10. Integration

10.1. Introducing Integration

10.1.1. The “anti-derivative”

Consider the function \(\cos(x)\). We know that this function can be obtained by differentiating \(\sin(x)\). Therefore, we could think of \(\sin(x)\) as the “anti-derivative” of \(\cos(x)\). After differentiating \(\sin(x)\), the “anti-derivative” takes us back to where we started.

Well… nearly. We could also obtain \(\cos(x)\) by differentiating \(\sin(x)+2\), so we cannot say that \(\sin(x)\) is the unique anti-derivative of \(f(x)\).

In general, for a given function \(f(x)\),

if \(\frac{\mathrm{d}F}{\mathrm{d}x} = f(x)\), then \(F(x)+c\) is an antiderivative of \(x\),

where \(c\) is an arbitrary constant.

Being able to “spot” anti-derivatives is a tremendously useful (and often under-utilised) trick.

10.1.2. Area under the curve

Definition

Fig. 10.1 shows how the area under a curve \(f(x)\) between \(x=a,b\) may be estimated by drawing rectangles of equal width \(\Delta x\). Formally, we may write that:

(10.1)\[\textrm{Area} = \sum_{j=0}^{n-1}f(x_n)\Delta x \textrm{, where }x_0=a, x_n=b, \Delta x = \frac{b-a}{n}\]

The expression simply states that we add up the rectangles with width \(\Delta x\) and height \(f(x_n)\). As \(\Delta x\rightarrow 0\) the area approximation becomes more accurate.

We call:

\[\lim_{\Delta x \rightarrow 0}\sum_{j=0}^{n-1}f(x_j)\Delta x\]

the integral of \(f\) with respect to \(x\) and we denote it by:

(10.2)\[ \displaystyle \int_a^b f(x)\mathrm{d}x\]
../_images/areaUnderTheCurve.png

Fig. 10.1 Rectangles drawn under the curve \(f(x)\) can be used to approximate the area. The thinner the rectangles, the better the approximation. Integration is the equivalent to this approach as the width of the rectangles, \(\Delta x \rightarrow 0\).

It can be proved that integration and anti-differentiation are essentially the same thing. This result is known as the Fundamental Theorem of Calculus. In fact, the theorem comes in two parts. The first part, given in the box below, allows the definite integral to be determined by evaluating an anti-derivative at integration limits. The second part (not given here) essentially proves that every continuous function has an anti-derivative, which can be found by integration. For some functions, such as \(e^{-x^2}\), we might not be able to express the integral in terms of elementary functions, but we know it exists (and we can use successive approximation techniques either analytically or numerically to estimate them).

Definition

The First Fundamental Theorem of Calculus (FTC1)

If \(F\) is an anti-derivative of \(f\), then \(\displaystyle \int_a^b f(x) \textrm{d}x = F(b) - F(a)\) gives the area between the curve \(f(x)\) and the x-axis between \(x=a\) and \(x=b\).

Symbol

Meaning

Notes

\(\textrm{d}x\)

“with respect to x” The integration is along the x-axis. We say that x is the “integration variable”

Do not forget to write this, or the integral has no meaning.

\(\displaystyle \int\)

Integration operator

An operator acs on a function in some way. For example, \(\frac{\textrm{d}}{\textrm{d}x}\) is also an operator.

\(f(x)\)

Integrand

The function to be integrated.

\(a, b\)

The limits of integration.

If the integral has limits then it is called a definite integral and has a single value for the solution. If the integration does not have limits then it is called an indefinite integral, and you can only compute the result to within an arbitrary constant.

10.1.3. Signed and unsigned areas

Note that the heights of the rectangles in Fig. 10.1 are given by \(f(x)\), and so if \(f(x)<0\) then the result will be negative. In many examples, positive and negative areas cancel. For instance,

(10.3)\[\displaystyle \int_0^{2\pi}\sin(x)\mathrm{d}x = 0 \textrm{, since } \displaystyle \int_0^{\pi}\sin(x)\mathrm{d}x = 2 \textrm{ and } \displaystyle \int_{\pi}^{2\pi}\sin(x)\mathrm{d}x = -2\]

If you want the total (unsigned) area between the curve and the axis, then you can work with the modulus of the function,

(10.4)\[\displaystyle \int_a^b |f(x)|\mathrm{d}x\]

The easiest way to do this is to split up the region of integration into areas where \(f(x)>0\) and areas where \(f(x)<0\). For example,

(10.5)\[\displaystyle \int_0^{2\pi}|\sin(x)|\mathrm{d}x = \displaystyle \int_0^{\pi}\sin(x)\mathrm{d}x+ \displaystyle \int_{\pi}^{2\pi}-\sin(x)\mathrm{d}x = 2+2 =4\]

The results are illustrated in Fig. 10.2:

../_images/my_humps.png

Fig. 10.2 The plot on the left shows \(\sin(x)\) for \(x\in[0,2\pi]\). The black and red signed areas (2,-2) cancel when the integral is performed. The plot on the right shows \(|\sin(x)|\) for \(x\in[0,2\pi]\). The black signed are both positive and so the integral is 2+2=4.

Similarly, if we perform the integration along the x-axis in the negative direction, then the quantities \(\Delta x\) are negative, and so we end up with a negative signed area when \(f(x) > 0\).

FTC1 also confirms that \(\displaystyle \int_b^a f(x) \textrm{d}x = -\displaystyle \int_a^b f(x) \textrm{d}x\).

10.1.4. Area between two curves

../_images/between_curves.png

Fig. 10.3 Where \(f(x) >= g(x)\) the integral \(\displaystyle \int (f(x)-g(x)) \textrm{d}x\) gives the area between the two curves.

10.2. Integration by substitution

Definition

For an indefinite integral:

\[\int f(x) \textrm{d}x = \int f(x) \frac{\mathrm{d}x}{\mathrm{d}u}\mathrm{d}u+c\]

For a definite integral:

\[\int_a^b f\mathrm{d}x= \int_{u(a)}^{u(b)}f\frac{\mathrm{d}x}{\mathrm{d}u}\mathrm{d}u\]

Don’t forget to change the limits for definite integration!

These results can be proved using the chain rule and fundamental theorem of calculus.

The idea of using a substitution is that we convert a difficult integral w.r.t \(x\) to an easier one w.r.t \(u\).

Note that \(\displaystyle f \frac{\mathrm{d}x}{\mathrm{d}u}=f\biggr/\frac{\mathrm{d}u}{\mathrm{d}x}\) and so the substitution \(u\) can be chosen to eliminate a factor from \(f\).

Worked example

To calculate \( \displaystyle \int x\sqrt{x^2-1}\mathrm{d}x\) we can try the substitution \(u=x^2-1\). This substitution will eliminate the factor \(x\) and will simplify the expression \(\sqrt{x^2-1}\). We find that:

\(\displaystyle \frac{\mathrm{d}u}{\mathrm{d}x}=2x\) and so \(\displaystyle \frac{\mathrm{d}x}{\mathrm{d}u}=\frac{1}{2x}\), hence:

\[\int x\sqrt{x^2-1}\mathrm{d}x=\displaystyle \int\frac{1}{2}u^{1/2}\mathrm{d}u=\frac{1}{3}u^{3/2}+c=\frac{1}{3}(x^2-1)^{3/2}+c\]

This is the same result we can find by inspection.

10.3. Integral of \(\displaystyle \frac{1}{x}\)

The function \(\displaystyle \frac{1}{x}\) falls into a special category, because it cannot be solved using the polynomial antiderivative \(\displaystyle \int x^n = \frac{x^{n+1}}{n+1}\) as \(n=-1\) here.

../_images/oneOverx.png

Fig. 10.4 The plot of \(\displaystyle y=\frac{1}{x}\), showing limits of integration \(\pm a\) and \(\pm b\).

However we can think about the fact that \(\displaystyle \frac{\mathrm{d}}{\mathrm{d}x}\Big(\ln(x)\Big) = \frac{1}{x}\) and therefore use this as an antiderivative. From Fig. 10.4, we see that:

\[\int_{-b}^{-a}\frac{1}{x}\mathrm{d}x = -\int_a^b\frac{1}{x}\mathrm{d}x = \int_b^a\frac{1}{x}\mathrm{d}x = \int_{-b}^{-a}\frac{1}{|x|\mathrm{d}x}\]

So, \(\displaystyle \int\frac{1}{x}=\ln(|x|)+\mathrm{const},\, \forall x\), provided that the range of integration does not cross the vertical axis.

This is because formally \(\displaystyle \int_{-a}^a\frac{1}{x}\mathrm{d}x\) is undefined, although we can define what is called the Cauchy principal value of this integral as:

\[\lim_{\epsilon\rightarrow 0^{+}}\left( \int_{-a}^{-\epsilon}\frac{1}{x}\mathrm{d}x+ \int_{\epsilon}^a\frac{1}{x}\mathrm{d}x\right)=0\]

10.4. Integration by parts

Definition

For an indefinite integral:

\[\int u\,\frac{\mathrm{d}v}{\mathrm{d}x}\,\mathrm{d}x = u\,v- \int v\,\frac{\mathrm{d}u}{\mathrm{d}x}\,\mathrm{d}x+c\]

For a definite integral:

\[\int_a^b u\,\frac{\mathrm{d}v}{\mathrm{d}x}\,\mathrm{d}x=\biggr[u\,v\biggr]_a^b- \int_a^b v\frac{\mathrm{d}u}{\mathrm{d}x}\,\mathrm{d}x\]

This result can be proved using the product rule and the fundamental theorem of calculus.

\[\begin{split}\mathrm{d}(u\,v) &= u\,\mathrm{d}\,v + v\,\mathrm{d}u \\ u\,\mathrm{d}\,v &= \mathrm{d}(u\,v) - v\,\mathrm{d}u \\ \int u\,\mathrm{d}\,v &= \int \mathrm{d}(u\,v) - \int v\,\mathrm{d}u \\ \Rightarrow \int u\,\mathrm{d}\,v &= u\,v - \int v\,\mathrm{d}u\end{split}\]

For a handy rule of thumb for choosing which term to differentiate when applying this method, we can use LIATE method:

  • Logarithmic Term

  • Inverse Term

  • Algebraic Term

  • Trignometric Term

  • Exponential Term

This list tells the order of preference for the term to be differentiated \(u(x) \rightarrow u'(x)\). The exponential function is at the bottom of the sit because it is usually the easiest to integrate, the logarithmic is usually the hardest, hence it sits at the top.

Worked Examples

1.

\[I_a = \int x \ln(x)\,\mathrm{d}x \]

Since L falls before A in LIATE, we pick:

\[\begin{split}u = \ln(x),\quad & v' = x\\ u' = \frac{1}{x}, \quad & v = \frac{1}{2}x^2\end{split}\]

and so integrating by parts we find:

\[I_a = \frac{1}{2} \,x^2 \,\ln(x) - \int \frac{1}{2} \,x \,\mathrm{d}x = \frac{1}{2} \, x^2 \left(\ln(x) - \frac{1}{2} \right) + C\]

2. Sometimes it is not always obvious that there are two functions in the integration expression, however this can be remedied using a simple trick:

\[I_b = \int \ln(x)\,\mathrm{d}x \]

This should be seen as:

\[I_b = \int x^0\,\ln(x)\,\mathrm{d}x \]

Since L falls before A in LIATE, we pick:

\[\begin{split}u = \ln(x),\quad & v' = x^0 \\ u' = \frac{1}{x},\quad &v = x^1\end{split}\]

and so integrating by parts we find:

\[I_b = x \ln(x) - \int \frac{x}{x}\,\mathrm{d}x = x\ln(x) - \int x^0\,\mathrm{d}x = x\ln(x) - x + C \]

10.5. Integration by partial fractions

Composing fractions means putting everything together, such as:

\[\frac{A}{x+a} + \frac{B}{x+b} = \frac{(A+B)x + (Ab+Ba)}{(x+a)(x+b)}\]

Decomposition is the reverse process, i.e. taking the fraction apart. This very useful for integration, since the resulting partial fractions are much easier to integrate to logarithmic type terms.

We can combine together our facts about partial fractions into:

Factor in denominator

Term in partial fraction decomposition

\(ax+b\)

\(\displaystyle \frac{A}{ax+b}\)

\((ax+b)^n\)

\(\displaystyle \frac{A_1}{ax+b} + \frac{A_2}{(ax+b)2} + \dots \frac{A_n}{(ax+b)^n}, \quad n \in \mathbb{N}\)

\(ax^2+bx+c\)

\(\displaystyle \frac{Ax+B}{ax^2+bx+c}\)

\((ax^2+bx+c)^n\)

\(\displaystyle \frac{A_1}{ax^2+bx+c} + \frac{A_2}{(ax^2+bx+c)^2} + \dots \frac{A_n}{(ax^2+bx+c)^n}, \quad n \in \mathbb{N}\)

Worked example

Find the value of the integral

\[\int \frac{5x+7}{x^2+3x+2}\,\mathrm{d}x\]

Since this is an algebraic fraction we can try to express as partial fractions:

\[\begin{split}\frac{5x+7}{x^2+3x+2} &= \frac{5x+7}{(x+1)(x+2)}\\ &= \frac{A}{x+1}+\frac{B}{x+2} = \frac{A(x+2) + B(x+1)}{(x+1)(x+2)}\\ &= \frac{(A+B)x + 2A+B}{(x+1)(x+2)} \\ \Rightarrow 5x + 7 &= (A+B)x + 2A+B\end{split}\]

Which leads to simultaneous equations:

\[\begin{split}A + B &=5 \\ 2A + B &= 7\end{split}\]

and is solved by \(A = 2,\, B = 3\), therefore:

\[\begin{split}\frac{5x+7}{x^2+3x+2} &= \frac{2}{x+1}+\frac{3}{x+2} \\ \, \\ \int \frac{5x+7}{x^2+3x+2}\,\mathrm{d}x &= 2\ln|x+1| + 3\ln|x+2| + C\end{split}\]

10.6. Reduction formula

Reduction formulae set up a recurrence relation between different integrands:

\[I_{n} = f(n,x) + I_{n-a} \]

where \(a,\,n\) are integers typically and \(a < n\). Usually we solve these problems by integration by parts, which by spotting the required pattern, makes finding the solution more straightforward.

Worked example

Find and evaluate the reduction formula for:

\[I_n = \int_0^\infty x^n\,e^{-x}\,\mathrm{d} x\]

We can integrate by parts (using LIATE if necessary):

\[\begin{split}\begin{array}{lcl} u(x) = x^n && v'(x) = e^{-x} \\ u'(x) = nx^{n-1} && v(x) = -e^{-x} \end{array}\end{split}\]

Which means that we can write:

\[\begin{split}I_n &= \bigg[-x^n\,e^{-x}\bigg]_0^\infty + n\int_0^\infty x^{n-1}\, e^{-x}\,\mathrm{d} x \\ I_n &= 0 + n I_{n-1}\end{split}\]

If we were to repeat the integration by parts repeatedly we would find:

\[I_n = nI_{n-1} = n(n-1)\,I_{n-2} = \dots = n(n-1)(n-2)\dots (2)(1)\,I_0 = n!\, I_0\]

\(I_0\) is a good place to stop, since we can evaluate it:

\[I_0 = \int_0^\infty e^{-x}\,\mathrm{d} x = \bigg[-e^{-x}\bigg]_0^\infty = 0 - (-1) = 1\]

Hence \(I_n = n!\).

10.7. Mean value of a function

../_images/mean.png

Fig. 10.5 The average of a discrete function is given by adding up the values and dividing by the number of points. We extend this concept to continuous functions in the limit as \(\Delta x\rightarrow 0\).

As shown in Fig. 10.5, for any function \(f(x)\) we can obtain the average output \(\bar{f(x)}\) by adding up a series of outputs and dividing by the number of datapoints used. In this case, the spacing between datapoints is \(\Delta x\).

\[\bar{f} = \frac{1}{n} \sum_{j=0}^{n-1}f(x_j) = \frac{\Delta x}{b-a}\sum_{j=0}^{n-1}f(x_j)\]

By taking the limit as \(\Delta x\rightarrow 0\), we can obtain an average of the function \(f(x)\) and we find that this result is equal to the limit defintion of the definite integral.

Definition

The mean value \(\bar{f}\) of the function \(f(x)\) on the domain \(x \in [a,\, b]\) is given by:

\[\bar{f} = \frac{1}{b-a} \int_a^b f(x)\mathrm{d}x\]

Worked example

Calculate the mean value of \(\cos ^3(\theta)\) on the interval \(\displaystyle \left[-\frac{\pi}{3}, \frac{\pi}{3}\right]\) by integration.

\[\begin{split}\frac{1}{\frac{\pi}{3} - \left(-\frac{\pi}{3}\right)} \int_{-\pi/3}^{\pi/3}\cos^{3}(\theta)\mathrm{d}\theta &= \frac{3}{2\pi}\displaystyle \int_{-\pi/3}^{\pi/3}\cos(\theta)(1-\sin^2(\theta))\mathrm{d}\theta\\ &=\frac{3}{2\pi}\left[\sin(\theta)-\frac{1}{3}\sin^3(\theta)\right]_{-\pi/3}^{\pi/3} = \frac{9\sqrt{3}}{8\pi}\end{split}\]

10.8. Arc Length

Suppose that we have a function \(y = f(x)\), we know through integration that we can find the area under the curve:

../_images/FunctionArea.png

Fig. 10.6 The area (in blue) under a function \(f(x)\) over the interval \(x \in [a,\,b]\) given by the integral \(\displaystyle \int_a^bf(x)\,\mathrm(d)x\).

Another question that we could ask however is what is the length of the path traced out by the function \(f(x)\) over the range \([a,\,b]\):

../_images/FunctionPathLength.png

Fig. 10.7 The path traced out by the function \(f(x)\) over the interval \(x \in [a,\,b]\).

To find an expression for this length \(S\), we can break up the path into infinitesimal segments and then integrate these over the whole function. We can see this process breaking down the change in arc length \(\Delta s\) by Pythagoras graphically as:

\[(\Delta s)^2 = (\Delta x)^2 + (\Delta y)^2 \Rightarrow \Delta s = \sqrt{(\Delta x)^2 + (\Delta y)^2} = \Delta x\,\sqrt{1 + \left(\frac{\Delta y}{\Delta x}\right)^2}\]
../_images/ArcLength2.png

Fig. 10.8 Left Pane: Breaking down path legnth segment \(\Delta s\) into \(x,\, y\) components \(\Delta x,\, \Delta y\). The distance \(\Delta s\) moved along a curve, corresponding to small changes \(\Delta x \Delta y\) can be calculated by Pythagoras’ formula.

Right Pane: Effect of taking the limit of smaller \(\Delta x\) in finding the path length. By decreasing the difference \(\Delta x\) between the joined up points, we obtain a better approximation.

Taking the limit of \(\Delta x\rightarrow 0\) we find:

\[\mathrm{d}s = \mathrm{d}x\, \sqrt{1 + \left(\frac{\mathrm{d}y}{\mathrm{d}x}\right)^2}\]

and therefore to find the path length:

\[L = \int_{x=a}^{x=b} \mathrm{d}s = \int_a^b \sqrt{1 + \left(\frac{\mathrm{d}y}{\mathrm{d}x}\right)^2}\,\mathrm{d}x\]

We might also have a function which is parameterised \(y = y(t),\, x=x(t)\), it is also possible to find an expression for the path length for \(t \in [t_a,\, t_b]\), taking the limit of \(\Delta t \rightarrow 0\):

\[\begin{split}\Delta s &= \Delta t\,\sqrt{\left(\frac{\Delta x}{\Delta t}\right)^2 + \left(\frac{\Delta y}{\Delta t}\right)^2} \\ \Rightarrow \mathrm{d}s &= \mathrm{d}t\, \sqrt{\left(\frac{\mathrm{d}x}{\mathrm{d}t}\right)^2 + \left(\frac{\mathrm{d}y}{\mathrm{d}t}\right)^2}\\ L &= \int_{t_a}^{t_b} \sqrt{\left(\frac{\mathrm{d}x}{\mathrm{d}t}\right)^2 + \left(\frac{\mathrm{d}y}{\mathrm{d}t}\right)^2}\,\mathrm{d}t\end{split}\]

Worked examples

1. Looking at \(y = \cosh(x)\) over the range \(x \in [0,\,1]\), so we find \(\mathrm{d}y/\mathrm{d}x = \sinh(x)\), and therefore:

\[L = \int_0^1 \sqrt{1 + \sinh^2(x)}\,\mathrm{d}x = \int_0^1 \cosh(x)\,\mathrm{d}x = \Big[ \sinh(x) \Big]_0^1 = \sinh(1) \simeq 1.175\dots\]

2. Consider a circle, parameterised by \(x = R\cos(t),\, y = R\sin(t)\), over the range \(t \in [0,\, 2\pi)\), giving \(\mathrm{d}x/\mathrm{d}t = -R\sin(t),\, \mathrm{d}y/\mathrm{d}t = R\cos(t)\) and therefore:

\[L = \int_0^{2\pi} \sqrt{R^2\sin^2(t) + R^2\cos^2(t)}\,\mathrm{d}t = \int_0^{2\pi} R\,\mathrm{d}x = \Big[ Rt \Big]_0^{2\pi} = 2\pi\,R\]

which gives the result for the circumference of a circle!

10.9. Surfaces of Revolution

Lets now go further, what happens if we take a function and rotate it around an axis:

../_images/AreaVolumeRevolution.png

Fig. 10.9 Left Pane: Rotating a function \(y = f(x)\) around the \(x\) axis, over \(x \in [a,\, b]\) producing a solid of revolution, Right Pane: Breaking down the volume into discrete slices, each of width \(\mathrm{d} x\).

To find the volume of such a rotated solid, we need to think about the cross sectional area along the range, which will be \(\pi \,y^2\) since the radius of each slice at \(x\) is \(y(x)\). Then we just need to integrate up these slices \(\pi\,y^2\,\mathrm{d}x\) over the range:

\[V_x = \int_a^b \pi\,y^2\,\mathrm{d}x\]

and we can swap round the axis we rotate over to the \(y\) axis and therefore the radius of each slice would be \(\pi\,x^2\) and therefore:

\[V_y = \int_{y(a)}^{y(b)} \pi\,x^2\,\mathrm{d}y\]

Note

Instead of using cylindrical slices, we could use “frustrums” (sections of cones). The result would be the same (in the limit \(\Delta x \rightarrow 0\)).

Similarly, when calculating the area under the curve, we used rectangles, but we could have used parallelograms.

We can also find the surface area of this solid of revolution, if we rotate over the \(x\) xcis then this surface area is the circumference of each slice \(2\pi\,y\) multiplied by the path length \(\mathrm{d} s\) along the surface, so we find:

\[A_x = \int_{x=a}^{x=b} 2\pi\,y\,\mathrm{d}s = \int_a^b 2\pi\,y\,\sqrt{1 + \left(\frac{\mathrm{d}y}{\mathrm{d}x}\right)^2}\,\mathrm{d}x\]

We could rotate over the \(y\) axis instead, in which case we just switch the circumference we are interested in to be \(2\pi\,x\) and integrate the path length over the \(y\) axis:

\[A_y = \int_{y(a)}^{y(b)} 2\pi\,x\,\mathrm{d}s = \int_{y(a)}^{y(b)} 2\pi\,x\,\sqrt{1 + \left(\frac{\mathrm{d}x}{\mathrm{d}y}\right)^2}\,\mathrm{d}y\]

Worked example

Lets find the volume of a cone depicted in Fig. 10.10, with height \(h\) and circular radius \(R\), so \(y = R x / h\) over the range \(x \in [0,\, h]\):

\[\begin{split}V_x = \int_0^h \pi\,\left(\frac{R x}{h}\right)^2\,\mathrm{d}x = \frac{R^2 \,\pi}{h^2}\Big[ \frac{1}{3}x^3\Big]_0^h = \frac{1}{3}\pi \,R^2\,h\\\end{split}\]

which matches the expression we expect and explains where this factor of \(1/3\) comes from!

Likewise we can look at the surface area of this cone, \(y = R x / h \rightarrow \mathrm{d}y/\mathrm{d}x = R / h\) over the range \(x \in [0,\, h]\):

\[\begin{split}A_x &= \int_0^h 2\pi\,\frac{R x}{h}\,\sqrt{1 + \left(\frac{R}{h}\right)^2}\,\mathrm{d}x \\ &= 2\pi \frac{R}{h}\sqrt{1 + \left(\frac{R}{h}\right)^2}\Bigg[ \frac{1}{2}x^2\Bigg]_0^h \\ &= \pi R h\sqrt{1 + \left(\frac{R}{h}\right)^2} = \pi R\sqrt{h^2 + R^2}\end{split}\]

where \(\sqrt{h^2 + R^2}\) is the slant length.

../_images/cone.png

Fig. 10.10 Depiction of a function \(y = R x / h\) and the solid of revolution around the \(x\) axis - a cone.

Likewise if we have parametrised expressions \(x = x(t),\, y = y(t)\), over the range \(t \in [t_1,\, t_2]\) then these expressions become:

\[\begin{split}V_x &= \int_{t_1}^{t_2} \pi\,y^2\,\frac{\mathrm{d}x}{\mathrm{d}t}\,\mathrm{d}t \\ V_y &= \int_{t_1}^{t_2} \pi\,x^2\,\frac{\mathrm{d}y}{\mathrm{d}t}\,\mathrm{d}t \\ A_x &= \int_{t_1}^{t_2} 2\pi\,y\,\sqrt{\left(\frac{\mathrm{d}x}{\mathrm{d}t}\right)^2 + \left(\frac{\mathrm{d}y}{\mathrm{d}t}\right)^2}\,\mathrm{d}t \\ A_y &= \int_{t_1}^{t_2} 2\pi\,x\,\sqrt{\left(\frac{\mathrm{d}x}{\mathrm{d}t}\right)^2 + \left(\frac{\mathrm{d}y}{\mathrm{d}t}\right)^2}\,\mathrm{d}t\end{split}\]