This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

An elementary multidimensional fundamental theorem of calculus

Joaquim Bruna Departament de Matemàtiques, Edifici Cc, Universitat Autònoma de Barcelona, 08193 Cerdanyola del Vallès, Barcelona, Spain. joaquim.bruna@uab.cat
Abstract.

We discuss a version of the fundamental theorem of calculus in several variables and some applications, of potential interest as a teaching material in undergraduate courses.

Key words and phrases:
Integral; Derivative; Interval Functions; Density
2020 Mathematics Subject Classification:
Primary: 26B05; Secondary: 26A24; 26B30; 28A10; 28A15.

1. Introduction and main result

In standard undergraduate courses on one variable calculus, differentiation and integration are presented as inverse processes, as stated in the fundamental theorem of calculus: if FF is differentiable at every point x[a,b]x\in[a,b] and F=fF^{\prime}=f is integrable in [a,b][a,b], then

F(x)F(a)=axf(t)𝑑t,axb.F(x)-F(a)=\int_{a}^{x}f(t)\,dt,\quad a\leqslant x\leqslant b.

This holds both in the context of Riemann and Lebesgue integration (see [4] for a proof in the Lebesgue integration context).

In this note we provide a multidimensional version of this statement. The proof is straightforward and can be included in an undergraduate course of multidimensional calculus.

The first basic concept is that of interval or cube function Φ\Phi on a domain UnU\subset\mathbb{R}^{n}. By an interval QQ in UU we understand a set Q=j=1nIjQ=\prod_{j=1}^{n}I_{j} with the Ij=II_{j}=I, II a one-dimensional closed interval, that is the faces of QQ are parallel to the coordinate axis. An interval function is a map defined on all intervals QnQ\subset\mathbb{R}^{n} assigning to each QQ a real or complex number Φ(Q)\Phi(Q) with the property that

Φ(Q)=iΦ(Qi),\Phi(Q)=\sum_{i}\Phi(Q_{i}),

whenever (Qi)(Q_{i}) is a finite partition of QQ, that is, Q=QiQ=\cup Q_{i} and the QiQ_{i} have disjoint interiors. The easiest example is

Φf(Q)=Qf𝑑x,\Phi_{f}(Q)=\int_{Q}f\,dx,

where ff is locally integrable.

We have in mind two other examples. For the first one, assume that T:UVT:U\rightarrow V is a measurable homeomorphism between two domains such that m(T(A))=0m(T(A))=0 if m(A)=0m(A)=0. Here and in the following m(A)m(A) denotes the Lebesgue measure of AA. Then Φ(Q)=m(T(Q))\Phi(Q)=m(T(Q)) is an interval function, because images of the faces have zero measure. For the second one assume that FF is a continuous vector field in the plane or in space and set

Φ(Q)=QF,N𝑑mn1,\Phi(Q)=\int_{\partial Q}\langle F,N\rangle\,dm_{n-1},

the flow of FF through the boundary Q\partial Q oriented with the outward normal NN. Then Φ(Q)\Phi(Q) is an interval function. This is because if Qi,QjQ_{i},Q_{j} are two intervals with a face SS in common, the outward normals are opposite one each other.

Notice that a Dirac delta at a point aUa\in U, that is, Φ(Q)=1\Phi(Q)=1 if aQa\in Q and zero otherwise, is not an interval function according to our definition, because if aa is a boundary point of both Q1,Q2Q_{1},Q_{2} then Φ(Q1)=Φ(Q2)=Φ(Q)=1\Phi(Q_{1})=\Phi(Q_{2})=\Phi(Q)=1.

In dimension n=1n=1, with U=(a,b)U=(a,b), if gg is defined on (a,b)(a,b), it is immediately seen that

Φ([c,d])=g(d)g(c),\Phi([c,d])=g(d)-g(c), (1)

defines an interval function on (a,b)(a,b). Indeed, a decomposition of [c,d][c,d] into pieces QiQ_{i} amounts to a selection of intermediate points (the end-points of the QiQ_{i}) c=t0<t1<<tm=dc=t_{0}<t_{1}<\cdots<t_{m}=d, and then

Φ(I)=g(d)g(c)=ig(ti+1)g(ti)=iΦ(Qi).\Phi(I)=g(d)-g(c)=\sum_{i}g(t_{i+1})-g(t_{i})=\sum_{i}\Phi(Q_{i}).

Conversely, given an interval function defined on (a,b)(a,b) and p(a,b)p\in(a,b), the function

g(x)=Φ([p,x]),px;g(x)=Φ([x,p]),xp,g(x)=\Phi([p,x]),p\leqslant x;g(x)=-\Phi([x,p]),x\leqslant p,

satisfies (1). Thus there is a one-to-one correspondence between interval functions and classical functions.

The second basic concept is that of density. For an interval function Φ\Phi we define its upper density

D¯Φ(x)=lim supxQ,δ(Q)0Φ(Q)m(Q)=infεsupδ(Q)ε,xQΦ(Q)m(Q),\overline{D}_{\Phi}(x)=\limsup_{x\in Q,\delta(Q)\to 0}\frac{\Phi(Q)}{m(Q)}=\inf_{\varepsilon}\sup_{\delta(Q)\leqslant\varepsilon,x\in Q}\frac{\Phi(Q)}{m(Q)},

where m(Q)m(Q) denotes the measure of QQ and δ(Q)\delta(Q) its diameter. Analogously the lower density is defined

D¯Φ(x)=lim infxQ,δ(Q)0Φ(Q)m(Q)=supεinfδ(Q)ε,xQΦ(Q)m(Q).\underline{D}_{\Phi}(x)=\liminf_{x\in Q,\delta(Q)\to 0}\frac{\Phi(Q)}{m(Q)}=\sup_{\varepsilon}\inf_{\delta(Q)\leqslant\varepsilon,x\in Q}\frac{\Phi(Q)}{m(Q)}.

In case both are finite and equal we say that Φ\Phi has a finite density DΦ(x)D_{\Phi}(x) at xx.

For instance, if ff is continuous in UU, the density of Φf\Phi_{f} is ff at all points. Indeed, given ε>0\varepsilon>0 there is τ\tau such that |f(y)f(x)|ε|f(y)-f(x)|\leqslant\varepsilon if |xy|τ|x-y|\leqslant\tau. Then, if δ(Q)<τ,xQ\delta(Q)<\tau,x\in Q one has |f(y)f(x)|ε|f(y)-f(x)|\leqslant\varepsilon for all yQy\in Q so

|Φf(Q)m(Q)f(x)|=|1m(Q)Q(f(y)f(x))𝑑y|1m(Q)Q|f(y)f(x)|𝑑yε.|\frac{\Phi_{f}(Q)}{m(Q)}-f(x)|=|\frac{1}{m(Q)}\int_{Q}(f(y)-f(x))\,dy|\leqslant\frac{1}{m(Q)}\int_{Q}|f(y)-f(x)|\,dy\leqslant\varepsilon.

Thus

limδ(Q)0Φf(Q)m(Q)=f(x).\lim_{\delta(Q)\to 0}\frac{\Phi_{f}(Q)}{m(Q)}=f(x).

A deeper result is Lebesgue’s differentiation theorem (see [3]) stating that Φf\Phi_{f} has density f(x)f(x) at almost all points xUx\in U under the sole assumption that ff is locally integrable.

For a better understanding of the density consider the following example. Assume U=(0,1)×(0,1)U=(0,1)\times(0,1) and let L={(x,x),0<x<1}L=\{(x,x),0<x<1\} be the diagonal. Define Φ(Q)\Phi(Q) as the length of LQL\cap Q, clearly an interval function. Then DΦ(x)=0D_{\Phi}(x)=0 for xLx\notin L while D¯Φ(x)=0,D¯Φ(x)=+\underline{D}_{\Phi}(x)=0,\overline{D}_{\Phi}(x)=+\infty for xLx\in L.

In dimension one, if Φ\Phi is given by (1)\eqref{dim1}, Φ\Phi has a finite density at xx if and only gg is differentiable at xx, because if x[c,d]x\in[c,d]

g(d)g(c)dc=g(d)g(x)dxdxdc+g(x)g(c)xcxcdc.\frac{g(d)-g(c)}{d-c}=\frac{g(d)-g(x)}{d-x}\frac{d-x}{d-c}+\frac{g(x)-g(c)}{x-c}\frac{x-c}{d-c}.

Next elementary theorem seems to be unnoticed, to the best of author’s knowledge. It holds both in the context of Riemann and Lebesgue’s integration.

Theorem 1.1.

If an interval function Φ\Phi has a finite upper density D¯Φ\overline{D}_{\Phi} at every point and D¯Φ\overline{D}_{\Phi} is locally integrable, then for every cube QUQ\subset U

Φ(Q)QD¯Φ(x)𝑑x.\Phi(Q)\leqslant\int_{Q}\overline{D}_{\Phi}(x)\,dx.

Analogously, if Φ\Phi has a finite lower density D¯Φ\underline{D}_{\Phi} at every point and D¯Φ\underline{D}_{\Phi} is locally integrable, then

Φ(Q)QD¯Φ(x)𝑑x.\Phi(Q)\geqslant\int_{Q}\underline{D}_{\Phi}(x)\,dx.

Thus,

Φ(Q)=QDΦ(x)𝑑x,\Phi(Q)=\int_{Q}D_{\Phi}(x)\,dx,

whenever Φ\Phi has a finite integrable density at every point.

In an informal way, if Φ(Q)\Phi(Q) is of the order of f(x)m(Q)f(x)m(Q) for infinitesimal cubes xQx\in Q, then Φ(Q)=Qf(x)𝑑x\Phi(Q)=\int_{Q}f(x)\,dx for big cubes.

In dimension one, in view of the remark before the theorem, this is the fundamental theorem of calculus stated in the beginning.

As a corollary we may state:

Corollary. For an interval function Φ\Phi and a continuous function ff on UU the following two statements are equivalent:

limxQ,δ(Q)0Φ(Q)m(Q)=f(x),Φ(Q)=Qf(x)𝑑x.\lim_{x\in Q,\delta(Q)\to 0}\frac{\Phi(Q)}{m(Q)}=f(x),\quad\Phi(Q)=\int_{Q}f(x)\,dx.

We point out some remarks. First, it is essential, as in one variable, that the density is assumed to exist at every point. If it exists just a.e. then the theorem does not hold. Secondly, in other type of results the a.e. existence of the density is actually proved like in Lebesgue’s differentiation theorem quoted before. In fact, the interval functions Φf\Phi_{f} are characterized as those being absolutely continuous, meaning that for every ε>0\varepsilon>0 there exists δ>0\delta>0 such that i|Φ(Qi)|<ε\sum_{i}|\Phi(Q_{i})|<\varepsilon whenever QiQ_{i} are non-overlapping cubes and im(Qi)<δ\sum_{i}m(Q_{i})<\delta. So the result can be rephrased by saying that interval functions having finite integrable density at all points are automatically absolutely continuous. A reference for all these results is [3].

2. Proof

Let QUQ\subset U be a cube and let us break it into 2n2^{n} cubes SiS_{i} of equal measure. Since Φ(Q)=Φ(Si)\Phi(Q)=\sum\Phi(S_{i}), one has

Φ(Si)Φ(Q)2n,\Phi(S_{i})\geqslant\frac{\Phi(Q)}{2^{n}},

whence

Φ(Si)m(Si)Φ(Q)m(Q),\frac{\Phi(S_{i})}{m(S_{i})}\geqslant\frac{\Phi(Q)}{m(Q)},

for at least one ii. Repeating the argument we find a sequence QkQ_{k} of cubes, QkQQ_{k}\subset Q, shrinking to some point pQp\in Q such that

Φ(Qk)m(Qk)Φ(Q)m(Q).\frac{\Phi(Q_{k})}{m(Q_{k})}\geqslant\frac{\Phi(Q)}{m(Q)}.

Therefore D¯Φ(p)Φ(Q)m(Q)\overline{D}_{\Phi}(p)\geqslant\frac{\Phi(Q)}{m(Q)}. So Φ(Q)D¯Φ(p)m(Q)\Phi(Q)\leqslant\overline{D}_{\Phi}(p)\,m(Q) for some point pQp\in Q. Similarly, Φ(Q)D¯Φ(p)m(Q)\Phi(Q)\geqslant\underline{D}_{\Phi}(p)\,m(Q) for some point. This holds for all cubes. So Φ0\Phi\leqslant 0 if D¯Φ0\overline{D}_{\Phi}\leqslant 0. Similarly, Φ0\Phi\geqslant 0 if D¯Φ0\underline{D}_{\Phi}\geqslant 0.

We complete now the proof in the Riemann integration context, where just the definition of the Riemann integral is used.

Let (Qi)(Q_{i}) be a partition of QQ; then

Φ(Q)=iΦ(Qi)iD¯Φ(pi)m(Qi).\Phi(Q)=\sum_{i}\Phi(Q_{i})\leqslant\sum_{i}\overline{D}_{\Phi}(p_{i})m(Q_{i}).

Therefore, if D¯Φ\overline{D}_{\Phi} is Riemann integrable, it follows that

Φ(Q)QD¯Φ(x)𝑑x.\Phi(Q)\leqslant\int_{Q}\overline{D}_{\Phi}(x)\,dx.

In a similar way we see that

Φ(Q)QD¯Φ(x)𝑑x,\Phi(Q)\geqslant\int_{Q}\underline{D}_{\Phi}(x)\,dx,

and so the theorem is proved when the density is Riemann integrable.

Assume now that D¯Φ\overline{D}_{\Phi} is Lebesgue integrable on UU. We may assume Φ\Phi real-valued and use semi-continuous functions as in [4]. Recall that a function gg is called lower semi-continuous at a point pp if lim infxpg(x)g(p)\liminf_{x\to p}g(x)\geqslant g(p) and upper semi-continuous if lim infxpg(x)g(p)\liminf_{x\to p}g(x)\leqslant g(p).

Given ε>0\varepsilon>0, by the Vitali-Carathedory theorem, there is a lower semi-continuous function vv such that D¯Φv\overline{D}_{\Phi}\leqslant v and U(vD¯Φ)𝑑x<ε\int_{U}(v-\overline{D}_{\Phi})\,dx<\varepsilon. Define

Ψ(Q)=Qv𝑑xΦ(Q).\Psi(Q)=\int_{Q}v\,dx-\Phi(Q).

Then, vv being lower semi-continuous,

D¯Ψ(x)=lim infxQΨ(Q)m(Q)lim infxQ1m(Q)Qvdxlim supxQΦ(Q)m(Q))v(x)D¯Φ(x)0.\underline{D}_{\Psi}(x)=\liminf_{x\in Q}\frac{\Psi(Q)}{m(Q)}\geqslant\liminf_{x\in Q}\frac{1}{m(Q)}\int_{Q}v\,dx-\limsup_{x\in Q}\frac{\Phi(Q)}{m(Q)})\geqslant v(x)-\overline{D}_{\Phi}(x)\geqslant 0.

Therefore Ψ(Q)0\Psi(Q)\geqslant 0, whence

Φ(Q)Qv𝑑x=QD¯Φ𝑑x+Q(vD¯Φ)𝑑x<QD¯Φ𝑑x+ε.\Phi(Q)\leqslant\int_{Q}v\,dx=\int_{Q}\overline{D}_{\Phi}\,dx+\int_{Q}(v-\overline{D}_{\Phi})\,dx<\int_{Q}\overline{D}_{\Phi}\,dx+\varepsilon.

Since ε\varepsilon is arbitrary, this shows that Φ(Q)QD¯Φ𝑑x\Phi(Q)\leqslant\int_{Q}\overline{D}_{\Phi}\,dx and applying the same argument to Φ-\Phi we are done.

3. Applications

1. As a first application we indicate a simplified proof of a version of the change of variables formula, with minimal assumptions and not relying in the one-dimensional version and Fubini’s theorem, the one stated in Theorem 7.26 in [4]:

Theorem 3.1.

Let T:UVT:U\rightarrow V be an homeomorphism between two domains in n\mathbb{R}^{n}, differentiable at every point xUx\in U. Assume that |detdT(x)||\det dT(x)| is integrable on UU. Then for a positive measurable function ff in VV one has

Vf(y)𝑑y=Uf(T(x))|detdT(x)|𝑑x.\int_{V}f(y)\,dy=\int_{U}f(T(x))|\det dT(x)|\,dx. (2)

Note that the assumption |detdT(x)|0|\det dT(x)|\neq 0 is not made, so this version includes Sard’s theorem.

We modify the proof in [4] replacing the more advanced Radon-Nikodym differentiation theorem for absolutely continuous measures by theorem 1.1.

First, Lemma 7.25 in [4] proves that TT maps sets of measure zero to sets of measure zero. As a consequence,

Φ(Q)=m(T(Q))\Phi(Q)=m(T(Q))

is an interval function.

Secondly, theorem 7.24 in [4] proves that Φ\Phi has density |detdT(p)||\det dT(p)| at every point pp. In fact, the proof in [4] uses balls, but it is easily checked that it holds for cubes too. We explain the basic idea for completeness. By hypothesis, we can approximate T(p+h)T(p+h) near pp by L=L(p+h)=T(p)+dT(p)(h)L=L(p+h)=T(p)+dT(p)(h), and use that m(L(Q))=|detL|m(Q)m(L(Q))=|\det L|\,m(Q) for all affine maps.

One has

T(p+h)=L+E,|E|τ(|h|)|h|,τ(h)0.T(p+h)=L+E,|E|\leqslant\tau(|h|)|h|,\tau(h)\to 0.

with decreasing τ(t)\tau(t) as t0t\to 0.

Let vj=Txj(p),j=1,,nv_{j}=\frac{\partial T}{\partial x_{j}}(p),j=1,\cdots,n be the columns of dT(p)dT(p). If QQ has side δ\delta, LL maps QQ onto a parallelepiped PP with spanning vectors δvj\delta v_{j}, whose measure is

m(P)=|detdT(p)|δn=|detdT(p)|m(Q),m(P)=|\det dT(p)|\delta^{n}=|\det dT(p)|m(Q),

so let us compare T(Q)T(Q) with P=L(Q)P=L(Q). Since |TL|=|E|τ(|h|)|h||T-L|=|E|\leqslant\tau(|h|)|h|, it is clear that T(Q)T(Q) is included in a parallelepiped P1P_{1} concentric with PP with spanning vectors (δ+o(δ))vj(\delta+o(\delta))v_{j}, whose measure is

δn|detdT(p)|+o(δn).\delta^{n}|\det dT(p)|+o(\delta^{n}).

Again by |TL|=|E|τ(|h|)|h||T-L|=|E|\leqslant\tau(|h|)|h|, the boundary b(T(Q))=T(bQ)b(T(Q))=T(bQ) is at distance less than τ(δ)δ\tau(\delta)\delta from bPbP. Since TT is an homeomorphism, this implies that T(Q)T(Q) contains a parallelepiped concentric with PP with spanning vectors (δo(δ))vj(\delta-o(\delta))v_{j} (see figure 1 for n=2n=2; a rigorous proof of this fact relies on Brouwer’s fixed point theorem and can be found in lemma 7.23 of [4]), whose measure is

δn|detdT(p)|o(δn).\delta^{n}|\det dT(p)|-o(\delta^{n}).

Altogether, since m(Q)=δnm(Q)=\delta^{n},

m(Q)|detdT(p)|o(m(Q))m(T(Q))m(Q)|detdT(p)|+o(m(Q)),m(Q)|\det dT(p)|-o(m(Q))\leqslant m(T(Q))\leqslant m(Q)|\det dT(p)|+o(m(Q)),

proving that Φ\Phi has density |detdT(p)||\det dT(p)| at pp.

By theorem 1.1, one has

m(T(A))=A|detdT(x)|𝑑x,m(T(A))=\int_{A}|\det dT(x)|\,dx,

when AA is a cube, whence when AA is a finite union of cubes too. Since every open set is a countable union of cubes, by the monotone convergence theorem this holds when AA is an open set and in turn when AA is a countable intersection of open sets, a GδG_{\delta} set. Since every measurable set differs from s GδG_{\delta} set in a set of zero measure and TT preserves those, we conclude that this holds for all measurable sets AUA\subset U, that is, (2)holds for the characteristic function of a measurable set. By linearity it then holds for simple functions, and by the monotone convergence theorem again, for a general measurable function.

Refer to caption
Figure 1.

Remark. As a first remark for the instructor, in case |detdT(x)|0|\det dT(x)|\neq 0 for all xUx\in U, the use of Brouwer’s fixed point theorem can be avoided as follows:

By the inclusion T(Q)P1T(Q)\subset P_{1},

m(T(Q))m(Q)|detdT(p)|+o(m(Q)),m(T(Q))\leqslant m(Q)|\det dT(p)|+o(m(Q)),

implying D¯Φ(p)|detdT(p)|\overline{D}_{\Phi}(p)\leqslant|\det dT(p)|. Then theorem 1.1 implies

m(T(A))A|detdT(x)|𝑑x,m(T(A))\leqslant\int_{A}|\det dT(x)|\,dx,

for all cubes, and as before this leads to

Vf(y)𝑑yUf(T(x))|detdT(x)|𝑑x.\int_{V}f(y)\,dy\leqslant\int_{U}f(T(x))|\det dT(x)|\,dx.

But since the same inequality applies to the inverse f1f^{-1}, the result follows.

Remark. As a second remark, to be eventually combined with the previous one, a proof in the context of Riemann integration can be further simplified as follows. To prove (2) say for a continuous function ff with compact support, introduce

Ψ(Q)=T(Q)f𝑑y.\Psi(Q)=\int_{T(Q)}f\,dy.

The continuity of ff implies

DΨ(p)=f(T(p))DΦ(p),D_{\Psi}(p)=f(T(p))D_{\Phi}(p),

so DΨ(p)=f(T(p))|detdT(p)|D_{\Psi}(p)=f(T(p))|\det dT(p)|. This leads using theorem 1.1 to

T(Q)f(y)𝑑y=Qf(T(x))|detdT(x)|𝑑x,\int_{T(Q)}f(y)\,dy=\int_{Q}f(T(x))|\det dT(x)|\,dx,

for all cubes. If KK is the support of ff, the compact T1(K)T^{-1}(K) can be covered by a finite number of cubes QQ, so (2) follows.

2. As a second application we analyze the divergence theorem. Assume that FF is a continuous vector field in space and set

Φ(Q)=QF,N𝑑mn1,\Phi(Q)=\int_{\partial Q}\langle F,N\rangle\,dm_{n-1},

the flow of FF through the boundary Q\partial Q oriented with the outward normal NN. We mentioned before that Φ\Phi is indeed an interval function. If its density exists, we call it the divergence divF\operatorname{div}F of FF. If integrable, the theorem implies

bQF,N𝑑mn1=QdivFdmn,\int_{bQ}\langle F,N\rangle\,dm_{n-1}=\int_{Q}\operatorname{div}F\,dm_{n},

and the same holds with QQ replaced by a finite union of cubes. From this it follows by approximations that the same holds with QQ replaced by a domain with piece-wise regular boundary (details can be found in [1]).

If FF is differentiable with components FiF_{i}, let us check that the density divF\operatorname{div}F exists at every point and equals ,F=iDiFi\langle\nabla,F\rangle=\sum_{i}D_{i}F_{i}.

First consider an affine field F(x)=M(XP)F(x)=M(X-P), where M=(mij)M=(m_{ij}) is a constant matrix and X,PX,P are the column vectors xt,ptx^{t},p^{t}, and let us compute the flux across the boundary Q\partial Q of a parallelepiped in space spanned by 33 vectors v1,v2,v3v_{1},v_{2},v_{3}

Q={p+t1v1+t2v2+t3v3,0ti1},Q=\{p^{\prime}+t_{1}v_{1}+t_{2}v_{2}+t_{3}v_{3},0\leqslant t_{i}\leqslant 1\},

containing pp, oriented by the outward normal. FF differs from M(XP)M(X-P^{\prime}) by a constant field, which obviously has zero flux, so we can replace pp by pp^{\prime} and assume p=0p^{\prime}=0. On the face t3=1t_{3}=1, the basis v1,v2v_{1},v_{2} is positively oriented and the flux is

0101det(M(t1v1+t2v2+v3),v1,v2)dt1dt2,\int_{0}^{1}\int_{0}^{1}\det(M(t_{1}v_{1}+t_{2}v_{2}+v_{3}),v_{1},v_{2})\,dt_{1}\,dt_{2},

while on the face t3=0t_{3}=0 it is

0101det(M(t1v1+t2v2),v1,v2)dt1dt2.-\int_{0}^{1}\int_{0}^{1}\det(M(t_{1}v_{1}+t_{2}v_{2}),v_{1},v_{2})\,dt_{1}\,dt_{2}.

Therefore they add up to

det(M(v3),v1,v2)\det(M(v_{3}),v_{1},v_{2})

If M(v3)=iλiviM(v_{3})=\sum_{i}\lambda_{i}v_{i}, this equals λ3det(v3,v1,v2)\lambda_{3}\det(v_{3},v_{1},v_{2}). The same applies to the other two couples of opposite sides, whence the flux is exactly

trace(M)det(v1,v2,v3),\operatorname{trace}(M)\det(v_{1},v_{2},v_{3}),

the trace of MM times the volume of QQ.

Now let FF be a differentiable field at pp, QQ a cube of size δ\delta containing pp. As before we expand FF around pp

F(x)=F(p)+dF(p)(XP)+E,E=o(|xp|).F(x)=F(p)+dF(p)(X-P)+E,E=o(|x-p|).

The contribution to the flux of FF across Q\partial Q of the constant field F(p)F(p) is zero, that of the linear field dF(p)(XP)dF(p)(X-P) is the trace of dF(p)dF(p) times m(Q)m(Q) while that of EE is o(δn)o(\delta^{n}), whence the flux equals

(D1F1++DnFn)(p)m(Q)+o(m(Q)),(D_{1}F_{1}+\cdots+D_{n}F_{n})(p)m(Q)+o(m(Q)),

thus proving that the density is ,F\langle\nabla,F\rangle.

Upon replacement of the field F=(A,B)F=(A,B) by JF=(B.A)JF=(-B.A) the divergence theorem in the plane amounts to Green’s formula. Using the language of line integrals, if Pdx+QdyP\,dx+Q\,dy is a differentiable 11-form and QxPyQ_{x}-P_{y} is integrable one has

bUP𝑑x+Qdy=U(QxPy)𝑑A,\int_{bU}P\,dx+Q\,dy=\int_{U}(Q_{x}-P_{y})\,dA,

with no assumption needed separately for Qx,PyQ_{x},P_{y}. A particular case are complex line integrals

Φ(Q)=bQf(z)𝑑z,\Phi(Q)=\int_{bQ}f(z)\,dz,

for ff continuous in the complex plane \mathbb{C}. For differentiable ff the density is ¯f\overline{\partial}f, and so if integrable one has

bUf(z)𝑑z=U¯f(z)𝑑A(z).\int_{bU}f(z)\,dz=\int_{U}\overline{\partial}f(z)\,dA(z).

Other general versions of Green’s theorem with minimal assumptions are known, but the proofs are far from elementary (see [2] and references herein).

3. In a surface SS in 3\mathbb{R}^{3} oriented by a unit normal field NN one can define cubes as those which are so in a local chart. If FF is a continuous field, the circulation

Φ(Q)=bQF,T𝑑s,\Phi(Q)=\int_{bQ}\langle F,T\rangle\,ds,

defines an interval function. If FF is differentiable, one can show along the same lines that the density is ×F,N\langle\nabla\times F,N\rangle and one gets Stoke’s theorem with minimal assumptions (see the details in the book [1]).

Acknowledgements

The author is partially supported by the Ministry of Science and Innovation–Research Agency of the Spanish Government through grant PID2021-123405NB-I00 and by AGAUR, Generalitat de Catalunya, through grant 2021-SGR-00087. The author thanks his colleague Julià Cufí for valuable comments.

References

  • [1] J. Bruna, Analysis in Euclidean Space, Essential Textbooks in Mathematics, World Scientific, November 2022
  • [2] R.M. Fesq, Green’s formula, linear continuity, and Hausdorff measure. Trans. Amer. Math. Soc. 118 (1965), 105-112.
  • [3] S. Lojasiewicz, An introduction to the theory of real functions, Wiley-Interscience, 1988.
  • [4] W. Rudin, Real and Complex Analysis, Mc Graw-hill, 1987.