Math 105: Solutions to Practice Problems

77 downloads 1007 Views 377KB Size Report
Math 105: Solutions to Practice Problems. Steven Miller. May 13, 2010. Abstract ... 5.3 The Double and Triple Integral Over More General Regions . . . 21.
Math 105: Solutions to Practice Problems Steven Miller May 13, 2010 Abstract Below are detailed solutions to some problems similar to some assigned homework problems.

Contents 1 The Geometry of Euclidean Space

2

2 Differentiation 2.1 The geometry of real-valued functions 2.2 Limits and continuity . . . . . . . . . 2.3 Differentiation . . . . . . . . . . . . . 2.4 Introduction to paths and curves . . . 2.5 Properties of the derivative . . . . . . 2.6 Gradients and directional derivatives . 2.7 Review Exercises - Page 173 . . . . .

. . . . . . .

2 2 2 2 2 2 2 6

. . . .

8 8 10 12 14

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

3 Higher-order derivatives; maxima and minima 3.1 Iterated partial derivatives . . . . . . . . . . . 3.2 Taylor’s theorem . . . . . . . . . . . . . . . 3.3 Extrema of real-valued functions . . . . . . . 3.4 Constrained extrema and Lagrange multipliers 4 Vector-valued functions

. . . . . . .

. . . .

. . . . . . .

. . . .

. . . . . . .

. . . .

. . . . . . .

. . . .

. . . . . . .

. . . .

. . . . . . .

. . . .

. . . . . . .

. . . .

. . . . . . .

. . . .

. . . . . . .

. . . .

. . . . . . .

. . . .

17

1

5 Double and Triple Integrals 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 The Double Integral over a Rectangle . . . . . . . . . . . . 5.3 The Double and Triple Integral Over More General Regions 5.4 Changing the order of integration . . . . . . . . . . . . . . . 5.5 Mathematical Modeling: Baseball / Sabermetrics Lecture . .

. . . . .

17 17 19 21 23 25

6 Change of variable formula and applications of integration 6.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2 The Change of Variable Theorem . . . . . . . . . . . . . . . . . .

27 27 27

7 Sequences and Series 7.1 Page 10.3 . . . . 7.2 Page 10.6 . . . . 7.3 Page 10.7 . . . . 7.4 Page 10.8 . . . . 7.5 Page 10.10 . . . .

34 34 36 37 37 38

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

8 From path integrals to Stokes’ Theorem

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

39

1 The Geometry of Euclidean Space 2 Differentiation 2.1 The geometry of real-valued functions 2.2 Limits and continuity 2.3 Differentiation 2.4 Introduction to paths and curves 2.5 Properties of the derivative 2.6 Gradients and directional derivatives The assignment is: Section 2.6: #2ab, #4a, #6a, #16 (the (in)famous Captain Ralph problem), #18. 2

π‘₯ Question: #2c: Compute the ( directional ) derivative of 𝑓 (π‘₯, 𝑦) = 𝑒 cos(πœ‹π‘¦), βˆ’ (π‘₯0 , 𝑦0) = (0, βˆ’1) and β†’ 𝑣 = βˆ’ √15 , √25 . Solution: We have ) ( βˆ‚π‘“ βˆ‚π‘“ (βˆ‡π‘“ )(π‘₯, 𝑦) = , = (𝑒π‘₯ cos(πœ‹π‘¦), βˆ’πœ‹π‘’π‘₯ sin(πœ‹π‘¦)) , βˆ‚π‘₯ βˆ‚π‘¦

and thus if we evaluate at (0, βˆ’1) we find (βˆ‡π‘“ )(0, βˆ’1) = (βˆ’1, 0) . The directional derivative in general is βˆ’ (βˆ‡π‘“ )(π‘₯0 , 𝑦0 ) β‹… β†’ 𝑣, so for this problem the answer is (

1 2 (βˆ’1, 0) β‹… βˆ’ √ , √ 5 5

)

1 = √ . 5

βˆ’ Note that β†’ 𝑣 is a unit length vector. 2 3 Question: #2d: Compute the ( directional ) derivative of 𝑓 (π‘₯, 𝑦) = π‘₯𝑦 + π‘₯ 𝑦, βˆ’ (π‘₯0 , 𝑦0) = (4, βˆ’2) and β†’ 𝑣 = √110 , √310 . Solution: We have ( ) (βˆ‡π‘“ )(π‘₯, 𝑦) = 𝑦 2 + 3π‘₯2 𝑦, 2π‘₯𝑦 + π‘₯3 ,

and thus if we evaluate at (4, βˆ’2) we find

(βˆ‡π‘“ )(4, βˆ’2) = (βˆ’92, 48) . The directional derivative in general is βˆ’ (βˆ‡π‘“ )(π‘₯0 , 𝑦0 ) β‹… β†’ 𝑣, so for this problem the answer is (βˆ’92, 48) β‹…

(

1 3 √ ,√ 10 10 3

)

52 = √ . 10

βˆ’ Note that β†’ 𝑣 is a unit length vector. Question: #3a: Compute 𝐷𝑓 for 𝑓 (π‘₯, 𝑦) = (π‘₯2 𝑦, π‘’βˆ’π‘₯𝑦 ). Solution: Note that 𝑓 : ℝ2 β†’ ℝ2 , and thus 𝐷𝑓 will be a 2 Γ— 2 matrix. Writing 𝑓 (π‘₯, 𝑦) as (𝑓1 (π‘₯, 𝑦), 𝑓2 (π‘₯, 𝑦)), we have the first row of 𝐷𝑓 is βˆ‡π‘“1 , while the second row is βˆ‡π‘“2 . Explicitly, ( ) ( ) βˆ‚π‘“1 βˆ‚π‘“1 2π‘₯𝑦 π‘₯2 βˆ‚π‘₯ βˆ‚π‘¦ (𝐷𝑓 )(π‘₯, 𝑦) = = . βˆ‚π‘“2 βˆ‚π‘“2 βˆ’π‘¦π‘’βˆ’π‘₯𝑦 βˆ’π‘₯π‘’βˆ’π‘₯𝑦 βˆ‚π‘₯ βˆ‚π‘¦ Question: #4c: Find the plane tangent to π‘₯𝑦𝑧 = 1 at the point (1, 1, 1). Solution: We use equation (1) on the bottom of page 167, which says that if 𝑓 (π‘₯, 𝑦, 𝑧) = π‘˜ (for some constant π‘˜) then the tangent plane at (π‘₯0 , 𝑦0 , 𝑧0 ) is given by (βˆ‡π‘“ )(π‘₯0 , 𝑦0 , 𝑧0 ) β‹… (π‘₯ βˆ’ π‘₯0 , 𝑦 βˆ’ 𝑦0 , 𝑧 βˆ’ 𝑧0 ) = 0. For our problem, 𝑓 (π‘₯, 𝑦, 𝑧) = π‘₯𝑦𝑧 and π‘˜ = 1. We have

(βˆ‡π‘“ )(π‘₯, 𝑦, 𝑧) = (𝑦𝑧, π‘₯𝑧, π‘₯𝑦), which yields (βˆ‡π‘“ )(1, 1, 1) = (1, 1, 1). Thus the tangent plane is all (π‘₯, 𝑦, 𝑧) satisfying (1, 1, 1) β‹… (π‘₯ βˆ’ 1, 𝑦 βˆ’ 1, 𝑧 βˆ’ 1) = 0, or equivalently it is π‘₯ + 𝑦 + 𝑧 βˆ’ 3 = 0. Question: #6b: Compute the gradient of 𝑓 (π‘₯, 𝑦, 𝑧) = π‘₯𝑦 + 𝑦𝑧 + π‘₯𝑧. Solution: The gradient is defined as the vector of partial derivatives: ( ) βˆ‚π‘“ βˆ‚π‘“ βˆ‚π‘“ grad(𝑓 ) = βˆ‡π‘“ = , , . βˆ‚π‘₯ βˆ‚π‘¦ βˆ‚π‘§ A straightforward computation shows βˆ‡π‘“ = (𝑦 + 𝑧, π‘₯ + 𝑧, π‘₯ + 𝑦). 4

Question: #6c: Compute the gradient of 𝑓 (π‘₯, 𝑦, 𝑧) = 1/(π‘₯2 + 𝑦 2 + 𝑧 2 ). Solution: By symmetry, it suffices to compute βˆ‚π‘“ , as βˆ‚π‘“ and βˆ‚π‘“ are obtained βˆ‚π‘₯ βˆ‚π‘¦ βˆ‚π‘§ βˆ‚π‘“ through analogous computations. To compute βˆ‚π‘₯ , we use the one-variable chain (or power) rule: βˆ‚ 2 (π‘₯ + 𝑦 2 + 𝑧 2 )βˆ’1 βˆ‚π‘₯

=

βˆ’(π‘₯2 + 𝑦 2 + 𝑧 2 )βˆ’2

=

βˆ’

=

2π‘₯ + 𝑦 2 + 𝑧 2 )2 βˆ’2𝑓 (π‘₯, 𝑦, 𝑧)2 β‹… π‘₯.

) βˆ‚ ( 2 π‘₯ + 𝑦2 + 𝑧2 βˆ‚π‘₯

(π‘₯2

Collecting yields βˆ‡π‘“ = βˆ’2𝑓 (π‘₯, 𝑦, 𝑧)(π‘₯, 𝑦, 𝑧). Question: #8b: Compute the equation of the tangent planes for 𝑓 (π‘₯, 𝑦, 𝑧) = π‘₯3 βˆ’ 2𝑦 3 + 𝑧 3 = 0 at (1, 1, 1). Solution: First, we note that the point (1, 1, 1) is on the surface. The tangent plane is given by equation (1) on page 167. Explicitly, it is all (π‘₯, 𝑦, 𝑧) satisfying (βˆ‡π‘“ )(1, 1, 1) β‹… (π‘₯ βˆ’ 1, 𝑦 βˆ’ 1, 𝑧 βˆ’ 1). As βˆ‡π‘“ = (3π‘₯2 , βˆ’6𝑦 2 , 3𝑧 2 ), we have (βˆ‡π‘“ )(1, 1, 1) = (3, βˆ’6, 3), which implies the tangent plane is (3, βˆ’6, 3) β‹… (π‘₯ βˆ’ 1, 𝑦 βˆ’ 1, 𝑧 βˆ’ 1) = 0, or 3π‘₯ βˆ’ 6𝑦 + 3𝑧 = 0.

5

βˆ’ βˆ’ Question: #19: A function 𝑓 : ℝ𝑛 β†’ ℝ is said to be even if 𝑓 (β†’ π‘₯ ) = 𝑓 (βˆ’β†’ π‘₯ ) for β†’ βˆ’ β†’ βˆ’ all π‘₯ . If 𝑓 is differentiable and even, find (𝐷𝑓 )( 0 ). Solution: Whenever we have to prove something in several variables, it is not a bad idea to look at some examples from one-variable calculus to build up our intuition. We first recall some even, differentiable functions: π‘₯2 , π‘₯4 , π‘₯2𝑛 , cos π‘₯. All of these have first derivative equal to 0 at the origin, and thus it is natural to β†’ βˆ’ β†’ βˆ’ guess that (𝐷𝑓 )( 0 ) = 0 . βˆ’ βˆ’ One way to prove this is by using the Chain Rule. Let 𝑔(β†’ π‘₯ ) = βˆ’β†’ π‘₯ (so 𝑛 𝑛 𝑔 : ℝ β†’ ℝ ). Then βˆ’ βˆ’ 𝐴(π‘₯) = 𝑓 (β†’ π‘₯ ) = 𝑓 (𝑔(β†’ π‘₯ )), so

βˆ’ βˆ’ βˆ’ βˆ’ (𝐷𝐴)(β†’ π‘₯ ) = (𝐷𝑓 )(β†’ π‘₯ ) = (𝐷𝑓 )(𝑔(β†’ π‘₯ ))(𝐷𝑔)(β†’ π‘₯ ). βˆ’ βˆ’ As 𝑔(β†’ π‘₯ ) = βˆ’β†’ π‘₯ , unwinding this we find 𝑔(π‘₯1 , . . . , π‘₯𝑛 ) = (βˆ’π‘₯1 , . . . , βˆ’π‘₯𝑛 ), which implies

βˆ’ βˆ’ (𝐷𝑔)(β†’ π‘₯ ) = (βˆ‡π‘”)(β†’ π‘₯ ) = βˆ’πΌ,

where 𝐼 is the 𝑛 Γ— 𝑛 identity matrix which is 1 along the main diagonal and 0 elsewhere. The reason this is the answer is that 𝑔 has 𝑛 inputs and 𝑛 outputs. Thus (𝐷𝑔) is a matrix with 𝑛 rows and 𝑛 columns. The first row is 𝐷𝑔1 or βˆ‡π‘”1 , where 𝑔1 (π‘₯1 , . . . , π‘₯𝑛 ) = βˆ’π‘₯1 , while the last row is 𝐷𝑔𝑛 or βˆ‡π‘”π‘› . β†’ βˆ’ β†’ βˆ’ β†’ βˆ’ At the origin, 𝑔( 0 ) = 0 and (𝐷𝑔)( 0 ) = βˆ’πΌ, and thus β†’ βˆ’ β†’ βˆ’ β†’ βˆ’ (𝐷𝑓 )( 0 ) = (𝐷𝑓 )(βˆ’ 0 )(𝐷𝑔)( 0 )

becomes

β†’ βˆ’ β†’ βˆ’ β†’ βˆ’ (𝐷𝑓 )( 0 ) = (𝐷𝑓 )( 0 ) (βˆ’πΌ) = βˆ’(𝐷𝑓 )( 0 ). β†’ βˆ’ βˆ’ βˆ’ βˆ’ We thus have an equation of the form β†’ 𝑒 = βˆ’β†’ 𝑒 ; the only solution is β†’ 𝑒 = 0 , or β†’ βˆ’ in other words since (𝐷𝑓 )( 0 ) equals its own negative, it must be the zero vector.

2.7 Review Exercises - Page 173 Question: #22: Find the direction in which the function 𝑀(π‘₯, 𝑦) = π‘₯2 + π‘₯𝑦 increases most rapidly at the point (βˆ’1, 1). What is the magnitude of βˆ‡π‘€ at this point. 6

Solution: We have βˆ‡π‘€ = (2π‘₯ + 𝑦, π‘₯), so (βˆ‡π‘€)(βˆ’1, 1) = (βˆ’1, βˆ’1). As the βˆ’ βˆ’ directional derivative in the direction β†’ 𝑣 at (βˆ’1, 1) is (βˆ‡π‘€)(βˆ’1, 1) β‹… β†’ 𝑣 , which is βˆ’ maximized when β†’ 𝑣 = (βˆ‡π‘€)(βˆ’1, 1) = (βˆ’1, 1). Question: #24: Let 𝑧(π‘₯, 𝑦) = 𝑓 (π‘₯ βˆ’ 𝑦)/𝑦 (where 𝑓 is differentiable and 𝑦 βˆ•= 0), βˆ‚π‘§ βˆ‚π‘§ show that the identity 𝑧 + 𝑦 βˆ‚π‘₯ + 𝑦 βˆ‚π‘¦ = 0. Solution: We have βˆ‚π‘§ 𝑓 β€² (π‘₯ βˆ’ 𝑦)(βˆ’1) β‹… 𝑦 βˆ’ 𝑓 (π‘₯ βˆ’ 𝑦) = , βˆ‚π‘¦ 𝑦2

βˆ‚π‘§ 𝑓 β€² (π‘₯ βˆ’ 𝑦) = . βˆ‚π‘₯ 𝑦

Thus 𝑧+𝑦

βˆ‚π‘§ 𝑓 (π‘₯ βˆ’ 𝑦) βˆ’π‘“ β€² (π‘₯ βˆ’ 𝑦)𝑦 βˆ’ 𝑓 (π‘₯ βˆ’ 𝑦) βˆ‚π‘§ +𝑦 = + 𝑓 β€² (π‘₯ βˆ’ 𝑦) + = 0. βˆ‚π‘₯ βˆ‚π‘¦ 𝑦 𝑦

In the arguments above, we frequently used the one-variable chain rule. For example, βˆ‚ βˆ‚ 𝑓 (π‘₯ βˆ’ 𝑦) = 𝑓 (𝑔(𝑦)), βˆ‚π‘¦ βˆ‚π‘¦ where 𝑔(𝑦) = π‘₯ βˆ’ 𝑦. We can now use the one-variable chain rule. As π‘₯ is fixed, the answer is just 𝑓 β€² (𝑔(𝑦)) β‹… 𝑔 β€²(𝑦), which is 𝑓 β€² (π‘₯ βˆ’ 𝑦) β‹… (βˆ’1). Question: #44: Verify the chain rule for the function 𝑓 (π‘₯, 𝑦) = π‘₯2 /(2 + cos 𝑦) and the path 𝑐(𝑑) = (π‘₯(𝑑), 𝑦(𝑑)) = (𝑒𝑑 , π‘’βˆ’π‘‘ ). Solution: Setting 𝐴(𝑑) = 𝑓 (𝑐(𝑑)), we have (𝐷𝐴)(𝑑) = (𝐷𝑓 )(𝑐(𝑑))𝑐′(𝑑). We have 𝑐′ (𝑑) = (𝑒𝑑 , βˆ’π‘’βˆ’π‘‘ ) (which should really be written as a column vector). For 𝐷𝑓 , we have ( ) ( ) βˆ‚π‘“ βˆ‚π‘“ 2π‘₯ π‘₯2 sin 𝑦 𝐷𝑓 = , = ,βˆ’ ; βˆ‚π‘₯ βˆ‚π‘¦ 2 + cos 𝑦 (2 + cos 𝑦)2 however, we want (𝐷𝑓 )(𝑐(𝑑)), which is ( ) 2𝑒𝑑 𝑒2𝑑 sin π‘’βˆ’π‘‘ (𝐷𝑓 )(𝑐(𝑑)) = ,βˆ’ . 2 + cos π‘’βˆ’π‘‘ (2 + cos 𝑒𝑑 )2 Taking the dot product, we obtain (𝐷𝐴)(𝑑) =

2𝑒𝑑 𝑒2𝑑 sin π‘’βˆ’π‘‘ 𝑑 β‹… 𝑒 + β‹… π‘’βˆ’π‘‘ . 2 + cos π‘’βˆ’π‘‘ (2 + cos 𝑒𝑑 )2 7

We can also compute this derivative directly, as 𝐴(𝑑) = 𝑓 (𝑐(𝑑)) =

𝑒2𝑑 . (2 + cos π‘’βˆ’π‘‘ )

Taking the derivative yields 𝐴′ (𝑑) =

2𝑒2𝑑 (2 + cos π‘’βˆ’π‘‘ ) + 𝑒2𝑑 sin π‘’βˆ’π‘‘ , (2 + cos π‘’βˆ’π‘‘ )2

which does agree with the Chain Rule.

3 Higher-order derivatives; maxima and minima 3.1 Iterated partial derivatives 2

2

2

βˆ‚ 𝑓 βˆ‚ 𝑓 Question: #3: Compute βˆ‚βˆ‚π‘₯𝑓2 , βˆ‚π‘₯βˆ‚π‘¦ , βˆ‚π‘¦βˆ‚π‘₯ and verify the equality of the mixed derivatives. Solution: We have

βˆ‚π‘“ βˆ‚π‘₯ βˆ‚2𝑓 βˆ‚π‘₯2 βˆ‚2𝑓 βˆ‚π‘¦βˆ‚π‘₯

βˆ‚2𝑓 βˆ‚π‘¦ 2

for 𝑓 (π‘₯, 𝑦) = cos(π‘₯𝑦 2 ), and

=

βˆ’π‘¦ 2 sin(π‘₯𝑦 2)

=

βˆ’π‘¦ 4 cos(π‘₯𝑦 2 )

=

βˆ’2𝑦 sin(π‘₯𝑦 2 ) βˆ’ 2π‘₯𝑦 3 cos(π‘₯𝑦 2).

=

βˆ’2π‘₯𝑦 sin(π‘₯𝑦 2 )

=

βˆ’2π‘₯ sin(π‘₯𝑦 2 ) βˆ’ 4π‘₯2 𝑦 2 cos(π‘₯𝑦 2 )

=

βˆ’2𝑦 sin(π‘₯𝑦 2 ) βˆ’ 2π‘₯𝑦 3 cos(π‘₯𝑦 2).

Similarly, we find βˆ‚π‘“ βˆ‚π‘¦ βˆ‚2𝑓 βˆ‚π‘¦ 2 βˆ‚2𝑓 βˆ‚π‘₯βˆ‚π‘¦ Note that we do have

βˆ‚2𝑓 βˆ‚π‘₯βˆ‚π‘¦

=

βˆ‚2𝑓 . βˆ‚π‘¦βˆ‚π‘₯

8

Question: #8b: Find all the second partial derivatives of 𝑧(π‘₯, 𝑦) = π‘₯2 𝑦 2𝑒2π‘₯𝑦 . Solution: We have βˆ‚π‘§ βˆ‚π‘₯ βˆ‚2𝑧 βˆ‚π‘₯2 βˆ‚2𝑧 βˆ‚π‘¦βˆ‚π‘₯ βˆ‚π‘§ βˆ‚π‘¦ βˆ‚2𝑧 βˆ‚π‘¦ 2

=

2π‘₯𝑦 2 𝑒2π‘₯𝑦 + 2π‘₯2 𝑦 3 𝑒2π‘₯𝑦

=

2𝑦 2 𝑒2π‘₯𝑦 + 4π‘₯𝑦 3 𝑒2π‘₯𝑦 + 4π‘₯𝑦 3 𝑒2π‘₯𝑦 + 4π‘₯2 𝑦 4 𝑒2π‘₯𝑦

=

4π‘₯𝑦𝑒2π‘₯𝑦 + 4π‘₯2 𝑦 2 𝑒2π‘₯𝑦 + 6π‘₯2 𝑦 2 𝑒2π‘₯𝑦 + 4π‘₯3 𝑦 3 𝑒2π‘₯𝑦

=

2π‘₯2 𝑦𝑒2π‘₯𝑦 + 2π‘₯3 𝑦 2 𝑒2π‘₯𝑦

=

2π‘₯2 𝑒2π‘₯𝑦 + 4π‘₯3 𝑦𝑒2π‘₯𝑦 + 4π‘₯3 𝑦𝑒2π‘₯𝑦 + 4π‘₯4 𝑦 2 𝑒2π‘₯𝑦 .

Question: Supplemental problem related to #11: Use the fact that the derivative of a sum is the sum of the derivatives to prove that the derivative of a sum of three terms is the sum of the three derivatives. Solution: The idea to solve this problem is quite useful in mathematics (and may be useful to attacking #11). We know that for any two functions 𝑓 (π‘₯) and 𝑑𝑓 𝑑𝑔 𝑑 𝑔(π‘₯) that 𝑑π‘₯ (𝑓 (π‘₯) + 𝑔(π‘₯)) = 𝑑π‘₯ + 𝑑π‘₯ . We now use this result to show a similar claim holds for the sum of three functions. We have 𝐴(π‘₯) 𝑑𝐴 𝑑π‘₯

= = = = = =

𝑓 (π‘₯) + 𝑔(π‘₯) + β„Ž(π‘₯) 𝑑 (𝑓 (π‘₯) + 𝑔(π‘₯) + β„Ž(π‘₯)) 𝑑π‘₯ 𝑑 (𝐡(π‘₯) + β„Ž(π‘₯)) , 𝐡(π‘₯) = 𝑓 (π‘₯) + 𝑔(π‘₯) 𝑑π‘₯ 𝑑𝐡 π‘‘β„Ž + 𝑑π‘₯ 𝑑π‘₯ ) ( 𝑑𝑓 𝑑𝑔 π‘‘β„Ž + + 𝑑π‘₯ 𝑑π‘₯ 𝑑π‘₯ 𝑑𝑓 𝑑𝑔 π‘‘β„Ž + + , 𝑑π‘₯ 𝑑π‘₯ 𝑑π‘₯

where we constantly used the fact that the derivative of a sum of two functions is the sum of the two derivatives.

9

Question: #11: Use Theorem 1 to show that if 𝑓 (π‘₯, 𝑦, 𝑧) is of class 𝐢 3 then βˆ‚3𝑓 βˆ‚3𝑓 = βˆ‚π‘¦βˆ‚π‘§βˆ‚π‘₯ . βˆ‚π‘₯βˆ‚π‘¦βˆ‚π‘§ Hint: Slowly switch orders of differentiation. For example, we know βˆ‚π‘“ = βˆ‚π‘“ , βˆ‚π‘§ βˆ‚π‘§ βˆ‚2𝑓 βˆ‚2𝑓 and so we may differentiate both sides with respect to π‘₯, obtaining βˆ‚π‘₯βˆ‚π‘§ = βˆ‚π‘₯βˆ‚π‘§ , βˆ‚2𝑓 . We now differentiate both and then we may rewrite the right hand side as βˆ‚π‘§βˆ‚π‘₯ sides with respect to 𝑦, and keep switching orders.

3.2 Taylor’s theorem Question: #1: Find the second order Taylor series expansion for 𝑓 (π‘₯, 𝑦) = (π‘₯ + 𝑦)2 about (π‘₯0 , 𝑦0 ) = (0, 0). Solution: We give two solutions. The first is the standard solution. We have ( ) βˆ‚π‘“ βˆ‚π‘“ , = (2(π‘₯ + 𝑦), 2(π‘₯ + 𝑦)) (βˆ‡π‘“ )(π‘₯, 𝑦) = βˆ‚π‘₯ βˆ‚π‘¦ and

(

βˆ‚2𝑓 βˆ‚π‘₯2 βˆ‚2𝑓 βˆ‚π‘¦βˆ‚π‘₯

βˆ‚2𝑓 βˆ‚π‘₯βˆ‚π‘¦ βˆ‚2𝑓 βˆ‚π‘¦ 2

)

2 2 2 2

)

.

1 𝑓 (0, 0) + (βˆ‡π‘“ )(0, 0) β‹… (π‘₯, 𝑦) + (π‘₯, 𝑦)(𝐻𝑓 )(0, 0) 2

(

π‘₯ 𝑦

(𝐻𝑓 )(π‘₯, 𝑦) =

=

(

Thus the second order expansion is )

,

which is 1 0 + (π‘₯𝑦) 2

(

2 2 2 2

)(

π‘₯ 𝑦

)

= = =

( ) 1 2π‘₯ + 2𝑦 (π‘₯𝑦) 2π‘₯ + 2𝑦 2 1 [π‘₯(2π‘₯ + 2𝑦) + 𝑦(2π‘₯ + 2𝑦)] 2 π‘₯(π‘₯ + 𝑦) + 𝑦(π‘₯ + 𝑦) = (π‘₯ + 𝑦)2.

We now present another solution. The Taylor series expansion of 𝑔(𝑒) = 𝑒2 is simply 0 + 0𝑒 + 𝑒2 , and thus taking π‘₯ + 𝑦 for 𝑒 gives the second order Taylor series is just (π‘₯ + 𝑦)2 . It isn’t surprising that this is the answer – 𝑓 is a polynomial of degree 2, and thus its second order Taylor series should equal itself!

10

Question: #4: Find the second order Taylor series expansion for 𝑓 (π‘₯, 𝑦) = 2 2 π‘’βˆ’(π‘₯ +𝑦 ) cos(π‘₯𝑦) about (π‘₯0 , 𝑦0) = (0, 0). Solution: The long way to do this is to compute ( ) βˆ‚π‘“ βˆ‚π‘“ (βˆ‡π‘“ )(π‘₯, 𝑦) = , βˆ‚π‘₯ βˆ‚π‘¦ and (𝐻𝑓 )(π‘₯, 𝑦) =

(

βˆ‚2𝑓 βˆ‚π‘₯βˆ‚π‘¦ βˆ‚2𝑓 βˆ‚π‘¦ 2

βˆ‚2𝑓 βˆ‚π‘₯2 βˆ‚2𝑓 βˆ‚π‘¦βˆ‚π‘₯

)

,

and then use our result that the second order expansion is 1 𝑓 (0, 0) + (βˆ‡π‘“ )(0, 0) β‹… (π‘₯, 𝑦) + (π‘₯, 𝑦)(𝐻𝑓 )(0, 0) 2

(

π‘₯ 𝑦

)

.

This is not pleasant; for instance, βˆ‚π‘“ 2 2 2 2 = βˆ’2π‘₯π‘’βˆ’(π‘₯ +𝑦 ) cos(π‘₯𝑦) βˆ’ π‘¦π‘’βˆ’(π‘₯ +𝑦 ) sin(π‘₯𝑦). βˆ‚π‘₯ There is a faster way. Rolling up our sleeves and doing the work, we find 𝑓 (0, 0) = 1, (βˆ‡π‘“ )(0, 0) = (0, 0) and after even more work we find (𝐻𝑓 )(0, 0) =

(

βˆ’2 0 0 βˆ’2

)

;

we can make our life a little easier by noting that 𝑓 is of class π’ž 2 , and thus βˆ‚2𝑓 . βˆ‚π‘¦βˆ‚π‘₯

Thus we have one fewer painful derivative to take. By Taylor’s theorem, the second order approximation is just ( ) 1 π‘₯ 𝑓 (0, 0) + (βˆ‡π‘“ )(0, 0) β‹… (π‘₯, 𝑦) + (π‘₯, 𝑦)(𝐻𝑓 )(0, 0) . 𝑦 2

Substituting gives

= =

1 1 + (0, 0) β‹… (π‘₯, 𝑦) + (π‘₯, 𝑦) 2 ( ) 1 βˆ’2π‘₯ 1 + (π‘₯, 𝑦) βˆ’2𝑦 2

1 βˆ’ π‘₯2 βˆ’ 𝑦 2 .

11

(

βˆ’2 0 0 βˆ’2

)(

π‘₯ 𝑦

)

βˆ‚2𝑓 βˆ‚π‘₯βˆ‚π‘¦

=

We can determine the Taylor series very easily using our trick. We have 𝑒𝑒 = 1 + 𝑒 +

𝑒2 +β‹…β‹…β‹… , 2

so π‘’βˆ’(π‘₯

2 +𝑦 2 )

= 1 βˆ’ (π‘₯2 + 𝑦 2 ) + β‹… β‹… β‹… ;

we stopped at this term as this term is already of order 2 in π‘₯ and 𝑦, and thus there is no need to keep further terms (as we only want up to second order). Similarly we find 𝑀2 +β‹…β‹…β‹… , cos(𝑀) = 1 βˆ’ 2 so cos(π‘₯𝑦) = 1 βˆ’ β‹… β‹… β‹… ;

here we only kept one term as the next term would be 𝑀 2 /2 = π‘₯2 𝑦 2 /2, which is a fourth order (and not a second order) term. We thus find the Taylor series expansion of order 2 at the origin is simply 1 βˆ’ (π‘₯2 + 𝑦 2), and this was obtained with significantly less work! You of course need to know how to compute a Taylor series in general, but this trick will work in most of the problems you need.

3.3 Extrema of real-valued functions Question: #4: Find the critical points of 𝑓 (π‘₯, 𝑦) = π‘₯2 + 𝑦 2 + 3π‘₯𝑦. β†’ βˆ’ Solution: We must solve βˆ‡π‘“ = 0 . We have ( ) βˆ‚π‘“ βˆ‚π‘“ (βˆ‡π‘“ ) = , = (2π‘₯ + 3𝑦, 2𝑦 + 3π‘₯). βˆ‚π‘₯ βˆ‚π‘¦ Thus an extremum occurs when 2π‘₯ + 3𝑦 = 0, 3π‘₯ + 2𝑦 = 0. There are several ways to proceed. Note, however, that at this point it is no longer a calculus problem, but rather an algebra one. A common approach is to solve 12

for one variable in terms of the other (i.e., the substitution method). Another is to multiply the equations by various constants and combine. Let’s solve for 𝑦 in terms of π‘₯. We have 𝑦 = βˆ’2π‘₯/3 from the first equation and 𝑦 = βˆ’3π‘₯/2 from the second. Thus the only solution is π‘₯ = 𝑦 = 0. Another way of arranging the algebra is to find 𝑦 = βˆ’2π‘₯/3 from the first equation, and then substitute this into the second, which becomes 3π‘₯ + 2(βˆ’2π‘₯/3) = 0, which clearly implies π‘₯ = 0. Alternatively, note 5π‘₯ + 5𝑦 = 0 so π‘₯ = βˆ’π‘¦ and then βˆ’2𝑦 + 3𝑦 = 0 yields 𝑦 = 0. Question: #6: Find the critical points of 𝑓 (π‘₯, 𝑦) = π‘₯2 βˆ’ 3π‘₯𝑦 + 5π‘₯ βˆ’ 2𝑦 + 6𝑦 2 + 8. β†’ βˆ’ Solution: The critical points are where βˆ‡π‘“ = 0 . For our function we have ( ) βˆ‚π‘“ βˆ‚π‘“ (βˆ‡π‘“ )(π‘₯, 𝑦) = , = (2π‘₯ βˆ’ 3𝑦 + 5, βˆ’3π‘₯ βˆ’ 2 + 12𝑦), βˆ‚π‘₯ βˆ‚π‘¦ so in order for this to equal the zero vector we must have 2π‘₯ βˆ’ 3𝑦 + 5 = 0, and βˆ’ 3π‘₯ + 12𝑦 βˆ’ 2 = 0. These are two equations in two unknowns. We have 2π‘₯ βˆ’ 3𝑦 = βˆ’5, βˆ’3π‘₯ + 12𝑦 = 2. There are lots of ways to solve this. We could multiply the first equation by 4 and add it to the second. This will cancel all the 𝑦 terms, and leave us with 8π‘₯ βˆ’ 3π‘₯ = βˆ’20 + 2, or 5π‘₯ = βˆ’18 or π‘₯ = βˆ’18/5. As 𝑦 = 2π‘₯+5 , this implies 3 11 𝑦 = βˆ’ 15 . Another way to solve this system of equations is to isolate 𝑦 as a function of π‘₯ using the first equation, and substitute this into the second. We find 2π‘₯βˆ’3𝑦 = βˆ’5, so 𝑦 = 2π‘₯+5 . Substituting this into the second equation yields 3 βˆ’3π‘₯ βˆ’ 2 + 12

2π‘₯ + 5 = 0, 3

which implies or

βˆ’3π‘₯ βˆ’ 2 + 8π‘₯ + 20 = 0, π‘₯ = βˆ’

exactly as before.

13

18 , 5

3.4 Constrained extrema and Lagrange multipliers Question: #1: Find the extrema of 𝑓 (π‘₯, 𝑦, 𝑧) = π‘₯ βˆ’ 𝑦 + 𝑧 subject to 𝑔(π‘₯, 𝑦, 𝑧) = π‘₯2 + 𝑦 2 + 𝑧 2 = 2. Solution: By the method of Lagrange multipliers, we need (βˆ‡π‘“ )(π‘₯, 𝑦, 𝑧) = πœ†(βˆ‡π‘”)(π‘₯, 𝑦, 𝑧) for (π‘₯, 𝑦, 𝑧) to be an extremum. We have βˆ‡π‘“ = (1, βˆ’1, 1) and βˆ‡π‘” = (2π‘₯, 2𝑦, 2𝑧). Thus we are searching for a πœ† and a point (π‘₯, 𝑦, 𝑧) where (1, βˆ’1, 1) = πœ†(2π‘₯, 2𝑦, 2𝑧). We find 2πœ†π‘₯ = 1, 2πœ†π‘¦ = βˆ’1, 2πœ†π‘§ = 1. As πœ† βˆ•= 0 (if πœ† = 0 then there is no way to have the two gradients equal), we have π‘₯ = 𝑧 = βˆ’π‘¦. We still have another equation to use, namely 𝑔(π‘₯, 𝑦, 𝑧) = 2. There are several ways to proceed. We can solve and find π‘₯ = 𝑧 = 1/2πœ†, 𝑦 = βˆ’1/2πœ†, and thus 1 1 1 + 2 + 2 = 2, 2 4πœ† 4πœ† 4πœ† √ which implies 3/4πœ†2 = 2 or πœ†2 = 3/8, which yields πœ† = Β± 3/8. There are thus two points where 𝑓 may have an extremum, namely √ √ √ √ √ √ (1/2 3/8, βˆ’1/2 3/8, 1/2 3/8), (βˆ’1/2 3/8, 1/2 3/8, βˆ’1/2 3/8). √ Evaluating 𝑓 at the first point gives 1/2 3/8, while evaluating 𝑓 at the second √ point gives βˆ’1/2 3/8. Question: #4: Find the extrema of 𝑓 (π‘₯, 𝑦) = π‘₯ subject to 𝑔(π‘₯, 𝑦) = π‘₯2 +2𝑦 2 = 3. Solution: We have βˆ‡π‘“ = (1, 0) and βˆ‡π‘” = (2π‘₯, 4𝑦). Thus at an extremum the point (π‘₯, 𝑦) must satisfy, for some πœ†, the equation (1, 0) = πœ†(2π‘₯, 4𝑦). This implies 1 = 2πœ†π‘₯ and 0 = 4πœ†π‘¦. We must therefore have 𝑦 = 0, but at this point we cannot determine π‘₯ and πœ†, only their product (which is 1/2). All is not 14

2 lost, however, as we know π‘₯2 + 2𝑦 2 = 3. As 𝑦 = 0, we then √ √ find π‘₯ = 3 so π‘₯ = Β± 3. We could now easily determine πœ† (it is just Β±1/2 3); however, there is no need to. The only reason we care about πœ† is that it is supposed to help us in finding where 𝑓 has an extremum. As we already know the π‘₯ and 𝑦 coordinates, √ we have all the information we need. Thus the extrema occur at π‘₯ = Β± 3. We could have predicted this answer in the beginning. We have our function depending only on π‘₯ and constrained to lie on an ellipse. We thus naturally want the π‘₯-extension as large as possible, which means taking 𝑦 = 0 and being at the extremes of the major-axis.

Question: Find the maximum value of 𝑓 (π‘₯, 𝑦, 𝑧) = π‘₯𝑦𝑧 given that 𝑔(π‘₯, 𝑦, 𝑧) = π‘₯ + 𝑦 + 𝑧 = 3 and π‘₯, 𝑦, 𝑧 β‰₯ 0. Solution: We may interpret this problem as saying we have a bar three units in length, and we can fold it twice at right angles to give a skeleton of part of a box; how should we divide it so that the volume is maximized? While it seems clear that the answer should be π‘₯ = 𝑦 = 𝑧 = 1, we must prove this. The main constraint is 𝑔(π‘₯, 𝑦, 𝑧) = 3; we need the other constraint so as to eliminate possible solutions such as (βˆ’100)(βˆ’100)(203). Using Lagrange multipliers, we want βˆ‡π‘“ = πœ†βˆ‡π‘”. As βˆ‡π‘“ = (𝑦𝑧, π‘₯𝑧, π‘₯𝑦) and βˆ‡π‘” = (1, 1, 1), this means (𝑦𝑧, π‘₯𝑧, π‘₯𝑦) = πœ†(1, 1, 1). If πœ† = 0 then at least one of π‘₯, 𝑦 and 𝑧 equals zero, and the volume π‘₯𝑦𝑧 is zero; thus this clearly cannot be the maximum. We may thus assume πœ† βˆ•= 0. We have 𝑦𝑧 = π‘₯𝑧 = π‘₯𝑦 = πœ†, and we may assume none of π‘₯, 𝑦 or 𝑧 vanish. As 𝑦𝑧 = π‘₯𝑧, since 𝑧 βˆ•= 0 we have 𝑦 = π‘₯. Looking at the other equality yields 𝑦 = 𝑧, and hence π‘₯ = 𝑦 = 𝑧. As 𝑔(π‘₯, 𝑦, 𝑧) = π‘₯ + 𝑦 + 𝑧 = 3, since the three variables are equal we must have each of them equal to 1. More generally, if we have 𝑛 non-negative numbers with a fixed sum, then their product is maximized when they are all equal. The next, more advanced 15

question we could ask is what 𝑛 maximizes the product for a given sum. This question is related to what base we should use in building computers. Interestingly, this implies that if we are primarily concerned with data storage, we should work in base 3 and not base 2. The answer is related to 𝑒, and the fact that 3 is closer to 𝑒 than 2. This was an extra credit problem earlier in the semester; for a non-multivariable calculus solution, see http://www.williams.edu/go/math/sjmiller/public html/ 105/extracredit/ExtraCredit SummandsN.pdf

Question: Maximize the function 𝑓 (π‘₯, 𝑦, 𝑧) = π‘₯𝑦 + 𝑦𝑧 + π‘₯𝑧 on the unit sphere 𝑔(π‘₯, 𝑦, 𝑧) = π‘₯2 + 𝑦 2 + 𝑧 2 = 1. Note this is a hard problem, but looking through the arguments below will give you a great grounding in how to handle the algebra that can arise. Solution: We need βˆ‡π‘“ = πœ†βˆ‡π‘”. Differentiating yields βˆ‡π‘“ = (𝑦 + 𝑧, π‘₯ + 𝑧, π‘₯ + 𝑦) = πœ†(2π‘₯, 2𝑦, 2𝑧) = πœ†βˆ‡π‘”. We thus have four equations in four unknowns: 𝑦+𝑧 π‘₯+𝑧 π‘₯+𝑦 2 2 π‘₯ + 𝑦 + 𝑧2

= = = =

2πœ†π‘₯ 2πœ†π‘¦ 2πœ†π‘§ 1.

There are many ways to solve these equations. We describe a few. First, note that if we take ratios of any two of the first three equations that the πœ† disappears. (Note πœ† cannot equal 0. If it did, we would have π‘₯ + 𝑧 = 0 and π‘₯ + 𝑦 = 0. This would force 𝑦 to equal 𝑧, which when substituted into 𝑦 + 𝑧 = 0 would give 𝑦 = 𝑧 = 0. We would then have π‘₯ = 0, and hence the constraint 𝑔(π‘₯, 𝑦, 𝑧) = 1 𝑦+𝑧 could not be satisfied.) Dividing the first equation by the second gives π‘₯+𝑧 = π‘₯𝑦 . Cross multiplying gives 𝑦 2 + 𝑦𝑧 = π‘₯2 + π‘₯𝑧. Looking at the ratio of the second π‘₯+𝑧 and the third equations gives π‘₯+𝑦 = 𝑦𝑧 , or 𝑧 2 + π‘₯𝑧 = 𝑦 2 + π‘₯𝑦. We thus have π‘₯2 + π‘₯𝑧 = 𝑦 2 + 𝑦𝑧,

𝑧 2 + π‘₯𝑧 = 𝑦 2 + π‘₯𝑦.

Subtracting these two equations from each other gives π‘₯2 βˆ’ 𝑧 2 = 𝑦𝑧 βˆ’ π‘₯𝑦, 16

or (π‘₯ βˆ’ 𝑧)(π‘₯ + 𝑧) = 𝑦(𝑧 βˆ’ π‘₯). There are thus two solutions: either π‘₯ βˆ’ 𝑧 = 0 or 𝑦 = βˆ’(π‘₯ + 𝑧). We leave the rest of this approach to the reader. Another way to attack this problem is to add the first three equations to each other, which gives (𝑦 + 𝑧) + (π‘₯ + 𝑧) + (π‘₯ + 𝑦) = 2πœ†π‘₯ + 2πœ†π‘¦ + 2πœ†π‘§, or equivalently 2(π‘₯ + 𝑦 + 𝑧) = 2πœ†(π‘₯ + 𝑦 + 𝑧). Thus either π‘₯+𝑦 +𝑧 = 0 or πœ† = 1. If πœ† = 1 then squaring the first three equations gives (𝑦 + 𝑧)2 + (π‘₯ + 𝑧)2 + (π‘₯ + 𝑦)2 = 4π‘₯2 + 4𝑦 2 + 4𝑧 2 = 4, where the last follows from the fact that π‘₯2 + 𝑦 2 + 𝑧 2 = 1. If we expand the squares we find 𝑦 2 + 2𝑦𝑧 + 𝑧 2 + π‘₯2 + 2π‘₯𝑧 + 𝑧 2 + π‘₯2 + 2π‘₯𝑦 + 𝑦 2 = 4. Note the left hand side has 2(π‘₯2 + 𝑦 2 + 𝑧 2 ), which is 2. Thus we have 2 + 2𝑦𝑧 + 2π‘₯𝑧 + 2π‘₯𝑦 = 4, or 𝑦𝑧 + π‘₯𝑧 + 𝑦𝑧 = 1. Note, however, that 𝑦𝑧 + π‘₯𝑧 + 𝑦𝑧 is just our function 𝑓 (π‘₯, 𝑦, 𝑧)! We leave the rest of the details of this problem to the reader.

4 Vector-valued functions 5 Double and Triple Integrals 5.1 Introduction Question: #1b: Find

∫ πœ‹/2 ∫ 1 0

0

(𝑦 cos π‘₯ + 2)𝑑𝑦𝑑π‘₯. 17

Solution: We first do the 𝑦-integral, and then the π‘₯-integral. We have ] ∫ πœ‹/2 ∫ 1 ∫ πœ‹/2 [∫ 1 (𝑦 cos π‘₯ + 2)𝑑𝑦𝑑π‘₯ = (𝑦 cos π‘₯ + 2)𝑑𝑦 𝑑π‘₯ 0 0 0 0 ⎑ 1 ⎀ 1 ∫ πœ‹/2 2 𝑦 ⎣ = cos π‘₯ + 2𝑦 ⎦ 𝑑π‘₯ 2 0 0 0 ∫ πœ‹/2 ( ) cos π‘₯ = + 2 𝑑π‘₯ 2 0 ] πœ‹/2 [ sin π‘₯ = + 2π‘₯ 2 0

=

1 + πœ‹. 2

∫0 ∫2 Question: #1d: Find βˆ’1 1 (βˆ’π‘₯ ln 𝑦)𝑑𝑦𝑑π‘₯. Solution: Again, we do the 𝑦-integral first, followed by the π‘₯-integral. We need to find a function whose derivative is ln 𝑦. It is natural (forgive the pun) to try 𝑦 ln 𝑦. Why is this a reasonable guess? When we take the derivative, we use the product rule and the first piece is just 1 β‹… ln 𝑦. Thus this is close to what we want, though not quite the correct answer. The problem is the full derivative is 1 β‹… ln 𝑦 + 𝑦 β‹…

1 = ln 𝑦 + 1; 𝑦

again, this is almost correct, but we are off by 1. We may interpret this as saying our guess is off by a function whose derivative is 1; one example of such a function is 𝑦. If we subtract this from our original guess, we should end up with the correct anti-derivative. Specifically, (𝑦 ln 𝑦 βˆ’ 𝑦)β€² = 1 β‹… ln 𝑦 + 𝑦 β‹…

1 βˆ’ 1 = ln 𝑦; 𝑦

we have thus found the sought-after anti-derivative. This is the Method of Guess and Check, and it is a powerful way to find anti-derivatives.

18

Armed with the anti-derivative for ln 𝑦, we can solve the problem. We have ] ∫ 0∫ 2 ∫ 0 [∫ 2 (βˆ’π‘₯ ln 𝑦)𝑑𝑦𝑑π‘₯ = (βˆ’π‘₯ ln 𝑦)𝑑𝑦 𝑑π‘₯ βˆ’1 1 βˆ’1 1 ] ∫ 0 [∫ 2 = βˆ’ π‘₯ ln 𝑦𝑑𝑦 𝑑π‘₯ βˆ’1 1 2 ∫ 0 = βˆ’ π‘₯ [𝑦 ln 𝑦 βˆ’ 𝑦] 𝑑π‘₯ βˆ’1 1 ∫ 0 = βˆ’ π‘₯ [(2 ln 2 βˆ’ 2) βˆ’ (1 ln 1 βˆ’ 1)] 𝑑π‘₯ βˆ’1 ∫ 0 = βˆ’ π‘₯(2 ln 2 βˆ’ 1)𝑑π‘₯ βˆ’1 ∫ 0 = βˆ’(2 ln 2 βˆ’ 1) π‘₯𝑑π‘₯ βˆ’1 ⎑ ⎀ 0 2 π‘₯ = βˆ’(2 ln 2 βˆ’ 1) ⎣ ⎦ 2 βˆ’1 [ ] 1 = βˆ’(2 ln 2 βˆ’ 1) βˆ’ 2 2 ln 2 βˆ’ 1 . = 2

5.2 The Double Integral over a Rectangle ∫∫ Question: #1c: Evaluate (π‘₯𝑦)2 cos(π‘₯3 )𝑑𝐴, where 𝑅 = [0, 1] Γ— [0, 1]. 𝑅 Solution: We need to choose whether or not we want to integrate first with respect to π‘₯ or with respect to 𝑦. For this problem, it does not matter as we can write the integral as 𝑓 (π‘₯)𝑔(𝑦) for some functions 𝑓 and 𝑔 (here 𝑓 (π‘₯) = π‘₯2 cos(π‘₯3 )

19

and 𝑔(𝑦) = 𝑦 2 ). Let’s do the integration with respect to 𝑦 first. We have ] ∫ ∫ ∫ 1 [∫ 1 2 3 2 2 3 (π‘₯𝑦) cos(π‘₯ )𝑑𝐴 = π‘₯ 𝑦 cos(π‘₯ )𝑑𝑦 𝑑π‘₯ 𝑅 0 0 [∫ 1 ] ∫ 1 3 2 2 = cos(π‘₯ )π‘₯ 𝑦 𝑑𝑦 𝑑π‘₯ 0 0 ⎑ ⎀ 1 ∫ 1 3 𝑦 = cos(π‘₯3 )π‘₯2 ⎣ ⎦ 𝑑π‘₯ 3 0 0 ∫ 1 1 = cos(π‘₯3 )π‘₯2 𝑑π‘₯ 3 0 ∫ 1 1 cos(π‘₯3 )3π‘₯2 𝑑π‘₯; 9 0 where we multiplied by 1 in the form 3/3 to facilitate the application of 𝑒-substitution below (though of course this is not needed). Let 𝑒 = π‘₯3 . Then 𝑑𝑒 = 3π‘₯2 𝑑π‘₯, and as π‘₯ : 0 β†’ 1 we have 𝑒 : 0 β†’ 1. (Note it is very important that our function 𝑒 = π‘₯3 is monotonic or strictly increasing in this domain). Thus we have ∫ ∫ ∫ 1 1 2 3 (π‘₯𝑦) cos(π‘₯ )𝑑𝐴 = cos 𝑒𝑑𝑒 9 0 𝑅 1 sin 𝑒 9 0

=

sin 1 . 9

∫1∫1 Question: Compute 0 0 𝑦 cos(π‘₯𝑦)𝑑𝐴. Solution: We have a choice as to whether or not we want to integrate with respect to π‘₯ first or with respect to 𝑦. Note the integrand is 𝑦 cos(π‘₯𝑦). If we integrate with respect to π‘₯ first, then everything will work out nicely through 𝑒substitution; if we do the 𝑦 integral first we have to use the method of Guess and Check to figure out an anti-derivative (with respect to 𝑦) of 𝑦 cos(π‘₯𝑦). Thus let’s

20

integrate with respect to π‘₯ first. We have ] ∫ 1∫ 1 ∫ 1 [∫ 1 𝑦 cos(π‘₯𝑦)𝑑𝐴 = 𝑦 cos(π‘₯𝑦)𝑑π‘₯ 𝑑𝑦 0 0 0 0 ] ∫ 1 [∫ 1 = cos(π‘₯𝑦)𝑦𝑑π‘₯ 𝑑𝑦. 0

0

Let 𝑒 = π‘₯𝑦, so 𝑑𝑒 = 𝑦𝑑π‘₯ and π‘₯ : 0 β†’ 1 means 𝑒 : 0 β†’ 𝑦. We find ] ∫ 1∫ 1 ∫ 1 [∫ 𝑦 𝑦 cos(π‘₯𝑦)𝑑𝐴 = cos 𝑒𝑑𝑒 𝑑𝑦 0 0 0 0 𝑦 ] ∫ 1[ = sin 𝑒 𝑑𝑦 0 0 ∫ 1 = sin 𝑦 0 1 = [βˆ’ cos 𝑦] 0

= =

(βˆ’ cos 1) βˆ’ (βˆ’1) 1 βˆ’ cos 1.

5.3 The Double and Triple Integral Over More General Regions Question: #1a: Evaluate the iterated integral ∫

0

1

∫

π‘₯2

𝑑𝑦𝑑π‘₯,

0

state whether or not the region is π‘₯-simple, 𝑦-simple or simple. Draw the region. Solution: Solution: The region is drawn in Figure 1. The region is 𝑦-simple, as for 0 ≀ 𝑦 ≀ 1 we have πœ™1 (𝑦) ≀ π‘₯ ≀ πœ™2 (𝑦) with πœ™1 (𝑦) = 0 and πœ™2 (𝑦) = π‘₯2 . Similarly we see the region is π‘₯-simple. For √ √ 0 ≀ 𝑦 ≀ 1 we have 𝑦 ≀ π‘₯ ≀ 1; we take πœ“1 (𝑦) = 𝑦 and πœ“2 (𝑦) = 1. As the region is both π‘₯-simple and 𝑦-simple, it is simple.

21

1.0

0.8

0.6

0.4

0.2

0.0 0.0

0.2

0.4

0.6

0.8

1.0

Figure 1: Region corresponding to 0 ≀ π‘₯ ≀ 1 and 0 ≀ 𝑦 ≀ π‘₯2 . We now evaluate the integral. We have ∫

0

1

∫

π‘₯2

𝑑𝑦𝑑π‘₯

=

0

∫

1

∫

1

∫

1

0

=

π‘₯2

]

1𝑑𝑦 𝑑π‘₯

0

0

=

[∫

⎑ 2⎀ π‘₯ βŽ£π‘¦ ⎦ 𝑑π‘₯ 0

π‘₯2 𝑑π‘₯

0

1 π‘₯3 . 3

=

0

Question: #1a: Evaluate the iterated integral ∫

2 βˆ’3

∫

𝑦2

(π‘₯2 + 𝑦)𝑑π‘₯𝑑𝑦, 0

state whether or not the region is π‘₯-simple, 𝑦-simple or simple. Draw the region. Solution: The region is drawn in Figure 2. The region is clearly π‘₯-simple, as for βˆ’3 ≀ 𝑦 ≀ 2 we have πœ“1 (𝑦) ≀ π‘₯ ≀ πœ“2 (𝑦), where πœ“1 (𝑦) = 0 and πœ“2 (𝑦) = 𝑦 2 (and of course πœ“1 (𝑦) ≀ πœ“2 (𝑦). The region is not 𝑦-simple (and hence it is not simple). The reason it is not 𝑦-simple 22

2

1

0

-1

-2

-3 0

2

4

6

8

Figure 2: Region corresponding to βˆ’3 ≀ 𝑦 ≀ 2 and 0 ≀ π‘₯ ≀ 𝑦 2 . is that for π‘₯ ∈ [0, √ 1] the values √ of 𝑦 vary discontinuously. For example, if π‘₯ = 2 then βˆ’3 ≀ 𝑦 ≀ βˆ’ 2 and 2 ≀ 𝑦 ≀ 2. We now evaluate the integral. We have ] ∫ 2 ∫ 𝑦2 ∫ 2 [∫ 𝑦 2 (π‘₯2 + 𝑦)𝑑π‘₯𝑑𝑦 = (π‘₯2 + 𝑦)𝑑π‘₯ 𝑑𝑦 βˆ’3

0

0

βˆ’3

= = =

𝑦2 ⎀ ⎣ π‘₯ + π‘₯𝑦 ⎦ 𝑑𝑦 3 βˆ’3 0 0 ) ∫ 2( 6 𝑦 + 𝑦 3 𝑑𝑦 3 βˆ’3 2 2 𝑦 7 𝑦 4 + 21 4 ∫

2

𝑦2 3

⎑

βˆ’3

= =

βˆ’3

128 729 81 + +4βˆ’ 21 7 4 7895 . 84

5.4 Changing the order of integration Question: #1a: Sketch the region and evaluate 23

∫1∫1 0

π‘₯

π‘₯𝑦𝑑𝑦𝑑π‘₯ both ways.

1.0

0.8

0.6

0.4

0.2

0.0 0.0

0.2

0.4

0.6

0.8

1.0

Figure 3: Region corresponding to π‘₯ ≀ 𝑦 ≀ 1 and 0 ≀ π‘₯ ≀ 1. Solution: See Figure 3 for a sketch of the region. We have ] ∫ 1∫ 1 ∫ 1 [∫ 1 π‘₯𝑦𝑑𝑦𝑑π‘₯ = π‘₯ 𝑦𝑑𝑦 𝑑π‘₯ 0 π‘₯ 0 π‘₯ ⎑ ⎀ 1 ∫ 1 2 𝑦 = π‘₯ ⎣ ⎦ 𝑑π‘₯ 2 0 π‘₯ ] ∫ 1 [ 1 π‘₯2 = π‘₯ βˆ’ 𝑑π‘₯ 2 2 0 ) ∫ 1( π‘₯ π‘₯3 βˆ’ 𝑑π‘₯ = 2 2 0 1 1 π‘₯2 π‘₯4 = βˆ’ 4 8 0

=

0

1 1 1 βˆ’ = . 4 8 8

We now do the integration in the opposite order. We fix 𝑦 now, and 𝑦 varies

24

from 0 to 1. It is now π‘₯ that varies, and π‘₯ ranges from 0 to 𝑦. We thus find ∫ 1∫ 1 ∫ 1∫ 𝑦 π‘₯𝑦𝑑𝑦𝑑π‘₯ = π‘₯𝑦𝑑π‘₯𝑑𝑦 0 π‘₯ 0 0 ] ∫ 1 [∫ 𝑦 = 𝑦 π‘₯𝑑π‘₯ 𝑑𝑦 0 0 ∫ 1 [ 2 𝑦 ] π‘₯ = 𝑦 𝑑𝑦 2 0 0 ∫ 1 ( 2) 𝑦 = 𝑦 𝑑𝑦 2 0 1 𝑦 4 1 = = . 8 8 0

Note that the two orders of integration lead to the same answer for this problem.

5.5 Mathematical Modeling: Baseball / Sabermetrics Lecture The following are some problems related to the ones from the baseball lecture. The slides are online at http://www.williams.edu/go/math/sjmiller/public html /105/talks/PythagWLTalk GeneralCalcVersion.pdf Question: #1: Let 𝑓 (π‘₯) = 6π‘₯(1 βˆ’ π‘₯) for 0 ≀ π‘₯ ≀ 1 and 0 otherwise, and let 𝑔(𝑦) = 12𝑦 2(1 βˆ’ 𝑦) for 0 ≀ 𝑦 ≀ 1 and zero otherwise. Prove 𝑓 and 𝑔 are probability distributions. Let 𝑋 be a random variable whose probability density of taking on the value π‘₯ is 𝑓 (π‘₯), and let π‘Œ be a random variable whose probability density of taking on the value 𝑦 is 𝑔(𝑦). Compute the probability that 𝑋 > π‘Œ (assuming, of course, that 𝑋 and π‘Œ are independent). Solution: To prove that 𝑓 and 𝑔 are probability distributions, we must show that each is non-negative and integrates to 1. Both are clearly non-negative; we are left with showing each integrates to one. The fastest way to do this is to note

25

that it suffices to study the integral from 0 to 1 of π‘₯π‘˜ (1 βˆ’ π‘₯). We have ∫ 1 ∫ 1 [ π‘˜ ] π‘˜ π‘₯ (1 βˆ’ π‘₯) = π‘₯ βˆ’ π‘₯π‘˜+1 𝑑π‘₯ 0 0 1 1 π‘₯π‘˜+2 π‘₯π‘˜+1 = βˆ’ π‘˜ + 1 π‘˜ + 2 0

= =

Thus

0

1 1 βˆ’ π‘˜+1 π‘˜+2 1 . (π‘˜ + 1)(π‘˜ + 2)

{ (π‘˜ + 1)(π‘˜ + 2)π‘₯π‘˜ (1 βˆ’ π‘₯) β„Ž(π‘₯) = 0

if 0 ≀ π‘₯ ≀ 1 otherwise

is a probability distribution for any positive integer π‘˜. In particular, if we take π‘˜ = 1 we see that we should have 6π‘₯(1 βˆ’ π‘₯), while if π‘˜ = 2 we should have 12π‘₯2 (1 βˆ’ π‘₯), which we do. This thus verifies both distributions are probability distributions simultaneously. We now compute the probability that 𝑋 > π‘Œ . We are integrating over the triangle 0 ≀ 𝑦 ≀ π‘₯ ≀ 1, and have ∫ 1 ∫ π‘₯ Prob(𝑋 β‰₯ π‘Œ ) = 𝑓 (π‘₯)𝑔(𝑦)𝑑𝑦𝑑π‘₯ π‘₯=0 𝑦=0 [∫ π‘₯ ] ∫ 1 𝑔(𝑦)𝑑𝑦 𝑑π‘₯ = 𝑓 (π‘₯) π‘₯=0 𝑦=0 [∫ π‘₯ ] ∫ 1 2 = 6π‘₯(1 βˆ’ π‘₯) 12𝑦 (1 βˆ’ 𝑦)𝑑𝑦 𝑑π‘₯ π‘₯=0 𝑦=0 [∫ π‘₯ ] ∫ 1 2 3 = 6π‘₯(1 βˆ’ π‘₯)12 (𝑦 βˆ’ 𝑦 )𝑑𝑦 𝑑π‘₯ π‘₯=0 𝑦=0 π‘₯ ] [ π‘₯ ∫ 1 3 4 𝑦 𝑦 = 72(π‘₯ βˆ’ π‘₯2 ) βˆ’ 𝑑π‘₯ 3 4 π‘₯=0 [ 3 0 4] 0 ∫ 1 π‘₯ π‘₯ = 72(π‘₯ βˆ’ π‘₯2 ) βˆ’ 𝑑π‘₯. 3 4 π‘₯=0

26

The simplest way to evaluate this is to expand, and we find ] ∫ 1[ 6 π‘₯ 7π‘₯5 π‘₯4 Prob(𝑋 β‰₯ π‘Œ ) = 72 βˆ’ + 𝑑π‘₯ 4 12 3 0 1 1 1 7 6 5 7π‘₯ π‘₯ π‘₯ = βˆ’ + 28 72 15 0

=

0

0

13 . 2520

6 Change of variable formula and applications of integration 6.1 6.2 The Change of Variable Theorem Question: #1: Consider the change of variables 𝑒 = 2π‘₯ + 3𝑦 and 𝑣 = 4𝑦. Show that this map takes the unit square [0, 1] Γ— [0, 1] (i.e., the set of points (π‘₯, 𝑦) with 0 ≀ π‘₯, 𝑦 ≀ 1) to a parallelogram. Use the change of variables formula to find the area of the parallelogram. Solution: The unit square is mapped to the parallelogram shown in Figure 4. To see this, look and see where each vertex of the unit square is sent. We have (0, 0) goes to (0, 0), we have (1, 0) goes to (2, 0), (0, 1) goes to (3, 4) and finally (1, 1) goes to (5, 4). More generally, if we take a point of the form (π‘₯, 0) it is mapped to the point (2π‘₯, 0), so we see the interval [0, 1] on the π‘₯-axis is mapped to the interval [0, 2] in the 𝑒-axis. A similar analysis shows all the other lines of the unit square are mapped to lines in the 𝑒𝑣-plane. For example, consider the line (π‘₯, 1) with 0 ≀ π‘₯ ≀ 1. This is mapped to the line (2π‘₯ + 3, 4) in the 𝑒𝑣-plane, or equivalently the line from (3, 4) (corresponding to π‘₯ = 0) to the point (5, 4) (corresponding to π‘₯ = 1). We need the inverse transformation 𝑇 βˆ’1 , which gives us the π‘₯ and 𝑦 corresponding to a choice of 𝑒 and 𝑣. We have to invert the relations 𝑒 = 2π‘₯ + 3𝑦, 𝑣 = 4𝑦. The second is the easiest; we clearly need to have 𝑦 = 𝑣/4. Knowing this, we then find 3𝑣 𝑒 = 2π‘₯ + , 4 27

Figure 4: Region the unit square is mapped to under 𝑇 (π‘₯, 𝑦) = (2π‘₯ + 3𝑦, 4𝑦).

or π‘₯ =

𝑒 3𝑣 βˆ’ . 2 4

In other words, we have 𝑇

βˆ’1

(𝑒, 𝑣) = (π‘₯(𝑒, 𝑣), 𝑦(𝑒, 𝑣)) =

(

𝑒 3𝑣 𝑣 βˆ’ , 2 4 4

)

.

We now find the determinant of the derivative. First we compute ( βˆ‚π‘₯ βˆ‚π‘₯ ) ( 1 ) 3 βˆ’ βˆ’1 βˆ‚π‘’ βˆ‚π‘£ 2 4 (𝐷𝑇 )(𝑒, 𝑣) = = . βˆ‚π‘¦ βˆ‚π‘¦ 0 14 βˆ‚π‘’ βˆ‚π‘£ The determinant is det((𝐷𝑇 βˆ’1 )(𝑒, 𝑣)) =

1 1 3 1 β‹… + β‹…0 = , 2 4 4 8

and thus the absolute value of the determinant is det((𝐷𝑇 βˆ’1 )(𝑒, 𝑣)) = 1 , 8 which means

𝑑π‘₯𝑑𝑦

βˆ’β†’

det((𝐷𝑇 βˆ’1)(𝑒, 𝑣)) 𝑑𝑒𝑑𝑣 = 1 𝑑𝑒𝑑𝑣. 8 28

By the Change of Variables formula, if 𝑆 is the original unit square in π‘₯𝑦-space and 𝑃 = 𝑇 (𝑆) is the parallelogram in 𝑒𝑣-space, we have ∫ ∫ ∫ ∫ ∫ ∫ 1 βˆ’1 1𝑑π‘₯𝑑𝑦 = 1 det((𝐷𝑇 )(𝑒, 𝑣)) 𝑑𝑒𝑑𝑣 = 1 β‹… 𝑑𝑒𝑑𝑣. 8 𝑆 𝑇 (𝑆) 𝑇 (𝑆) As 1/8 is constant, we can pull it out of the integral and find ∫ ∫ ∫ ∫ 1 1𝑑π‘₯𝑑𝑦 = 1𝑑𝑒𝑑𝑣; 8 𝑆 𝑇 (𝑆)

the left double integral is the area of the unit square, while the right double integral is the area of our parallelogrma. We thus find Area(𝑆) =

1 1 Area(𝑇 (𝑆)) = Area(𝑃 ), 8 8

or equivalently that the area of the parallelogram is 8: Area(𝑃 ) = 8Area(𝑆) = 8 β‹… 1 = 8. We could consider more general maps from squares to parallelograms, but this illustrates the principle and proves a nice, known result: the area of a parallelogram is base times height. For our parallelogram, the has has length 2 and the height is 4, which do multiply to give an area of 8. Notice that we are able to deduce the formula for the parallelogram’s area by knowing the area of the square because the absolute value of the determinant of the derivative matrix is constant (i.e., independent of 𝑒 and 𝑣). This allows us to pull out that common factor of 1/8 and leaves us with the integral of 1 over the parallelogram, which is thus its area. Whenever we have a change of variables where the determinant is constant, these calculations can often allow us to deduce the area of one region from knowing another. This is true in the homework problem, where you are asked to find the area of an ellipse knowing the area of another region. For that problem, consider the ellipse ( π‘₯ )2 ( 𝑦 )2 + ≀ 1. π‘Ž 𝑏 Consider the change of variables 𝑒 = π‘₯/π‘Ž and 𝑣 = 𝑦/𝑏, so

𝑇 (π‘₯, 𝑦) = (𝑒(π‘₯, 𝑦), 𝑣(π‘₯, 𝑦)) = (π‘₯/π‘Ž, 𝑦/𝑏) 29

or equivalently the inverse map 𝑇 βˆ’1 would be 𝑇 βˆ’1 (𝑒, 𝑣) = (π‘₯(𝑒, 𝑣), 𝑦(𝑒, 𝑣)) = (π‘Žπ‘’, 𝑏𝑣). Note this maps the ellipse to the unit disk 𝑒2 + 𝑣 2 ≀ 1, and we know the area of the unit disk is just πœ‹12 = πœ‹!

Question: #1: This is a slight modification of Problem #1 from Section 6.2: Let 𝐷 be the unit disk π‘₯2 + 𝑦 2 ≀ 1. Consider the integral ∫ ∫ cos(π‘₯2 + 𝑦 2)𝑑π‘₯𝑑𝑦. 𝐷

Evaluate this using polar coordinates. Solution: We have 𝑑π‘₯𝑑𝑦 goes to π‘Ÿπ‘‘π‘Ÿπ‘‘πœƒ, and the unit disk becomes 0 ≀ π‘Ÿ ≀ 1 and 0 ≀ πœƒ ≀ 2πœ‹. We replace 𝑓 (π‘₯, 𝑦) with 𝑓 (π‘Ÿ cos πœƒ, π‘Ÿ sin πœƒ), and thus find ∫ 2πœ‹ ∫ 1 ∫ ∫ 2 2 cos(π‘₯ + 𝑦 )𝑑π‘₯𝑑𝑦 = cos(π‘Ÿ 2 cos2 πœƒ + π‘Ÿ 2 sin2 πœƒ)π‘Ÿπ‘‘π‘Ÿπ‘‘πœƒ 𝐷 π‘Ÿ=0 βˆ«πœƒ=0 2πœ‹ ∫ 1 = cos(π‘Ÿ 2 )π‘Ÿπ‘‘π‘Ÿπ‘‘πœƒ πœƒ=0 π‘Ÿ=0 ] ∫ 2πœ‹ [∫ 1 1 2 = cos(π‘Ÿ )2π‘Ÿπ‘‘π‘Ÿ π‘‘πœƒ πœƒ=0 2 π‘Ÿ=0 ⎑ 1 ⎀ ∫ 2πœ‹ 1⎣ sin(π‘Ÿ 2 ) ⎦ π‘‘πœƒ = 2 πœƒ=0 0 ∫ 2πœ‹ sin 1 = π‘‘πœƒ πœƒ=0 2 sin 1 = β‹… 2πœ‹ = πœ‹ sin 1. 2

Question: #13: This is a problem similar to Problem #13 from Section 6.2. Consider the cylinder 𝐢 given by π‘₯2 + 𝑦 2 ≀ 9 and βˆ’1 ≀ 𝑧 ≀ 2. Evaluate ∫ ∫ ∫ 𝑓 (π‘₯, 𝑦, 𝑧)𝑑π‘₯𝑑𝑦𝑑𝑧 𝐢

30

where 𝑓 (π‘₯, 𝑦, 𝑧) = 𝑧

√

π‘₯2 + 𝑦 2.

Solution: If we wanted to write down the integral explicitly in Cartesian coordinates, we would have ∫ 2 ∫ 3 ∫ √9βˆ’π‘¦2 √ 𝑧 π‘₯2 + 𝑦 2𝑑π‘₯𝑑𝑦𝑑𝑧. √ 𝑧=βˆ’1

𝑦=βˆ’3

π‘₯=βˆ’

9βˆ’π‘¦ 2

2 2 To see this, note that on the √boundary π‘₯ √ + 𝑦 = 9, so if we have chosen a value of 𝑦 then π‘₯ ranges from βˆ’ 9 βˆ’ 𝑦 2 to 9 βˆ’ 𝑦 2 ; these are not integrals we desire to evaluate! For cylindrical coordinates, we have

𝑑π‘₯𝑑𝑦𝑑𝑧

βˆ’β†’

π‘Ÿπ‘‘π‘Ÿπ‘‘πœƒπ‘‘π‘§,

and π‘₯ = π‘Ÿ cos πœƒ, 𝑦 = π‘Ÿ sin πœƒ, 𝑧 = 𝑧. Our function 𝑓 (π‘₯, 𝑦, 𝑧) becomes 𝑓 (π‘Ÿ cos πœƒ, π‘Ÿ sin πœƒ, 𝑧), or in our case √ √ 𝑧 π‘₯2 + 𝑦 2 βˆ’β†’ 𝑧 π‘Ÿ 2 cos2 πœƒ + π‘Ÿ 2 sin2 πœƒ = π‘§π‘Ÿ.

The bounds of integration are 𝑧 ranges from βˆ’1 to 2, πœƒ ranges from 0 to 2πœ‹, and π‘Ÿ

31

ranges from 0 to 3. We thus have ∫ ∫ ∫ ∫ 2 ∫ 2πœ‹ ∫ 3 𝑓 (π‘₯, 𝑦, 𝑧)𝑑π‘₯𝑑𝑦𝑑𝑧 = 𝑓 (π‘Ÿ cos πœƒ, π‘Ÿ sin πœƒ, 𝑧)π‘Ÿπ‘‘π‘Ÿπ‘‘πœƒπ‘‘π‘§ 𝐢 𝑧=βˆ’1 πœƒ=0 π‘Ÿ=0 ∫ 2 ∫ 2πœ‹ ∫ 3 √ = 𝑧 π‘Ÿ 2 cos2 πœƒ + π‘Ÿ 2 sin2 πœƒπ‘Ÿπ‘‘π‘Ÿπ‘‘πœƒπ‘‘π‘§ 𝑧=βˆ’1 πœƒ=0 π‘Ÿ=0 ∫ 2 ∫ 2πœ‹ ∫ 3 = π‘§π‘Ÿ β‹… π‘Ÿπ‘‘π‘Ÿπ‘‘πœƒπ‘‘π‘§ 𝑧=βˆ’1 πœƒ=0 π‘Ÿ=0 ] ∫ 2 ∫ 2πœ‹ [∫ 3 2 = 𝑧 π‘Ÿ π‘‘π‘Ÿ π‘‘πœƒπ‘‘π‘§ 𝑧=βˆ’1 πœƒ=0 π‘Ÿ=0 ⎑ ⎀ ∫ 2 ∫ 2πœ‹ 3 3 ⎣ π‘Ÿ ⎦ π‘‘πœƒπ‘‘π‘§ = 𝑧 3 𝑧=βˆ’1 πœƒ=0 0 ∫ 2 ∫ 2πœ‹ 27 = 𝑧 π‘‘πœƒπ‘‘π‘§ 𝑧=βˆ’1 πœƒ=0 3 [∫ 2πœ‹ ] ∫ 2 = 9 𝑧 π‘‘πœƒ 𝑑𝑧 𝑧=βˆ’1 πœƒ=0 ∫ 2 = 9 𝑧2πœ‹π‘‘π‘§ 𝑧=βˆ’1 ∫ 2 = 18πœ‹ 𝑧𝑑𝑧 𝑧=βˆ’1 2 2 𝑧 = 18πœ‹ 2 [ βˆ’1 ] 4 1 βˆ’ = 18πœ‹ 2 2 3 = 18πœ‹ β‹… 2 = 27πœ‹.

Question: #21: This is a problem similar to Problem #21 from Section 6.2. Consider the unit sphere 𝑆 given by π‘₯2 + 𝑦 2 + 𝑧 2 ≀ 1. Evaluate ∫ ∫ ∫ 𝑓 (π‘₯, 𝑦, 𝑧)𝑑π‘₯𝑑𝑦𝑑𝑧 𝑆

32

for 𝑓 (π‘₯, 𝑦, 𝑧) =

(π‘₯2

1 . + 𝑦2 + 𝑧2)

If we were to write the integral out explicitly in Cartesian coordinates, we would find it equals ∫ ∫ √ ∫ √ 1βˆ’π‘§ 2

1

𝑧=βˆ’1

√ 𝑦=βˆ’ 1βˆ’π‘§ 2

1βˆ’π‘¦ 2 βˆ’π‘§ 2

π‘₯=βˆ’

√

𝑓 (π‘₯, 𝑦, 𝑧)𝑑π‘₯𝑑𝑦𝑑𝑧,

1βˆ’π‘¦ 2 βˆ’π‘§ 2

and these bounds of integration should look horrible! We now convert to spherical coordinates. We have π‘₯ = 𝜌 sin πœ™ cos πœƒ, 𝑦 = 𝜌 sin πœ™ sin πœƒ, 𝑧 = 𝜌 cos πœ™, with 0 ≀ 𝜌 ≀ 1, 0 ≀ πœƒ ≀ 2πœ‹, 0 ≀ πœ™ ≀ πœ‹. Our function 𝑓 (π‘₯, 𝑦, 𝑧) becomes 𝑓 (𝜌 sin πœ™ cos πœƒ, 𝜌 sin πœ™ sin πœƒ, 𝜌 cos πœ™) =

1 𝜌2

after some simple algebra. Finally, 𝑑π‘₯𝑑𝑦𝑑𝑧

βˆ’β†’

𝜌2 sin πœ™π‘‘πœŒπ‘‘πœƒπ‘‘πœ™.

Note: other textbooks change the role of πœƒ and πœ™, especially physics books. We

33

thus have ∫ ∫ ∫

= = = = = = =

∫

∫

∫

πœ‹

πœ™=0 πœ‹ πœ™=0 πœ‹

𝑓 (π‘₯, 𝑦, 𝑧)𝑑π‘₯𝑑𝑦𝑑𝑧 ∫ ∫ 1 𝑓 (𝜌 sin πœ™ cos πœƒ, 𝜌 sin πœ™ sin πœƒ, 𝜌 cos πœ™)𝜌2 sin πœ™π‘‘πœŒπ‘‘πœƒπ‘‘πœ™ 𝑆 2πœ‹

πœƒ=0 2πœ‹

∫

πœƒ=0 2πœ‹

∫

𝜌=0 1

∫

𝜌=0 1

1 2 𝜌 sin πœ™π‘‘πœŒπ‘‘πœƒπ‘‘πœ™ 𝜌2

∫

sin πœ™π‘‘πœŒπ‘‘πœƒπ‘‘πœ™ [∫ 1 ] sin πœ™ π‘‘πœŒ π‘‘πœƒπ‘‘πœ™ πœ™=0 πœƒ=0 𝜌=0 [∫ 2πœ‹ ] ∫ πœ‹ sin πœ™ π‘‘πœƒ π‘‘πœ™ πœ™=0 πœƒ=0 ∫ πœ‹ 2πœ‹ sin πœ™π‘‘πœ™ πœ™=0 πœ‹ ] [ 2πœ‹ βˆ’ cos πœ™

πœ™=0 ∫ πœ‹

πœƒ=0 ∫ 2πœ‹

𝜌=0

0

= 2πœ‹ [(βˆ’ cos πœ‹) βˆ’ (βˆ’ cos 0)] = 2πœ‹ (1 + 1) = 4πœ‹.

7 Sequences and Series 7.1 Page 10.3 Question: #2: Give an example of a sequence that does not have a limit. Solution: Let π‘Žπ‘› = (βˆ’1)𝑛 . Clearly this sequence does not have a limit, as half the time the sequence is 1 and half the time the sequence is -1. For another example, consider π‘Žπ‘› = 𝑛!, which clearly grows without bound.

34

Question: #4: Compute the limit of the sequence π‘Žπ‘› = 3/𝑛2 or explain why it does not converge. Solution: Note limπ‘›β†’βˆž 3 = 3 and limπ‘›β†’βˆž 𝑛2 = ∞. Technically we cannot use the limit of a quotient is the quotient of the limit as the denominator tends to infinity and thus doesn’t converge; however, as the numerator is bounded (it is in fact always 3) and the denominator becomes arbitrarily large, we can see that the sequence does converge to 0. For example, if 𝑛 β‰₯ 55 then 3/𝑛2 ≀ 1/1000, if 𝑛 β‰₯ 174 then 3/𝑛2 ≀ 1/10000, and if 𝑛 β‰₯ 548 then 3/𝑛2 ≀ 1/100000. 3

2

βˆ’π‘›βˆ’2 Question: Similar Problem to #5: Find the limit of π‘Žπ‘› = 𝑛 3𝑛+2𝑛 3 +π‘›βˆ’11 , or prove it does not exist. Solution: There are several ways to do this. We cannot use the limit of a quotient is the quotient of the limits, as both the numerator and denominator tend to infinity as 𝑛 β†’ ∞. One approach is to pull out the largest power of 𝑛 in the numerator and denominator: ) ( 1 + 𝑛2 βˆ’ 𝑛12 βˆ’ 𝑛23 𝑛3 1 + 𝑛2 βˆ’ 𝑛12 βˆ’ 𝑛23 𝑛3 + 2𝑛2 βˆ’ 𝑛 βˆ’ 2 ( ) π‘Žπ‘› = = = . 3𝑛3 + 𝑛 βˆ’ 11 3 + 𝑛12 βˆ’ 𝑛113 𝑛3 3 + 𝑛12 βˆ’ 𝑛113

After pulling out the 𝑛3 , we see the numerator tends to 1 as 𝑛 β†’ ∞ and the denominator tends to 3 as 𝑛 β†’ ∞. We can now use the limit of a quotient is the quotient of the limit, and find ( ) limπ‘›β†’βˆž 1 + 𝑛2 βˆ’ 𝑛12 βˆ’ 𝑛23 1 + 𝑛2 βˆ’ 𝑛12 βˆ’ 𝑛23 1 ( ) = = . lim π‘Žπ‘› = lim 1 11 1 11 π‘›β†’βˆž π‘›β†’βˆž 3 3 + 𝑛2 βˆ’ 𝑛3 limπ‘›β†’βˆž 3 + 𝑛2 βˆ’ 𝑛3

Alternatively, we can use L’Hopital’s rule to evaluate the limit; we keep taking derivatives until we no longer have infinity over infinity: 𝑛3 + 2𝑛2 βˆ’ 𝑛 βˆ’ 2 π‘›β†’βˆž 3𝑛3 + 𝑛 βˆ’ 11 lim

= = =

3𝑛2 + 4𝑛 βˆ’ 1 π‘›β†’βˆž 9𝑛2 + 1 6𝑛 + 4 lim π‘›β†’βˆž 18𝑛 6 1 lim = . π‘›β†’βˆž 18 3 lim

Question: Similar Problem to #7: Find the limit of π‘Žπ‘› = limit does not exist. 35

4𝑛2 βˆ’11𝑛+1 , 5𝑛6 +12

or prove the

Solution: The limit is zero. One way to see this is to pull out the highest power of 𝑛 from the numerator and the denominator; it is 𝑛2 for the numerator and 𝑛6 for the denominator. We have ( ) + 𝑛12 𝑛2 4 βˆ’ 11 + 𝑛12 4 βˆ’ 11 4𝑛2 βˆ’ 11𝑛 + 1 𝑛 𝑛 ( ( ) ) . π‘Žπ‘› = = = 5𝑛6 + 12 𝑛6 +5 𝑛126 𝑛4 +5 𝑛126

Note the numerator tends to 4 as 𝑛 β†’ ∞ while the denominator tends to infinity; thus the ratio tends to 0. Alternatively, we could use L’Hopital’s rule, taking derivatives until we no longer have infinity over infinity: 4𝑛2 βˆ’ 11𝑛 + 1 π‘›β†’βˆž 5𝑛6 + 12 lim

= =

8𝑛 βˆ’ 11 30𝑛5 8 lim = 0. π‘›β†’βˆž 150𝑛4 lim

π‘›β†’βˆž

7.2 Page 10.6 Question: Similar Problem to #8: Find the limit of the series it does not exist. Solution: Note that this sum is the same as ∞ ( )𝑛 βˆ‘ 2 ; 3 𝑛=0

2𝑛 𝑛=0 3𝑛

βˆ‘βˆž

or prove

this is the same as a geometric series with ratio which is less than 1 in βˆ‘π‘Ÿ = 2/3, 𝑛 absolute value. We know the geometric series ∞ π‘Ÿ converges if βˆ£π‘Ÿβˆ£ < 1; thus 𝑛=0 this series converges. Question: #10: Find 𝑛 such that 1 + 21 + 13 + β‹… β‹… β‹… + 𝑛1 > 106 . Solution: The 𝑛th harmonic number, 𝐻𝑛 , is defined by 1 1 1 + +β‹…β‹…β‹…+ . 2 3 𝑛 We have 𝐻100 β‰ˆ 5.2, 𝐻1000 β‰ˆ 7.5, 𝐻1010 β‰ˆ 23.6, 𝐻10100 β‰ˆ 230.8, 𝐻101000 β‰ˆ 2303.2, and so on. Note how slowly this grows! In fact, 𝐻𝑛 β‰ˆ ln 𝑛 for 𝑛 large. The sought after value of 𝑛 is about 10434295 , which is quite large! 𝐻𝑛 = 1 +

36

There are ways to find 𝑛 that will work without knowing 𝐻𝑛 β‰ˆ ln 𝑛. One way is to note that 1/3 + 1/4 β‰₯ 1/2, 1/5 + 1/6 + 1/7 + 1/8 β‰₯ 1/2, 1/9 + β‹… β‹… β‹…+ 1/16 β‰₯ 1/2 and so on. Thus we can keep getting at least 1/2....

7.3 Page 10.7 Question: Similar Problem to #13: Find all 𝑝 such that the sequence π‘Žπ‘› = 𝑛 ln1𝑝 𝑛 converges. Solution: For any fixed 𝑝, once 𝑛 is large the sequence is strictly decreasing and we can use the integral test. Thus the series converges or diverges depending on whether or not ∫ ∞ 1 𝑑π‘₯ 𝑝 π‘₯=big π‘₯ ln π‘₯ converges or diverges; we write β€˜big’ to indicate that the lower bound does not really matter – what matters is the behavior at infinity. We integrate by parts. Let 𝑒 = ln π‘₯ so 𝑑𝑒 = 𝑑π‘₯/π‘₯, and thus our integral becomes ∫ ∞ π‘’βˆ’π‘ 𝑑𝑒. 𝑒=ln(big)

1βˆ’π‘

The integral of π‘’βˆ’π‘ is 𝑒 𝑝 if 𝑝 βˆ•= 1 and ln 𝑒 if 𝑝 = 1. Thus the integral converges if 𝑝 > 1 and diverges 𝑝 ≀ 1.

7.4 Page 10.8 βˆ‘ 1√ Question: Similar Problem to #14: Determine if the series ∞ 𝑛=1 2𝑛 + 𝑛 converges or diverges. βˆ‘ √1 Solution: We use the comparison test. Note that while ∞ 𝑛=1 𝑛 diverges, βˆ‘βˆž 1 √ 𝑛 𝑛 β‰₯ 2𝑛 , we have 𝑛=1 2𝑛 converges. As 2 + 0≀

1 1 √ ≀ . 2𝑛 2𝑛 + 𝑛

Thus the series converges by the comparison test. 37

Question: Similar Problem to #15: Determine if the series or diverges. √ Solution: We have 𝑛 ≀ 2𝑛 . Thus ( )𝑛 √ 𝑛 2𝑛 2 ≀ 𝑛 = . 𝑛 5 5 5

√

𝑛 𝑛=1 5𝑛

βˆ‘βˆž

converges

Our series is thus bounded term by term by the geometric series with ratio 2/5, and thus converges by the comparison. βˆ‘ 𝑛 √2 ⋅𝑛! Question: Similar Problem to #16: Determine if the series ∞ 𝑛=1 𝑛+11 converges or diverges. Solution: This series diverges. Note the numerator is growing much faster than the denominator. The easiest way to see this is that the denominator is at most 𝑛 for large 𝑛, and 𝑛!/𝑛 = (𝑛 βˆ’ 1)!. In other words, the terms in the sequence tend to infinity, and thus the sum cannot converge.

7.5 Page 10.10 βˆ‘ 2π‘˜+1 Question: #20: Determine if π‘›π‘˜=0 310π‘˜ converges or diverges. Solution: We can re-write the terms in a more illuminating manner. We have 𝑛 βˆ‘ 32π‘˜+1 π‘˜=0

10π‘˜

𝑛 βˆ‘ 32π‘˜ 3 10π‘˜

=

π‘˜=0 𝑛 βˆ‘

9π‘˜ 10π‘˜ π‘˜=0 )π‘˜ 𝑛 ( βˆ‘ 9 3 . 10 π‘˜=0

=

3

=

Note the above sum is just three times the geometric series with ratio 3, and thus converges.

38

Instead of using the comparison test we can also use the Ratio Test. We look at π‘Žπ‘˜+1 /π‘Žπ‘˜ , which for us is π‘Žπ‘˜+1 32π‘˜+3 /10π‘˜+1 9 = = . π‘Žπ‘˜ 32π‘˜+1 /10π‘˜ 10 Thus the limit of π‘Žπ‘˜+1 /π‘Žπ‘˜ is 9/10 < 1, and the series therefore converges by the Ratio Test. βˆ‘ π‘˜ 4 Question: #22: Determine if π‘›π‘˜=0 3 (π‘˜ 5+π‘˜+1) converges or diverges. π‘˜ Solution: If we didn’t have the factor π‘˜ 4 + π‘˜ + 1, it would be straightforward, as the series would just be the geometric series with ratio 3/5. As π‘˜ 4 grows polynomially but 3π‘˜ and 5π‘˜ grow exponentially, we expect the series to still converge. Thus we look for an upper bound for the numerator such that, even when multiplied by 3π‘˜ , it grows slower than the denominator by a significant margin. For example, let’s try and show the numerator is bounded by 𝐢 β‹… 4π‘˜ for some constant 𝐢. We want to show for π‘˜ large that 3π‘˜ (π‘˜ 4 + π‘˜ + 1) ≀ 𝐢 β‹… 4π‘˜ , or π‘˜ 4 + π‘˜ + 1 ≀ 𝐢 β‹… (4/3)π‘˜ .

We have π‘˜ 4 + π‘˜ + 1 ≀ 3π‘˜ 4 , and thus if we take 𝐢 = 3 we need only show, for π‘˜ large, that π‘˜ 4 ≀ (4/3)π‘˜ . While this is not true for small π‘˜, it is true for large π‘˜ and thus the series is bounded by the geometric series with ratio 4/5, and hence converges. We now provide an alternative proof using the ratio test. We look at π‘Žπ‘˜+1 /π‘Žπ‘˜ , which for us is (3/5)π‘˜+1((π‘˜ + 1)4 + π‘˜ + 2) 3 (π‘˜ + 1)4 + π‘˜ + 2 π‘Žπ‘˜+1 = = β‹… . π‘Žπ‘˜ (3/5)π‘˜ (π‘˜ 4 + π‘˜ + 1) 5 π‘˜4 + π‘˜ + 1 If we take the limit as π‘˜ β†’ ∞, we see that the limit is 3/5. As this is less than 1, by the Ratio Test the series converges.

8 From path integrals to Stokes’ Theorem The final homework assignment of the semester is: Section 4.2: #1 (see formula at the bottom of the page for help). Section 4.4: #1, #14. Section 7.1: #3b. Section 7.2: #1c. Section 8.1: #3a. The problems below are similarly chosen problems. 39

Question: Section 4.2: #2: Find the arc length of the curve 𝑐(𝑑) = (1, 3𝑑2 , 𝑑3 ) for 0 ≀ 𝑑 ≀ 1. Solution: The answer is ∫ 1 βˆ£βˆ£π‘β€² (𝑑)∣∣ 𝑑𝑑, 0

where

𝑐′ (𝑑) = (0, 6𝑑, 3𝑑2) so βˆ£βˆ£π‘β€² (𝑑)∣∣ =

√

√ 36𝑑2 + 9𝑑4 = 3𝑑 4 + 𝑑2 ;

this will lead to a straightforward integral because of the factor of 𝑑 outside the square-root. We have ∫ 1 ∫ 1 √ β€² βˆ£βˆ£π‘ (𝑑)∣∣ 𝑑𝑑 = 3𝑑 4 + 𝑑2 𝑑𝑑 0 0 ∫ 3 1 (4 + 𝑑2 )1/2 2𝑑𝑑𝑑 = 2 0 1 2 3/2 = (4 + 𝑑 ) 0

=

5

3/2

βˆ’4

3/2

.

Question: Section 4.4: #2: Find the divergence and the curl of 𝑉 (π‘₯, 𝑦, 𝑧) = (𝑦𝑧, π‘₯𝑧, π‘₯𝑦) = (𝑉1 (π‘₯, 𝑦, 𝑧), 𝑉2 (π‘₯, 𝑦, 𝑧), 𝑉3(π‘₯, 𝑦, 𝑧)). Solution: The divergence is div(𝑉 ) = βˆ‡ β‹… 𝑉 =

βˆ‚π‘‰1 βˆ‚π‘‰2 βˆ‚π‘‰3 + + = 0; βˆ‚π‘₯ βˆ‚π‘¦ βˆ‚π‘§

the fact that the divergence is zero has physical interpretations. For the curl, we have β†’ βˆ’ βˆ’ β†’ βˆ’π‘– β†’ 𝑗 π‘˜ βˆ‚ βˆ‚ βˆ‚ curl(𝑉 ) = βˆ‡ Γ— 𝑉 = βˆ‚π‘₯ . βˆ‚π‘¦ βˆ‚π‘§ 𝑉 𝑉 𝑉 1 2 2 Expanding gives ( ) βˆ‚π‘‰3 βˆ‚π‘‰2 βˆ‚π‘‰1 βˆ‚π‘‰3 βˆ‚π‘‰2 βˆ‚π‘‰1 βˆ’ , βˆ’ , βˆ’ = (0, 0, 0) ; βˆ‚π‘¦ βˆ‚π‘§ βˆ‚π‘§ βˆ‚π‘₯ βˆ‚π‘₯ βˆ‚π‘¦ 40

thus this vector field has both zero curl and zero divergence! ∫ Question: Section 7.1: #3a: Find the path integral 𝑓 (π‘₯, 𝑦, 𝑧)𝑑𝑠 where 𝑓 (π‘₯, 𝑦, 𝑧) = 𝑐 √ 2 exp( 𝑧) and 𝑐(𝑑) = (1, 2, 𝑑 ) for 0 ≀ 𝑑 ≀ 1. Solution: The path integral is ∫ 1 𝑓 (𝑐(𝑑))βˆ£βˆ£π‘β€² (𝑑)∣∣ 𝑑𝑑. 0

We have 𝑐′ (𝑑) = (0, 0, 2𝑑), βˆ£βˆ£π‘β€² (𝑑)∣∣ = 2βˆ£π‘‘βˆ£ (which is 2𝑑 as 𝑑 β‰₯ 0). Further,

√ 𝑓 (𝑐(𝑑)) = 𝑓 (1, 2, 𝑑2) = exp( 𝑑2 ) = exp(𝑑).

Thus the path integral is ∫ 1 𝑓 (𝑐(𝑑))βˆ£βˆ£π‘β€²(𝑑)∣∣ 𝑑𝑑

=

0

∫

1

exp(𝑑)2𝑑𝑑𝑑. 0

The integral (or anti-derivative) of exp(𝑑)𝑑 is just exp(𝑑)(𝑑 βˆ’ 1), and thus we have 1 ∫ 1 𝑓 (𝑐(𝑑))βˆ£βˆ£π‘β€²(𝑑)∣∣ 𝑑𝑑 = 2 exp(𝑑)(𝑑 βˆ’ 1) = 2. 0 0

Question: Section 7.2: #1b: Let 𝐹 (π‘₯, 𝑦, 𝑧) = (π‘₯, 𝑦, 𝑧). Evaluate the integral of 𝐹 along the path 𝑐(𝑑) = (sin 𝑑, 0, cos 𝑑) for 0 ≀ 𝑑 ≀ 2πœ‹. Solution: We have 𝑐′ (𝑑) = (cos 𝑑, 0, βˆ’ sin 𝑑) , 𝐹 (𝑐(𝑑)) = 𝐹 (sin 𝑑, 0, cos 𝑑) = (sin 𝑑, 0, cos 𝑑). Thus the line integral is ∫ 2πœ‹ 𝐹 (𝑐(𝑑)) β‹… 𝑐′ (𝑑)𝑑𝑑

=

0

=

∫

∫

1 0

(sin 𝑑, 0, cos 𝑑)(cos 𝑑, 0, βˆ’ sin 𝑑)𝑑𝑑 2πœ‹

0𝑑𝑑 = 0. 0

41

Question: Section 8.1: #3b: Verify Green’s theorem for the disk with center (0, 0) and radius 𝑅 and the functions 𝑃 (π‘₯, 𝑦) = π‘₯ + 𝑦, 𝑄(π‘₯, 𝑦) = 𝑦. Solution: We have βˆ‚π‘„ βˆ‚π‘ƒ βˆ’ = 0 βˆ’ 1 = βˆ’1; βˆ‚π‘₯ βˆ‚π‘¦ thus

∫ ∫ ( 𝐷

βˆ‚π‘„ βˆ‚π‘ƒ βˆ’ βˆ‚π‘₯ βˆ‚π‘¦

)

𝑑π‘₯𝑑𝑦 =

∫ ∫

𝐷

βˆ’1𝑑π‘₯𝑑𝑦 = βˆ’Area(𝐷);

of course, the area of the disk is πœ‹π‘…2 so this double integral is βˆ’πœ‹π‘…2 . For the other part of Green’s theorem, we note the boundary curve is 𝑐(𝑑) = (𝑅 cos 𝑑, 𝑅 sin 𝑑), 0 ≀ 𝑑 ≀ 2πœ‹ (remember we must travel so that the region 𝐷 is on our left). Thus 𝑐′ (𝑑) = (βˆ’π‘… sin 𝑑, 𝑅 cos 𝑑). Further, β†’ βˆ’ β†’ βˆ’ 𝐹 (𝑐(𝑑)) = 𝐹 (𝑅 cos 𝑑, 𝑅 sin 𝑑) = (𝑅 cos 𝑑 + 𝑅 sin 𝑑, 𝑅 sin 𝑑), and hence ∫ β†’ βˆ’ β†’ 𝐹 β‹… π‘‘βˆ’ 𝑠

=

𝑐

=

∫

2πœ‹

∫0 2πœ‹ 0

=

∫

0

β†’ βˆ’ 𝐹 (𝑐(𝑑)) β‹… 𝑐′ (𝑑)𝑑𝑑 (𝑅 cos 𝑑 + 𝑅 sin 𝑑, 𝑅 sin 𝑑) β‹… (βˆ’π‘… sin 𝑑, 𝑅 cos 𝑑)𝑑𝑑

2πœ‹

βˆ’π‘…2 sin2 𝑑𝑑𝑑.

We evaluated the sine-integral many ways, and found it equals πœ‹; one could also use trig identities and find sin2 𝑑 = 1βˆ’cos(2𝑑) . Thus the integral equals πœ‹ times 2 2 βˆ’π‘… , which does match.

42