and \(1\) (with known answer \(1\)):

f(x) = 5x^4
integrate(f, 0, 1, 1000, trapezoid)
## 1.0000016666665

Pretty close to 1 with just 1,000 subintervals.

We now compare the error with the left Riemann sum for the same size \(n\):

ns = 10 .^ (2:5);           
left_r = [integrate(f, 0, 1, n) for n in ns];
trapezoid_r = [integrate(f, 0, 1, n, trapezoid) for n in ns];
[ns 1-left_r 1-trapezoid_r]
## 4x3 Array{Any,2}:
##     100  0.0248333    -0.000166665
##    1000  0.00249833   -1.66667e-6 
##   10000  0.000249983  -1.66667e-8 
##  100000  2.49998e-5   -1.66667e-10

One can see that the errors are much smaller for the trapezoid method.

Simpson's rule

The trapezoid rule can be viewed as a simple linear approximation to the function \(f(x)\) over the subinterval \([a, b]\). That is, replace the function with the secant line between these two values and integrate the replacement. With this viewpoint, it is possible that other easy-to-integrate function approximations will lead to improved approximate integrals. Simpson's method can be viewed in just this way. It replaces \(f\) by the parabola going through \((a, f(a))\), \((c, f( c))\) and \((b, f(b))\) where \(c=(a+b)/2\) is the midpoint between \(a\) and \(b\). The resulting area after this approximation is:

simpsons(f, a, b) = (f(a) + 4f((a+b)/2) + f(b))*(b-a)/6

We compare how accurate we get with this rule for the same f as before:

simpsons_r = [integrate(f, 0, 1, n, simpsons) for n in ns];
[ns 1-left_r 1-trapezoid_r 1-simpsons_r]
## 4x4 Array{Any,2}:
##     100  0.0248333    -0.000166665  -4.16667e-10
##    1000  0.00249833   -1.66667e-6   -4.17444e-14
##   10000  0.000249983  -1.66667e-8    0.0        
##  100000  2.49998e-5   -1.66667e-10   0.0        

As can be seen, for this function approximating with a parabola is much quicker to converge. That is, \(n\) can be smaller yet the same accuracy is maintained. (Of course, there are more computations involved for each, so the number of operations needed may or may not be fewer, that would require some analysis.)

Error

It can be shown that the error for Simpson's method is bounded by

\[ \frac{1}{90}\frac{1}{2^5} M (b-a)^5 \frac{1}{n^4}, \]

where \(M\) is a bound on the fourth derivative. As we increase \(n\), the error gets small at a quick rate. By contrast, the error for the trapezoid method will be like \(n^{-2}\) and the left Riemann sum like \(n^{-1}\).

Practice

Question

The trapezoid rule has no error for linear functions and Simpson's rule has no error for quadratic functions. Verify the latter by computing the following:

f(x) = x^2; F(x) = x^3/3
integrate(f, 0, 10, 100, simpsons) - (F(10) - F(0));

How accurate is the approximation? Around

Question

Compare the difference between the trapezoid rule and Simpson's rule when integrating \(\cos(x)\) from \(0\) to \(\pi/6\). How big is the difference when \(n=1000\)?

integrate(cos, 0, pi/6, 1000, trapezoid) - integrate(cos, 0, pi/6, 1000, simpsons);

How big is the difference?

Question ....

Using Simpson's rule and n= 3800 compute the integral of \(f(x) = 1/(1+x^2)\) between \(0\) and \(1\).

The quadgk function

Julia provides the quadgk function to do adaptive Gauss-Konrod quadrature, a modern, fast and accurate means to compute 1-dimensional integrals numerically. The use is straightforward, and similar to integrate above: you specify a function object, and the limits of integration. You don't specify \(n\) -- as this is computed adaptively -- but you can optionally specify a tolerance which controls the accuracy, though we don't do so here. For example, as typical usage might be:

quadgk(sin, 0, pi)      ## 2 is exact
## (2.0000000000000004,1.7896795156957523e-12)

Two values are returned, the answer and an estimate of the error. In the above, \(2\) is the exact answer to this integral, the estimated value a just a bit more \(2\), but is guaranteed to be off my no more than the second value, \(1.78 \cdot 10^{-12}\).

The known answer here is \(1/3\), and quadgk gets it right for all the digits:

quadgk(x -> x^2, 0, 1)
## (0.3333333333333333,5.551115123125783e-17)

For other integration routines, the Cubature package is an interface to the Cubature library (http://ab-initio.mit.edu/wiki/index.php/Cubature) which provides serveral. Cubature is the term for higher dimensional integrals, quadrature refers to finding area. In that package, the function hquadrature is similar to quadgk. (The two are written by the same author.)

Practice

Question Use quadgk

Let \(f(x) = \exp(-4 \cdot |x-1/2|)\). Find the integral over \([0,1]\) using quadgk:

Question Use quadgk

Let \(f(x) = \sin(100\pi x)/(\pi x)\). Find the integral over \([0,1]\) using quadgk:

Question Use quadgk

Let \(f(x) = \sin(100\pi x)/(100\pi x)\). Using \(1,000\) points, find the Riemann integral with right hand endpoints

begin
  f(x) = sin(100*pi*x)/(100*pi*x)
  right_riemann(f::Function, a, b) = f(b) * (b-a)
  val = integrate(f, 0, 1, 1000, right_riemann);
end
## 0.0044899515576561425

What is it?

(The answer via Riemann sums isn't even correct to 4 decimal points, due to the highly oscillatory nature of the function.)

Question

How far off is this?

begin
  f(x) = 1/(1 + x^4)
  quadgk(f, 0, 1)[1] - integrate(f, 0, 1, 100000);
end
## -2.499991666571333e-6

Question

The quadgk function allows you to specify issues where there are troubles. For example, we know that \(f(x) = \sin(x)/x\) has an issue at 0. Directly trying this integral quadgk(x->sin(x)/x, -pi, pi) will fail, but you can specify the issue at \(0\) as follows quadgk(x -> sin(x)/x, -pi, 0, pi). Do so. What is the value of the result:

Question

Let \(f(x) = |x - 0.3|^{-1/4}\). We wish to find \(\int_0^1 f(x) dx\). The problem with this function is the singularity at \(x=0.3\). (That is, the function is not continuous, so has no guarantee that an integral over a closed domain exists.) However, some such integrals do exist, and the quadgk function can integrate around such singularities by spelling them out in the domain of integration. Just specify the trouble spots between the endpoints:

f(x) = abs(x - 0.3)^(-1/4)
val = quadgk(f, 0, 0.3, 1);

Following the above, what answer do you get?

Applications

There are many more applications of the integral beyond computing areas under the curve. Here we discuss two:

  • finding the volume of a figure with rotational symmetry (a glass in our example) and
  • finding the arc length of a line.

In each case one integrates a function related to the one describing the problem. If you keep this straight, the applications are no different than above.

Volume as a function of radius.

For a symmetrical drinking vessel, like most every glass you drink from, the Volume can be computed from a formula if a function describing the radius is known. For a given glass, let \(r(h)\) give the radius as a function of height. Then the volume of the vessel as a function of height, \(b\), is given by an integral:

\[ V(b) = \int_0^b \pi (r(h))^2 dh \]

We wish to look at our intuition relating the height of the fluid in the vessel compared to the percentage of fluid of the whole. A basic question might be: If the vessel is filled half way by height, is the volume half of the total, more or less?

The answer, of course, depends on the shape of the glass. That is the shape of the function \(r(h)\). Note, if \(r(h)\) is a constant -- the glass is a cylinder -- then the half-height mark is also the half-volume mark. Not so in general.

In Glass Shape Influences Consumption Rate for Alcoholic Beverages the authors demonstrate that the shape of the glass can have an effect on the rate of consumption, presumably people drink faster when they aren't sure how much they have left. In particular, they comment that people have difficulty judging the half finished by volume mark.

We work with metric units, as there is a natural relation between volume in cm\(^3\) and liquid measure (1 liter = 1000 cm\(^3\), so a 16-oz pint glass is roughly \(450\) cm\(^3\).)

Let two glasses be given as follows. A typical pint glass with linearly increasing radius:

\[ r(h) = 3 + \frac{1}{5}h, \quad 0 \leq h \leq b; \]

and a curved edge one:

\[ s(h) = 3 + \log(1 + h), \quad 0 \leq h \leq b \]

Embedded Image

One could also consider a fluted one, such as appears in the comparison noted in the article.

Question

Which of these functions might describe a fluted glass where the radius changes faster as the height gets bigger, that is the radius is a concave up function?


For the two types of glasses in the figure, we create functions in julia as follows:

r(h) = 3 + h/5
s(h) = 2 + log(1 + h)
r_vol(b) = quadgk(x -> pi*r(x)^2, 0, b)[1]
s_vol(b) = quadgk(x -> pi*s(x)^2, 0, b)[1]

Then we can easily find the volume as a function of height. For example at 10cm we have:

(r_vol(10), s_vol(10))
## (513.1268000863329,427.26481657392833)

However, to find \(b\) that makes the glass \(450\) cm\(^3\) requires us to solve an equation involving an integral for \(b\):

\[ V(b) = \int_0^b \pi r(h)^2 dh = 450. \]

Not to worry, we can use fzero from the Roots package for that. To solve for when V(b) = r_vol(b) - 450 = 0 we have

using Roots
r_b = fzero(x -> r_vol(x) - 450,  10)
## 9.168923214523723

So \(b\) is basically \(9.17\). Given this, how much volume is left at b/2?

r_vol(r_b/2)
## 173.27527610482952

That is about j r_vol(r_b/2) / r_vol(r_b) *100 percent (\(\approx 173.28/450 \cdot 100\)). As this height is often mistaken for the half-way by volume mark, people tend to drink these pints faster than they think.

Now compare to the height to get half the volume (225 ml):

fzero(x -> r_vol(x) - 225,  5)
## 5.603662378152274

Or about \(5.6038\). That is, when you are at j x/r_b*100 percent \(\approx 5.6038/9.169 \cdot 100\) of the height you have only half the volume remaining (and not at 50% of the height.)

Practice

Question

Compare the above for the curved glass, where \(s(h) = 3 + \log(1 + h)\).

What is the height of the glass, b, needed to make the volume 450?

Question

Find the volume of the glass represented by \(s(h) = 3 + \log(1 + h), 0 \leq h \leq b\) when the glass is filled to half its height. Report the value as a percentage of the total volume.

Question

Now, what height of filling will produce half the volume when? Report your answer in terms of a percentage of \(b\), height of the glass.

Example, arc length

The basic indefinite integral for a positive function answers the amount of area under the curve over a given interval. However, the integral can be interpreted in many different ways. For example, one can use an integral to answer how long a curve is. Of course one can estimate this answer. For example, consider this curve:

plot(x -> x^2, 0, 1)

This curve has length no more than \(2 = 1 + 1\) -- the distance along the \(x\) axis starting at \(0\) to \(1\) and then going up. It is also longer than \(\sqrt{2} = \sqrt{1^2 + 1^2}\) -- the straight line distance between the two endpoint. But how long is it?

In general, the arc length of the curve \(y=f(x)\) between \(a \leq x \leq b\) (or how long is the curve) is given through the formula

\[ l = \int_a^b \sqrt{1 + f'(x)^2} dx \]

The formula is from the length of the hypotenuse of a right triangle with lengths \(1\) and \(f'(x)\), though why is left for another day.

The arc length is easily computed using numeric integration. For example, our answer for \(f(x) = x^2\) is given by

f(x) = x^2
quadgk(x -> sqrt(1 + f'(x)^2), 0, 1)
## "MethodError(ctranspose,(f,))"

Whereas, the length of the \(f(x) = \sin(x)\) over \([0, \pi]\) would be:

f(x) = sin(x)
quadgk(x -> sqrt(1 + f'(x)^2), 0, pi)
## "MethodError(ctranspose,(f,))"

Next we look at a more practical problem.

Example: the caternary shape

A caternary shape (http://en.wikipedia.org/wiki/Catenary) is the shape a hanging chain will take as it is suspended between two posts. Of course, power wires will also have this shape between towers. A formula for a caternary can be written in terms of the hyperbolic cosine, cosh in julia:

\[ y = a \cosh(x/a) = a \cdot \frac{e^{x/a} + e^{-x/a}}{2}. \]

Suppose we have the following wire hung between \(x=-1\) and \(x=1\) with \(a = 2\):

a = 2;
f(x) = a * cosh(x/a)
plot(f, -1, 1)

How long is the chain? Looking at the graph we can guess an answer is between \(2\) and \(2.5\), say, but it isn't much work to get much closer to the answer:

quadgk(x -> sqrt(1 + f'(x)^2), -1, 1)
## "MethodError(ctranspose,(f,))"

Practice

Question

The sag in the chain is adjusted through the parameter \(a\) -- chains with larger \(a\) have less sag.

Suppose your chain has parameter a= 2.58 what is the length? (Use quadgk)

Question Bridges are parabolas

Suspension bridges, like the Verrazano bridge, have different loading than a cable and hence a different shape. A parabola is the shape the cable takes under uniform loading (cf. page 19 of http://calteches.library.caltech.edu/4007/1/Calculus.pdf for a picture).

The Verrazano-Narrows bridge has a span of 1298m. Suppose the drop of the main cables is 147 meters over this span. Then the cable itself can be modeled as a parabola with

  • \(x\)-intercepts \(a = 1298/2\) and \(-a\) and
  • vertex \((0,b)\) with \(b=-147\).

The parabola that fits these three points is

\[ y = \frac{-b}{a^2}(x^2 - a^2) \]

Find the arc length of the cable in meters.

Question The tractrix

A boat sits at the point \((a, 0)\) and a man holds a rope taut attached to the boat at the origin \((0,0)\). The man walks on the \(y\) axis. The position \(y\) depends then on the position \(x\) of the boat, and if the rope is taut, the position satisfies:

\[ y = a \ln\frac{a + \sqrt{a^2 - x^2}}{x} - \sqrt{a^2 - x^2} \]

This can be entered into julia as:

g(x, a) = a * log((a + sqrt(a^2 - x^2))/x) - sqrt(a^2 - x^2)

Let \(a=\)16, \(f(x) = g(x, a)\). Compute the length the bow of the boat has traveled between \(x=1\) and \(x=a\) using quadgk.

(The most elementary description of this curve is in terms of the relationship \(dy/dx = -\sqrt{a^2-x^2}/x\) which could be used in place of D(f) in your work.)

More on the tractrix

Watch this video "bicycle tracks" to see an example of how the tractrix can be found in an everyday observation.

Gauss quadrature, Using weights

What components go into the quadgk function? This section covers some of the background.

The trapezoid rule can be rearranged to become:

\[ \delta f(x_0) + 2\delta f(x_2) + 2 \delta f(x_3) + \cdots + 2 \delta f(x_{n}) + \delta f(x_{n}) \]

where \(\delta = (b-a)/n\).

Whereas for even \(n\), Simpson's rule can be written with:

\[ \delta f(x_0) + 4\delta f(x_1) + 2 \delta f(x_2) + \cdots + 4 \delta f(x_{n-2}) + 2 \delta f(x_{n-1}) + \delta f(x_{n}) \]

with \(\delta = (b-a)/n \cdot (1/6)\).

These both have the general form of

\[ \sum_k w_k \cdot f(x_k) \]

where \(w_k\) are weights and the \(x_k\) some choice of points -- not necessarily evenly spaced, though that is so in the examples we've seen. So, an alternative way to do the trapezoid formula in julia for \(n=4\) might be:

a, b, n = 0, 1, 4
## (0,1,4)
delta = (b-a)/n
## 0.25
f(x) = x^2
w = delta * [1, 2, 2, 2, 1] ## delta * [1, repeat([2],n-1), 1]'
## [0.25,0.5,0.5,0.5,0.25]
x = a + (0:n) * delta
## 0.0:0.25:1.0
sum(w .* [f(xi) for xi in x])         ## note .* has the leading dot
## 0.6875

The compact code to compute the approximate integral, sum(w .* [f(xi) for xi in x]), shows there are three important things in this form of the integral: the weights, the nodes or \(x\) values, and the function. The use of equally spaced nodes has been used by us so far, but it need not be the case. If fact Gauss showed he could get similar answers faster if it wasn't the case.

The Gauss nodes and weights are computable (http://en.wikipedia.org/wiki/Gaussian_quadrature). The main tools are the so-called Legendre polynomials, which can be defined recursively with Bonnet's formula:

\[ P_0(x) = 1; P_1(x) = x; \quad n P_{n}(x) = (2(n-1)+1) x P_{n-1}(x) -(n-1) P_{n-2}(x). \]

Using julia's Polynomial package this can be implemented almost verbatim:

using Polynomial
function lgp(n::Integer) 
    if n == 0 return Poly([1]) end
    if n == 1 return Poly([1, 0]) end
    
    (2*(n-1) + 1) / n * lgp(1) * lgp(n-1) - (n-1)/n * lgp(n-2)
end

This is used as,

p4 = lgp(4)
## Poly(4.375x^4 + -3.75x^2 + 0.375)

The term recursion is applied to a function when it makes a reference to itself during a computation. With this function, don't try it with values much bigger than \(20\), as the recursion can take a long time.

The nodes are the roots of the right polynomial. Here we have the values for p4

xk = roots(p4)
## [-0.3399810435848559,0.3399810435848554,-0.8611363115940524,0.8611363115940529]

(The Konrod part of quadgk changes the nodes so they can be reused during the refinement.)

Finally, the weights involve the derivative of \(P_n\) through:

\[ w_i = \frac{2}{(1 - x_i^2) \cdot(P^{'}_n(x_i)/P_n(1))^2} \]

These can be done simply with:

P =  p4 /polyval(p4, 1) ## normalize p4 so it is 1 at 1.
## Poly(4.375x^4 + -3.75x^2 + 0.375)
weights(x) = 2 / (1 - x^2) / (polyval(polydir(P), x))^2
wk = [weights(xi) for xi in xk]
## {0.6521451548625462,0.6521451548625466,0.34785484513745457,0.34785484513745296}

From here gauss_quadrature will do the integration of f over the interval \([-1,1]\), though we can do it ourself quickly enough. Here we compute the integral of \(\cos(\pi/2 x)\) over \([-1,1]\) (you can check this is very close to the answer \(4/\pi\) even with just 4 nodes):

f(x) = cos(pi/2*x)
sum( wk .* [f(xi) for xi in xk])
## 1.2732295042595088

Adaptive integration

Next, we a have a brief discussion about an alternative means to compute integrals. The following function adapt implements a basic adaptive quadrature method for integration. The basic idea is that for a subinterval \([a,b]\) if the area of the trapezoid is not close to the area of Simpson's parabolic estimate then the subinterval is split into two pieces \([a,c]\) and \([c,b]\) and the same question is asked. If the area is close the Simpson's parabolic estimate is used to estimate the integral of \(f\) over that subinterval.

Again, we see recursion when programming this algorithm. To avoid infinite loops during this, we use a limit below to keep track.

In general, the value of adaptive methods like this, is the function calls concentrate on areas where \(f\) is not well approximated and where it is well approximated it just moves on. This approach works well for poorly behaved functions, as it has a more refined grid there.

function adapt(f, a, b, limit)
    close_enough(x, y) = abs(x - y) < sqrt(eps())

    ##  println("adapt called with a=$a, b=$b, limit=$limit")

    h = b-a
    c = (a + b)/2
 
    a1 = (f(a) + f(b)) * h/2          ## trapezoid
    a2 = (f(a) + 4f(c) + f(b)) * h/6  ## Simpson's parabola

    if close_enough(a1, a2) 
        return(a2)
    end

    if limit == 0
        println("limit reached for this interval [$a, $b]") 
    return(a2)
    end

    adapt(f, a, c, limit - 1) + adapt(f, c, b, limit-1)
end

Does it work? Let's see it for the area of \(f(x) = x^2(1-x)^{10}\) which is known to satisfy \(\beta(2+1, 10+1)\)

out = adapt(x -> x^2 * (1 -x)^10, 0, 1, 10)
## 0.0011655014497573913

Not too far off (1e-10) from the known answer which is a beta function:

out - beta(2 + 1, 10 + 1)
## 2.842562207752003e-10