How to prove that $limlimits_{xto0}frac{sin x}x=1$?












392












$begingroup$


How can one prove the statement
$$lim_{xto 0}frac{sin x}x=1$$
without using the Taylor series of $sin$, $cos$ and $tan$? Best would be a geometrical solution.



This is homework. In my math class, we are about to prove that $sin$ is continuous. We found out, that proving the above statement is enough for proving the continuity of $sin$, but I can't find out how. Any help is appreciated.










share|cite|improve this question











$endgroup$








  • 51




    $begingroup$
    l'Hôpital's rule is easiest: $limlimits_{xto0}sin x = 0$ and $limlimits_{xto0}x = 0$, so $limlimits_{xto 0}frac{sin x}x = limlimits_{xto 0}frac{cos x}1 = 1 $
    $endgroup$
    – Joren
    Oct 23 '11 at 20:41






  • 142




    $begingroup$
    @Joren: I'm extremely curious how will you prove then that $sin ' x = cos x$
    $endgroup$
    – Ilya
    Oct 24 '11 at 9:10






  • 3




    $begingroup$
    @FUZx44xl: sure, but to be fare you first prove that $sin xsim x$ with $xto 0$. Geometrically
    $endgroup$
    – Ilya
    Oct 24 '11 at 15:42








  • 11




    $begingroup$
    Recent changes in what? The definition of $lim$, of $sin$ or of $0$?
    $endgroup$
    – Asaf Karagila
    Dec 18 '13 at 20:03






  • 2




    $begingroup$
    @FUZxxl:Exactly what was your definition by "geometrical means"?
    $endgroup$
    – Platonix
    Dec 22 '13 at 18:45
















392












$begingroup$


How can one prove the statement
$$lim_{xto 0}frac{sin x}x=1$$
without using the Taylor series of $sin$, $cos$ and $tan$? Best would be a geometrical solution.



This is homework. In my math class, we are about to prove that $sin$ is continuous. We found out, that proving the above statement is enough for proving the continuity of $sin$, but I can't find out how. Any help is appreciated.










share|cite|improve this question











$endgroup$








  • 51




    $begingroup$
    l'Hôpital's rule is easiest: $limlimits_{xto0}sin x = 0$ and $limlimits_{xto0}x = 0$, so $limlimits_{xto 0}frac{sin x}x = limlimits_{xto 0}frac{cos x}1 = 1 $
    $endgroup$
    – Joren
    Oct 23 '11 at 20:41






  • 142




    $begingroup$
    @Joren: I'm extremely curious how will you prove then that $sin ' x = cos x$
    $endgroup$
    – Ilya
    Oct 24 '11 at 9:10






  • 3




    $begingroup$
    @FUZx44xl: sure, but to be fare you first prove that $sin xsim x$ with $xto 0$. Geometrically
    $endgroup$
    – Ilya
    Oct 24 '11 at 15:42








  • 11




    $begingroup$
    Recent changes in what? The definition of $lim$, of $sin$ or of $0$?
    $endgroup$
    – Asaf Karagila
    Dec 18 '13 at 20:03






  • 2




    $begingroup$
    @FUZxxl:Exactly what was your definition by "geometrical means"?
    $endgroup$
    – Platonix
    Dec 22 '13 at 18:45














392












392








392


166



$begingroup$


How can one prove the statement
$$lim_{xto 0}frac{sin x}x=1$$
without using the Taylor series of $sin$, $cos$ and $tan$? Best would be a geometrical solution.



This is homework. In my math class, we are about to prove that $sin$ is continuous. We found out, that proving the above statement is enough for proving the continuity of $sin$, but I can't find out how. Any help is appreciated.










share|cite|improve this question











$endgroup$




How can one prove the statement
$$lim_{xto 0}frac{sin x}x=1$$
without using the Taylor series of $sin$, $cos$ and $tan$? Best would be a geometrical solution.



This is homework. In my math class, we are about to prove that $sin$ is continuous. We found out, that proving the above statement is enough for proving the continuity of $sin$, but I can't find out how. Any help is appreciated.







calculus limits trigonometry proof-explanation limits-without-lhopital






share|cite|improve this question















share|cite|improve this question













share|cite|improve this question




share|cite|improve this question








edited Feb 4 '18 at 18:19









Michael Hardy

1




1










asked Oct 23 '11 at 16:21









FUZxxlFUZxxl

3,78052143




3,78052143








  • 51




    $begingroup$
    l'Hôpital's rule is easiest: $limlimits_{xto0}sin x = 0$ and $limlimits_{xto0}x = 0$, so $limlimits_{xto 0}frac{sin x}x = limlimits_{xto 0}frac{cos x}1 = 1 $
    $endgroup$
    – Joren
    Oct 23 '11 at 20:41






  • 142




    $begingroup$
    @Joren: I'm extremely curious how will you prove then that $sin ' x = cos x$
    $endgroup$
    – Ilya
    Oct 24 '11 at 9:10






  • 3




    $begingroup$
    @FUZx44xl: sure, but to be fare you first prove that $sin xsim x$ with $xto 0$. Geometrically
    $endgroup$
    – Ilya
    Oct 24 '11 at 15:42








  • 11




    $begingroup$
    Recent changes in what? The definition of $lim$, of $sin$ or of $0$?
    $endgroup$
    – Asaf Karagila
    Dec 18 '13 at 20:03






  • 2




    $begingroup$
    @FUZxxl:Exactly what was your definition by "geometrical means"?
    $endgroup$
    – Platonix
    Dec 22 '13 at 18:45














  • 51




    $begingroup$
    l'Hôpital's rule is easiest: $limlimits_{xto0}sin x = 0$ and $limlimits_{xto0}x = 0$, so $limlimits_{xto 0}frac{sin x}x = limlimits_{xto 0}frac{cos x}1 = 1 $
    $endgroup$
    – Joren
    Oct 23 '11 at 20:41






  • 142




    $begingroup$
    @Joren: I'm extremely curious how will you prove then that $sin ' x = cos x$
    $endgroup$
    – Ilya
    Oct 24 '11 at 9:10






  • 3




    $begingroup$
    @FUZx44xl: sure, but to be fare you first prove that $sin xsim x$ with $xto 0$. Geometrically
    $endgroup$
    – Ilya
    Oct 24 '11 at 15:42








  • 11




    $begingroup$
    Recent changes in what? The definition of $lim$, of $sin$ or of $0$?
    $endgroup$
    – Asaf Karagila
    Dec 18 '13 at 20:03






  • 2




    $begingroup$
    @FUZxxl:Exactly what was your definition by "geometrical means"?
    $endgroup$
    – Platonix
    Dec 22 '13 at 18:45








51




51




$begingroup$
l'Hôpital's rule is easiest: $limlimits_{xto0}sin x = 0$ and $limlimits_{xto0}x = 0$, so $limlimits_{xto 0}frac{sin x}x = limlimits_{xto 0}frac{cos x}1 = 1 $
$endgroup$
– Joren
Oct 23 '11 at 20:41




$begingroup$
l'Hôpital's rule is easiest: $limlimits_{xto0}sin x = 0$ and $limlimits_{xto0}x = 0$, so $limlimits_{xto 0}frac{sin x}x = limlimits_{xto 0}frac{cos x}1 = 1 $
$endgroup$
– Joren
Oct 23 '11 at 20:41




142




142




$begingroup$
@Joren: I'm extremely curious how will you prove then that $sin ' x = cos x$
$endgroup$
– Ilya
Oct 24 '11 at 9:10




$begingroup$
@Joren: I'm extremely curious how will you prove then that $sin ' x = cos x$
$endgroup$
– Ilya
Oct 24 '11 at 9:10




3




3




$begingroup$
@FUZx44xl: sure, but to be fare you first prove that $sin xsim x$ with $xto 0$. Geometrically
$endgroup$
– Ilya
Oct 24 '11 at 15:42






$begingroup$
@FUZx44xl: sure, but to be fare you first prove that $sin xsim x$ with $xto 0$. Geometrically
$endgroup$
– Ilya
Oct 24 '11 at 15:42






11




11




$begingroup$
Recent changes in what? The definition of $lim$, of $sin$ or of $0$?
$endgroup$
– Asaf Karagila
Dec 18 '13 at 20:03




$begingroup$
Recent changes in what? The definition of $lim$, of $sin$ or of $0$?
$endgroup$
– Asaf Karagila
Dec 18 '13 at 20:03




2




2




$begingroup$
@FUZxxl:Exactly what was your definition by "geometrical means"?
$endgroup$
– Platonix
Dec 22 '13 at 18:45




$begingroup$
@FUZxxl:Exactly what was your definition by "geometrical means"?
$endgroup$
– Platonix
Dec 22 '13 at 18:45










22 Answers
22






active

oldest

votes


















479












$begingroup$

sinc and tanc at 0



The area of $triangle ABC$ is $frac{1}{2}sin(x)$. The area of the colored wedge is $frac{1}{2}x$, and the area of $triangle ABD$ is $frac{1}{2}tan(x)$. By inclusion, we get
$$
frac{1}{2}tan(x)gefrac{1}{2}xgefrac{1}{2}sin(x)tag{1}
$$
Dividing $(1)$ by $frac{1}{2}sin(x)$ and taking reciprocals, we get
$$
cos(x)lefrac{sin(x)}{x}le1tag{2}
$$
Since $frac{sin(x)}{x}$ and $cos(x)$ are even functions, $(2)$ is valid for any non-zero $x$ between $-frac{pi}{2}$ and $frac{pi}{2}$. Furthermore, since $cos(x)$ is continuous near $0$ and $cos(0) = 1$, we get that
$$
lim_{xto0}frac{sin(x)}{x}=1tag{3}
$$
Also, dividing $(2)$ by $cos(x)$, we get that
$$
1lefrac{tan(x)}{x}lesec(x)tag{4}
$$
Since $sec(x)$ is continuous near $0$ and $sec(0) = 1$, we get that
$$
lim_{xto0}frac{tan(x)}{x}=1tag{5}
$$






share|cite|improve this answer











$endgroup$









  • 71




    $begingroup$
    Do my homework! Note that $(1)$ says that for $0le xlefrac{pi}{2}$ we have $0lesin(x)le x$; therefore, $limlimits_{xto0^+}sin(x)=0$. Then $cos(x)=1-2sin^2(x/2)$ should finish the job.
    $endgroup$
    – robjohn
    Oct 23 '11 at 18:37








  • 3




    $begingroup$
    Thank you very much. I know that proverb, but I really wasn't able to find that out on my own.
    $endgroup$
    – FUZxxl
    Oct 23 '11 at 18:41






  • 46




    $begingroup$
    From your comment, I wasn't expecting that you could find it on your own, but "Read the question!" seemed a bit rough around the edges.
    $endgroup$
    – robjohn
    Oct 23 '11 at 19:04






  • 22




    $begingroup$
    Sorry for that.
    $endgroup$
    – FUZxxl
    Oct 23 '11 at 19:35






  • 6




    $begingroup$
    I think that your justification is slightly circular :). In an introductory calculus course, $(sin x,cos x)$ is probably defined to be the point we reach after traveling $x$ units counterclockwise along the unit circle from $(1,0)$.
    $endgroup$
    – Mike F
    Oct 23 '11 at 20:54





















119












$begingroup$

You should first prove that for $x > 0$ small that $sin x < x < tan x$. Then, dividing by $x$ you get
$$
{ sin x over x} < 1
$$
and rearranging $1 < {tan x over x} = {sin x over x cos x }$
$$
cos x < {sin x over x}.
$$
Taking $x rightarrow 0^+$ you apply the squeeze theorem. For $x < 0$ and small use that $sin(-x) = -sin x$ so that $${sin(-x) over -x} = {sin x over x}.$$
As far as why the first inequality I said is true, you can do this completely from triangles but I don't know how to draw the pictures here.



diagram






share|cite|improve this answer











$endgroup$









  • 2




    $begingroup$
    But how to prove that $sin x<x<tan x$?
    $endgroup$
    – FUZxxl
    Oct 23 '11 at 16:31










  • $begingroup$
    It is in the picture. The definition of radians makes the picture above true. Maybe that is worth mentioning: this limit explicitly depends on "$x$" being measured in radians.
    $endgroup$
    – tkr
    Oct 23 '11 at 16:33








  • 2




    $begingroup$
    This is a strange picture! Normally you want the $tan(theta)$ side to be parallel to the $sin(theta)$ side.
    $endgroup$
    – user641
    Oct 23 '11 at 17:36






  • 9




    $begingroup$
    If you make $tan(theta)$ parallel then you need to make the points $S$ and $Q$ the same. For whatever reason, this is the proof I like the most because it relates the tangent line at the point on the circle to the value we call "tangent". To each his own...
    $endgroup$
    – tkr
    Oct 23 '11 at 18:22








  • 2




    $begingroup$
    Even with this picture, you could use the area inclusion principle and argue that area of triangle OPQ < Area of sector OPQ < Area of triangle OPS. This translates into $sin theta < theta < tan theta$ which is the first line of this answer and hence, nothing strange about it.
    $endgroup$
    – Deepak Gupta
    Sep 2 '15 at 21:05



















88












$begingroup$

Usually calculus textbooks do this using geometric arguments followed by squeezing.



Here's an Euler-esque way of looking at it---not a "proof" as that term is usually understood today, but still worth knowing about.



Let $theta$ be the length of an arc along the circle of unit radius centered at $(0,0)$, from the point $(1,0)$ in a counterclockwise direction to some point $(costheta,sintheta)$ on the circle. Then of course $sintheta$ is the height of the latter point above the $x$-axis. Now imagine what happens if $theta$ is an infinitely small positive number. Then the arc is just an infinitely short vertical line, and the height of the endpoint above the $x$-axis is just the length of the arc. I.e. when $theta$ is an infinitely small number, then $sintheta$ is the same as $theta$. It follows that when $theta$ is an infinitely small nonzero number, then $dfrac{sintheta}{theta}=1$.



That is how Euler viewed the matter. See his book on differential calculus.






share|cite|improve this answer











$endgroup$









  • 1




    $begingroup$
    I disagree with the recent edit to my answer and I have reverted to the previous version. "Infinitesmal" means "infinitely small". $qquad$
    $endgroup$
    – Michael Hardy
    Jan 21 '16 at 19:12










  • $begingroup$
    You read Euler's book ? Was it very difficult to read because of the notation and language of the time ?
    $endgroup$
    – user230452
    Feb 27 '16 at 4:03






  • 1




    $begingroup$
    @user230452 : Just some parts of it. I wouldn't say the differences in language and notation were the challenging part. $qquad$
    $endgroup$
    – Michael Hardy
    Feb 27 '16 at 18:18










  • $begingroup$
    Wouldn't one rather say that $frac{sin theta}{theta}$ is infinitely close to 1 if $theta$ is an infinitely small nonzero number?
    $endgroup$
    – Sven
    Apr 14 '17 at 13:05






  • 2




    $begingroup$
    @Sven : That is indeed how it's done in Robinson's "nonstandard analysis". There is another approach to rigorous infinitesimals in which it would be done the way Euler did it, saying that if $theta$ is infinitely small then $dfrac{sintheta}theta = 1.$ It's called "smooth infinitesimal analysis". In that approach, the square of an infinitesimal is $0$, so we have $$ frac{sintheta} theta = frac{theta - dfrac{theta^3} 6 + dfrac{theta^5}{120} - cdots cdots} theta = 1. $$ en.wikipedia.org/wiki/Smooth_infinitesimal_analysis $qquad$
    $endgroup$
    – Michael Hardy
    Apr 14 '17 at 18:48





















73












$begingroup$

Look at this link:



http://fatosmatematicos.blogspot.com/2010/08/provas-sem-palavras-parte-20.html



Here is the picture I copied from that blog:



Copy of the picture from the Fatos Matematicos blog






share|cite|improve this answer











$endgroup$













  • $begingroup$
    +1, nice site - What is the length BC?
    $endgroup$
    – NoChance
    Oct 24 '11 at 5:34






  • 11




    $begingroup$
    What is the argument for showing that $theta cos thetale sin theta$? The picture isn't immediately convincing.
    $endgroup$
    – Mark Viola
    Apr 15 '15 at 5:31










  • $begingroup$
    @Dr. MV: if you rescale the circle in $cos theta$ units, then the original $sin theta$ is a scaled $tan theta$ which is always longer than its arc.
    $endgroup$
    – Mattsteel
    May 30 '16 at 20:38



















49












$begingroup$

Here you may see an elementary approach that starts from a very interesting result, see this problem. All you need is a bit of imagination. When we take $lim_{nrightarrowinfty} frac{nsin(frac{pi}{n})}{1+sin(frac{pi}{n})}$ we may notice that we have infinitely many circles surrounding the unit circle with infinitely small diameters that lastly perfectly approximate the length of the unit circle when having it there infinity times . Therefore when multiplying n by the radius under the limit to infinity we get π. Let's denote $frac{pi}{n}$ by x.



$$lim_{xrightarrow0}frac{pisin(x)}{x(1+sin(x))}={pi}Rightarrowlim_{xrightarrow0}frac{sin(x)}{x(1+sin(x))}=1Rightarrowlim_{xrightarrow0}frac{sin(x)}{x}=1$$



The proof is complete.






share|cite|improve this answer











$endgroup$









  • 1




    $begingroup$
    this assumes the a priori knowledge of the existence of the limit $lim_{xto 0}frac{sin(x)}{x}$, how would you go about proving it before hand? (+1) by the way for the original and alternative proof
    $endgroup$
    – user153330
    Feb 24 '16 at 14:07



















49












$begingroup$

I claim that for $0<x<pi/2$ that the following holds
$$sin x lt x lt tan x$$
Figure 1

In the diagram, we let $OC=OA=1$. In other words, $Arc:CA=x$ is an arc of a unit circle. The shortest distance from point $C$ to line $AO$ is line $CE=sin x$ (because $CEperp OA$). Another path from point $C$ to line $OA$ is arc $CA$ (which is longer than CE because it is not the shortest path). So we have at the very least
$$sin x lt x$$
Now we need to show that line $BA=tan x gt x$.

Lines $AD$ and $CD$ are both tangent to arc $CA$. $CD+DA$ is longer than arc $CA$ because the set of points bound by sector $OCA$ is a subset of the set of points bound by quadrilateral $OCDA$, both of which are convex sets. This means that the perimeter of quadrilateral $OCDA$ must be longer than the perimeter of the sector $OCA$ (as per Archimedes, On the Sphere and Cylinder Book I). But both the sector and the quadrilateral both have sides $OC$ and $OA$, so we have
$$CA=x<DC+DA$$
But $BD>CD$ because it is the hypotenuse in $triangle BCD$ we have
$$tan x = BA = BD+DAgt CD+DA gt CA=x gt sin x$$



So we have
$$sin x lt x lt tan x$$
$$frac{sin x}{x} lt 1 lt frac{tan x}{x}=frac{sin x}{x}cdotsec x$$
From this we can extract
$$frac{sin x}{x} lt 1$$
and
$$1 lt frac{sin x}{x}cdotsec x$$
$$cos x lt frac{sin x}{x}$$
Putting these inequalities back together we have
$$cos x lt frac{sin x}{x} lt 1$$



Because $displaystylelim_{xto 0}cos x = 1$, by the squeeze theorem we have
$$lim_{xto 0}frac{sin x}{x}=1$$






share|cite|improve this answer









$endgroup$









  • 1




    $begingroup$
    The reason that I chose an arc length proof is because most derivations that I've seen of the area of a circle (see mathopenref.com/circleareaderive.html) assume that $displaystylelim_{xto 0}frac{sin x}{x}=1$. Using an area based proof, to me, seems like putting the cart before the horse.
    $endgroup$
    – John Joy
    Nov 9 '14 at 15:14










  • $begingroup$
    How did Archimedes avoid assuming $frac{chord}{arc}rightarrow 1$ as $arc$ tends to zero in his argument about perimeters that you referred to? I asked the following question because this issue has kept haunting me for a while.
    $endgroup$
    – String
    Feb 5 '15 at 9:34






  • 2




    $begingroup$
    Actually, what Archimedes assumed was that if 2 curves (both concave in the same direction) have the same endpoints and one of the curves is strictly between a line segment with the same end points and the other curve, then the length of the enclosed curve is shorter than the length of enclosing curve. See axiom #2 archive.org/stream/worksofarchimede00arch#page/4/mode/2up
    $endgroup$
    – John Joy
    Feb 5 '15 at 14:17








  • 2




    $begingroup$
    OK, so he essentially axiomatizes the idea that among all curves connecting endpoints $A$ and $B$ satisfying some kind of regularity condition - concavity in the same direction, we have this inequality based on enclosure. This makes $arc(CA)<CD+DA$ in your figure merely equivalent to the assumption itself, but it makes sense and is very nicely put by Archimedes! Thank you for pointing me to that!
    $endgroup$
    – String
    Feb 5 '15 at 14:52






  • 2




    $begingroup$
    If you don't have a copy of Archimedes's book, you can approximate sector OCA with a polygon inside OCDA and use cut-the-knot.org/m/Geometry/PerimetersOfTwoConvexPolygons.shtml
    $endgroup$
    – CopyPasteIt
    Jun 6 '17 at 16:06



















46












$begingroup$

I am not sure if it counts as proof, but I have seen this done by a High Schooler.
enter image description here



In the given picture above,
$displaystyle 2n text{ EJ} = 2nR sinleft( frac{pi}{n } right ) = text{ perimeter of polygon }$.



$displaystyle lim_{nto infty }2nR sinleft( frac{pi}{n } right ) = lim_{nto infty } (text{ perimeter of polygon }) = 2 pi R implies lim_{nto infty}frac{sinleft( frac{pi}{n } right )}{left( frac{pi}{n } right )} = 1$ and let $frac{pi}{n} = x$.






share|cite|improve this answer









$endgroup$









  • 7




    $begingroup$
    This method is usually used to prove that the perimeter of a circle is $2pi R$ using the fact $limlimits_{xto 0}frac{sin x}{x}=1$
    $endgroup$
    – BPP
    Jul 20 '13 at 19:15






  • 3




    $begingroup$
    Santosh, how does one prove that the perimeter of the polygon converges to the $ 2pi$ ?
    $endgroup$
    – Imago
    Feb 5 '16 at 19:45










  • $begingroup$
    @Imago math.stackexchange.com/questions/720935/…
    $endgroup$
    – user301988
    Jun 6 '16 at 3:32



















39












$begingroup$

Don't you feel strange about why most of the proofs are done with a figure? I've had this problem in the beginning, and realized after that this is due to the definition we use for the function $sin x$. Because the usual definition of $sin x$ we all study first in high schools depends on “classical geometry” and usually with a figure, you should depict out the figure and to make it clear.



However, if you use other definitions of $sin x$ that are equivalent to the former, you'll find it more simple. For example,



$$sin x = frac{x^1}{1!} - frac{x^3}{3!}+ frac{x^5}{5!} - cdots + cdots - cdots$$



and hence



$$frac{sin x}x = frac{x^0}{1!} - frac{x^2}{3!}+ frac{x^4}{5!} - cdots$$



which obviously tends to $1$ as $x$ approaches 0.






share|cite|improve this answer











$endgroup$









  • 4




    $begingroup$
    Indeed, it's easy to see that this holds if one uses a series, but this question starts on the prerequisite that one does not use a series.
    $endgroup$
    – FUZxxl
    Mar 13 '15 at 17:54






  • 2




    $begingroup$
    And how do you know that the derivative of an infinte series is equal to the sum of the derivatives of each term?
    $endgroup$
    – steven gregory
    Apr 16 '16 at 4:32






  • 3




    $begingroup$
    @StevenGregory: Where does he use that? He only divides by $x$. What he does use is that $frac{1}{x}lim_{ntoinfty} sum_{k=0}^n a_k = lim_{ntoinfty}sum_{k=0}^n frac{a_k}{x}$
    $endgroup$
    – celtschk
    Sep 24 '16 at 9:36












  • $begingroup$
    @celschk starting with the infinite series as a 'definition' causes more problems than it solves. Now we have to prove that this sine behaves like the sine we learned in high school. This proof is compelling but it is not really a proof.
    $endgroup$
    – steven gregory
    Sep 24 '16 at 14:44



















30












$begingroup$

Here's one more:
$$
lim_{x to 0} frac{sin x}{x}=lim_{x to 0} lim_{v to 0}frac{sin (x+v)-sin v}{x}\
=lim_{v to 0} lim_{x to 0}frac{sin (x+v)-sin v}{x}=lim_{v to 0}sin'v=lim_{v to 0} cos v=1
$$






share|cite|improve this answer









$endgroup$









  • 32




    $begingroup$
    Usually, this limit is used to compute the derivative of $sin(x)$.
    $endgroup$
    – robjohn
    Jul 20 '13 at 18:43



















28












$begingroup$

It depends on your definition of the sine function. I would suggest checking out the geometric proof in ProofWiki.






share|cite|improve this answer









$endgroup$





















    21












    $begingroup$

    The strategy is to find $frac{darcsin y}{dy}$ first. This can easily be done using the picture below.



    arcsin as area



    From the above picture, $arcsin y$ is twice the area of the orange bit. The area of the red bit is ${1 over 2}ysqrt{1-y^2}$. The area of the red bit plus the orange bit is $int_{0}^y sqrt{1-Y^2} dY$. So $$arcsin y = 2int_{0}^y sqrt{1-Y^2} dY - ysqrt{1-y^2}.$$ Differentiating with respect to $y$ gives $frac{darcsin y}{dy} = frac{1}{sqrt{1-y^2}}$. Using the theorem for the derivative of inverse functions yields $sin' theta = sqrt{1 - sin^2 theta} = cos theta$.



    (A similar thing can be done with the arc length definition of $arcsin$.)






    share|cite|improve this answer











    $endgroup$













    • $begingroup$
      While this is indeed an interesting approach, integrals haven't been taught at the point where this limit is proved. Thank you for your answer though.
      $endgroup$
      – FUZxxl
      Apr 11 '15 at 17:38



















    20












    $begingroup$

    Let $f:{yinmathbb{R}:yneq 0}tomathbb{R}$ be a function defined by $f(x):=dfrac{sin x}{x}$ for all $xin {yinmathbb{R}:yneq 0}$.



    We have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$ if and only if for every $varepsilon>0$, there exists a $delta>0$ such that $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



    Let $varepsilon>0$ be an arbitrary real number.



    Note that $sin x=displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n+1}}{(2n+1)!}$.



    If $x neq 0$, we have $dfrac{sin x}{x}=$$displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}=1+$$displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}$.



    We thus have



    $|f(x)-1|=left|dfrac{sin x}{x}-1right|=left|displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}right|leq left|displaystylesum_{n=1}^{infty} dfrac{x^{2n}}{(2n+1)!}right|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|$



    Therefore we have



    $|f(x)-1|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|leq displaystyle sum_{n=1}^{infty} |x^{2n}|=sum_{n=1}^{infty}|x^2|^n$



    If $0<|x|<1$, then $0<|x^2|<1$, and the infinite series $displaystylesum_{n=1}^{infty}|x^2|^n$ converges to $dfrac{x^2}{1-x^2}$.



    Choose $delta:=sqrt{dfrac{varepsilon}{1+varepsilon}}$.
    Then $0<|x-0|<delta$ implies that $0<|x|<$$sqrt{dfrac{varepsilon}{1+varepsilon}}<1$, and hence $x^2<varepsilon-varepsilon x^2$. But $x^2<varepsilon-varepsilon x^2$ implies that $dfrac{x^2}{1-x^2}<varepsilon$.



    We therefore have $sum_{n=1}^{infty}|x^2|^n<varepsilon$ whenever $0<|x-0|<delta$. But since $|f(x)-1|leqdisplaystylesum_{n=1}^{infty}|x^2|^n$, we have $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



    Since $varepsilon$ was arbitrary, we have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$.






    share|cite|improve this answer









    $endgroup$













    • $begingroup$
      The premise of the question was not to use power series.
      $endgroup$
      – FUZxxl
      Dec 4 '18 at 12:35



















    17












    $begingroup$


    Here is a different approach that uses the integral definition of the arcsine function. We will deduce the limit of interest without appeal to geometry or differential calculus.



    Instead, we only rely on elementary analysis of continuous functions and their inverses along with simple properties of the Riemann integral. To that end, we now proceed.






    We define the sine function, $sin(x)$, as the inverse function of the function $f(x)$ given by




    $$bbox[5px,border:2px solid #C0A000]{f(x)=int_0^x frac{1}{sqrt{1-t^2}},dt }tag 1$$




    for $|x|< 1$.




    NOTE: It can be shown that the sine function defined as the inverse of $f(x)$ given in $(1)$ has all of the familiar properties that characterize the circular function $sin(x)$.




    It is straightforward to show that since $frac{1}{sqrt{1-t^2}}$ is positive and continuous for $tin (-1,1)$, $f(x)$ is continuous and strictly increasing for $xin (-1,1)$ with $displaystylelim_{xto 0}f(x)=f(0)=0$.



    Therefore, since $f$ is continuous and strictly increasing, its inverse function, $sin(x)$, exists and is also continuous and strictly increasing with $displaystyle lim_{xto 0}sin(x)=sin(0)=0$.





    From $(1)$, we have the bounds (SEE HERE)




    $$bbox[5px,border:2px solid #C0A000]{1 le frac{f(x)}xle frac{1}{sqrt{1-x^2}}} tag 2$$




    for $xin (-1,1)$, whence applying the squeeze theorem to $(2)$ yields



    $$lim_{xto 0}frac{f(x)}{x}=1 tag 3$$





    Finally, let $y=f(x)$ so that $x=sin(y)$. As $xto 0$, $yto 0$ and we can write $(3)$ as



    $$lim_{yto 0}frac{y}{sin(y)}=1$$



    from which we have




    $$bbox[5px,border:2px solid #C0A000]{lim_{yto 0}frac{sin(y)}{y}=1}$$




    as was to be shown!






    NOTE:



    We can deduce the following set of useful inequalities from $(2)$. We let $x=sin(theta)$ and restrict $x$ so that $xin [0,1)$. In addition, we define new functions, $cos(theta)=sqrt{1-sin^2(theta)}$ and $tan(theta)=sin(theta)/cos(theta)$.



    Then, we have from $(2)$



    $$bbox[5px,border:2px solid #C0A000]{ycos(y)le sin(y)le yle tan(y)} $$



    which are the familiar inequalities often introduced in an introductory geometry or trigonometry course.







    share|cite|improve this answer











    $endgroup$









    • 1




      $begingroup$
      This is similar to my answer. You're formalizing the arc-length definition of $sin$ in the following three steps. 1) Defining $arcsin$ geometrically 2) Expressing that informal definition using integration 3) Defining $sin$ as the inverse of $arcsin$. Also, where you used the sqeeuze theorem, you could equally have used the Fundamental Theorem of Calculus and the Inverse Function Theorem. Then you'd get my answer
      $endgroup$
      – man on laptop
      Feb 5 '18 at 10:16








    • 1




      $begingroup$
      Another difference is you used the arc-length definition and I used the area definition
      $endgroup$
      – man on laptop
      Feb 5 '18 at 10:19






    • 1




      $begingroup$
      @ogogmad I don't use any geometrical argument. A function $f$ is defined without reference to the arc length interpretation. Both $f$ and its inverse has properties. It can be shown that the inverse function is the well-known sine function. And while one can apply the FOC instead of the squeeze theorem, there is no advantage in doing so. So, why on earth are you leaving these comments?
      $endgroup$
      – Mark Viola
      Feb 5 '18 at 14:50






    • 1




      $begingroup$
      Because I'm pointing out that your abstract definitions make geometric sense, whether or not your realize it. I'm also pointing out that your squeeze argument is just a special case of FOC
      $endgroup$
      – man on laptop
      Feb 5 '18 at 16:43






    • 1




      $begingroup$
      In fact, all of the arguments above are morally the same, AFAICT. The squeeze arguments are just unpacking the proof of FOC on a special case. The FOC itself is proved using a very similar squeeze argument to the one people are using. The areas of the sectors are simply the integrals that are being differentiated. In tkr's argument, he's instead differentiating an arclength integral, which gives the same result anyway. There's the remaining feature that our two answers uses $arcsin$ while everyone else's uses $sin$; they may be implicitly unpacking the proof of the inverse function theorem.
      $endgroup$
      – man on laptop
      Feb 5 '18 at 17:01



















    15












    $begingroup$

    Because $sin x$ has zeroes at $x=n pi$ for arbitrary integer $n$ including $x=0$, you can use Vieta's Theorem $sin x = A(cdots(x+2 pi)(x+pi)x(x-pi)(x-2 pi)cdots)$ with a constant $A$. Because $sin(frac{pi}{2})=1$ this constant can be determined by the equation: $$1=A(cdots(frac{pi}{2}+2 pi)(frac{pi}{2}+pi)frac{pi}{2}(frac{pi}{2}-pi)(frac{pi}{2}-2 pi)cdots).$$



    Now, in the Expression $f(x):= frac{sin(x)}{x}$ the $x$ cancels such that $$f(x)=A(cdots(x+2 pi)(x+pi)(x-pi)(x-2 pi)cdots),$$ hence: $$lim_{x rightarrow 0} f(x)=A(cdots(2 pi) cdot picdot(- pi)cdot(-2 pi)cdots) = A prod_{k=1}^infty (-k^2 pi^2).$$



    $frac{1}{A} = frac{pi}{2} prod_{k=1}^infty ((frac{pi}{2})^2-k^2 pi^2)$. The proof is completed when the Wallis product is used.






    share|cite|improve this answer











    $endgroup$









    • 5




      $begingroup$
      None of the infinite products in your answer converge.
      $endgroup$
      – Antonio Vargas
      Oct 15 '15 at 1:30






    • 2




      $begingroup$
      Ontop of Antonio Vargas comment, Vieta's theorem only works for finite number of zeros. You need the Weierstrass-Factorization Theorem or Hadamard-Factorization Theorem. Btw you also need to verify that $sin(x)$ does not have any complex zeros.
      $endgroup$
      – MrYouMath
      Jun 10 '16 at 12:52



















    14












    $begingroup$

    Usual proofs can be circular, but there is a simple way for proving such inequality.



    Let $theta$ be an acute angle and let $O,A,B,C,D,C'$ as in the following diagram:
    enter image description here



    We may show that:



    $$ CD stackrel{(1)}{ geq };stackrel{largefrown}{CB}; stackrel{(2)}{geq } CB,stackrel{(3)}{geq} AB $$



    $(1)$: The quadrilateral $OCDC'$ and the circle sector delimited by $O,C,C'$ are two convex sets. Since the circle sector is a subset of the quadrilateral, the perimeter of the circle sector is less than the perimeter of the quadrilateral.



    $(2)$: the $CB$ segment is the shortest path between $B$ and $C$.



    $(3)$ $CAB$ is a right triangle, hence $CBgeq AB$ by the Pythagorean theorem.



    In terms of $theta$ we get:
    $$ tantheta geq theta geq 2sinfrac{theta}{2} geq sintheta $$
    for any $thetainleft[0,frac{pi}{2}right)$. Since the involved functions are odd functions the reverse inequality holds over $left(-frac{pi}{2},0right]$, and $lim_{thetato 0}frac{sintheta}{theta}=1$ follows by squeezing.





    A slightly different approach might be the following one: let us assume $thetainleft(0,frac{pi}{2}right)$.
    By $(2)$ and $(3)$ we have
    $$ theta geq 2sinfrac{theta}{2}geq sintheta $$
    hence the sequence ${a_n}_{ngeq 0}$ defined by $a_n = 2^n sinfrac{theta}{2^n}$ is increasing and bounded by $theta$. Any increasing and bounded sequence is convergent, and we actually have $lim_{nto +infty}a_n=theta$ since $stackrel{largefrown}{BC}$ is a rectifiable curve and for every $ngeq 1$ the $a_n$ term is the length of a polygonal approximation of $stackrel{largefrown}{BC}$ through $2^{n-1}$ equal segments. In particular



    $$ forall thetainleft(0,frac{pi}{2}right), qquad lim_{nto +infty}frac{sinleft(frac{theta}{2^n}right)}{frac{theta}{2^n}} = 1 $$
    and this grants that if the limit $lim_{xto 0}frac{sin x}{x}$ exists, it is $1$. By $sin xleq x$ we get $limsup_{xto 0}frac{sin x}{x}leq 1$, hence it is enough to show that $liminf_{xto 0}frac{sin x}{x}geq 1$. We already know that for any $x$ close enough to the origin the sequence $frac{sin x}{x},frac{sin(x/2)}{x/2},frac{sin(x/4)}{x/4},ldots$ is convergent to $1$, hence we are done.



    Long story short: $lim_{xto 0}frac{sin x}{x}=1$ follows from the fact that a circle is a rectifiable curve, and a circle is a rectifiable curve because it is the boundary of a convex, bounded subset of $mathbb{R}^2$. The convexity of the disk follows from the triangle inequality: a disk is a closed ball for the euclidean distance.





    $(1)$ relies on this powerful Lemma:




    Lemma. If $A,B$ are convex bounded sets in $mathbb{R}^2$ and $Asubsetneq B$, the perimeter of $A$ is less than the perimeter of $B$.




    Proof: by boundedness and convexity, $partial A$ and $partial B$ are rectifiable, with lengths $L(A)=mu(partial A),,L(B)=mu(partial B)$. Always by convexity, there is some chord in $B$ that does not meet the interior of $A$ (a tangent to $partial A$ at a smooth point does the job, for instance). Assume that such chord has endpoints $B_1, B_2 in partial B$ and perform a cut along $B_1 B_2$: both the area and the perimeter of $B$ decrease, but $B$ remains a bounded convex set enclosing $A$. Since $A$ can be approximated through a sequence of consecutive cuts, $L(A)<L(B)$ follows.






    share|cite|improve this answer











    $endgroup$









    • 1




      $begingroup$
      Late to the party, but I think this is the non-circular (no puns intended) way of thinking that has been missing so far. Nice job!
      $endgroup$
      – String
      Jun 13 '17 at 9:05






    • 1




      $begingroup$
      Up voting! I think Archimedes used a weaker form of your Lemma to get the circumference of a circle.
      $endgroup$
      – CopyPasteIt
      Jun 21 '17 at 23:49










    • $begingroup$
      Great answer (+1)
      $endgroup$
      – MathLover
      Jan 25 '18 at 23:07



















    12












    $begingroup$

    Let $sin(x)$ is defined as solution of $frac{d^2}{dx^2}textrm{f}(x)=-textrm{f}(x)$ with $mathrm f(0)=0,,frac{d}{dx}mathrm f(0)=C$ initial conditions, so exact solution is $mathrm f(x)=Ccdotsin(x)$.
    Define second derivative as
    $$
    begin{align*}
    frac{d^2}{dx^2}textrm{f}(x)=lim_{Delta xto 0}{frac{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}-frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}{Delta x}}&=\=lim_{Delta xto 0}{frac{mathrm f(x)-2cdot mathrm f(x-Delta x)+mathrm f(x-2cdotDelta x)}{Delta x^2}}
    end{align*}
    $$
    we can easy check this limit for any (?) functions. Similarly, we can define the first derivative for right, middle and left points:
    $$
    frac{d}{dx}textrm{f}(x)left{
    begin{aligned}
    &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}}
    \
    &=lim_{Delta xto 0}{frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}\
    &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-2cdotDelta x)}{2cdotDelta x}}
    end{aligned}
    right.
    $$
    Let's use the finite elements method assuming $Td=Delta x,,y_n=mathrm f(x),,y_{n-1}=mathrm f(x-Delta x),,y_{n-2}=mathrm f(x-2cdot Delta x)$
    Override differential equation as
    $$
    frac{y_n-2cdot y_{n-1}+y_{n-2}}{Td^2}=-y_n
    $$
    Now solve this implicit equation for $y_n$ to obtain explicit recurrence relation:
    $$
    y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
    $$
    Using arbitrarily small but non-zero quantity Td we can plot exponentially decaying sampled sine function (because the poles are inside the unit circle of the transfer function corresponding to the given recurrence relation).
    Similarly we write three systems for the initial conditions:



    $$
    left{
    begin{aligned}
    &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
    \
    &C=frac{y_n-y_{n-1}}{Td}
    end{aligned}right.
    $$
    $$
    left{
    begin{aligned}
    &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
    \
    &C=frac{y_{n-1}-y_{n-2}}{Td}
    end{aligned}right.
    $$
    $$
    left{
    begin{aligned}
    &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
    \
    &C=frac{y_n-y_{n-2}}{2cdot Td}
    end{aligned}right.
    $$
    Solve this sequence of equations for $y_{n-1}$ and $y_{n-2}$:
    $$
    left{
    begin{aligned}
    &y_{n-1} = -Ccdot Td + y_{n}\
    &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1-Td^2right)
    end{aligned}right.
    $$
    $$
    left{
    begin{aligned}
    &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+Td^2right)\
    &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1+Td^2right)
    end{aligned}right.
    $$
    $$
    left{
    begin{aligned}
    &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+frac{Td^2}{2}right)\
    &y_{n-2}=-2cdot Ccdot Td + y_{n}
    end{aligned}right.
    $$
    At zero point $y_n=mathrm f(0)=0$ and we can see linear dependence:
    $$
    begin{aligned}
    &y_{n-1} = -Ccdot Td\
    &y_{n-2}=-2cdot Ccdot Td
    end{aligned}
    $$
    for all three solutions. Replace back:
    $$
    begin{array}{l}
    mathrm f(0)&=0\
    mathrm f(0-Delta x) &= -Ccdot Delta x\
    mathrm f(0-2cdot Delta x) &= -2cdot Ccdot Delta x
    end{array}
    $$
    So all three $frac{d}{dx}mathrm f(0)$ limits is equal to $C$ at $x=0$ and in accordance with $mathrm f(x)=Ccdotsin(x)$ by definition we can write
    $$
    lim_{Delta xto 0}{frac{mathrm f(0)-mathrm f(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{0-(-C cdot Delta x)}{Delta x}}=C
    $$
    Thus
    $$
    lim_{Delta xto 0}{frac{sin(0)-Ccdotsin(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{Ccdotsin(Delta x)}{Delta x}}=Ccdotlim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=C
    $$
    and $lim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=1$






    share|cite|improve this answer









    $endgroup$













    • $begingroup$
      While this is an exceptionally neat answer, it is not really appropriate for the level (high school) on which I asked the question, given that computing this limit is an introductory exercise into differentials. Plus we defined $sin$ geometrically (i.e. though the relationship between angles and sides in a right triangle), so your prerequisites must be established first.
      $endgroup$
      – FUZxxl
      Aug 19 '16 at 20:40










    • $begingroup$
      @FUZxxl do you realize that users with different levels and mathematical background may check this question? The answer given above might be helpful for them.
      $endgroup$
      – Xam
      Sep 23 '17 at 3:38



















    11












    $begingroup$

    Simple one is using sandwich theorem Which demonstrated earlier.In this method you have to show that $frac{sin x}{x} $ lies between other two functions. As $x longrightarrow 0$ both of them will tends to ONE.



    Then as in the case of sandwich (if both the bread part go to one stomach the middle portion will also go to the same stomach) $frac{sin x}{x}$ will go to ONE.



    You can use geogebra to see the visualization of this phenomena using geogebra.First you input $sin x$ and $x$ and observe that near to $0$ values of $sin x$ and $x$ are same.



    Secondly input $frac{sin x}{x}$ then observe function is approaching to $1$ as $x$ tends to $0$






    share|cite|improve this answer











    $endgroup$









    • 4




      $begingroup$
      This describes the Sandwich Theorem, but does not answer the question. At best, this should be a comment to the question,
      $endgroup$
      – robjohn
      Dec 13 '14 at 13:31



















    8












    $begingroup$

    Originally posted on the proofs without words post, here is a simple image that explains the derivative of $sin(x)$, which as we all know, is directly related to the limit at hand.



    enter image description here



    If one is not so convinced, take a look at the above picture and notice that if $upm h$ is in the first quadrant, then



    $$frac{sin(x+h)-sin(x)}h<cos(x)<frac{sin(x-h)-sin(x)}{-h}$$





    Notice that



    $$
    begin{align}frac{d}{dx}sin(x)&=lim_{hto0}frac{sin(x+h)-sin(x)}h\text{picture above}&=lim_{hto0}frac{sin(x)cos(h)+cos(x)sin(h)-sin(x)}h\cos(x)&=lim_{hto0}sin(x)frac{cos(h)-1}h+cos(x)frac{sin(h)}h\cos(0)&=lim_{hto0}frac{sin(h)}hend{align}
    $$






    share|cite|improve this answer











    $endgroup$









    • 2




      $begingroup$
      Note that your reason is circular: In order to prove the derivative of $sin$, we need to know the limit of $sin hover h.$ Attempting to prove the limit the other way round is counterproductive.
      $endgroup$
      – FUZxxl
      Nov 16 '16 at 0:47






    • 1




      $begingroup$
      @FUZxxl no actually, the whole point of this was a geometric proof of the fact.
      $endgroup$
      – Simply Beautiful Art
      Nov 16 '16 at 17:23










    • $begingroup$
      @FUZxxl If you are still unsatisfied about it, there is a whole post on the topic.
      $endgroup$
      – Simply Beautiful Art
      Nov 17 '16 at 0:44










    • $begingroup$
      @FUZxxl Since you still seem unhappy, I've added a better "squeeze theorem" type thing for the derivative of $sin(x)$.
      $endgroup$
      – Simply Beautiful Art
      Jan 5 '17 at 23:14



















    4












    $begingroup$

    Here is another approach.



    enter image description here(1)
    picture 2(2)



    In the large triangle, $$tan(theta)=frac{opp}{adg}=frac{z}{1}=z$$ So the triangle has height $$z=tan(theta)$$ and base $1$
    so it's area is $$Area(big triangle)=frac{1}{2}z=frac{1}{2}tan(theta)$$



    Next the sector area as a fraction of the entire circle, the sector is (see the right hand side of picture (1))$$frac{theta}{2pi}$$ of the entire circle so it's area is



    $$Area(sector)=frac{theta}{2pi}*(pi)(1)^2=frac{theta}{2}$$
    The triangle within the sector has height $y$. But $y=frac{y}{1}=frac{opp}{hyp}=sin(theta)$ so the small triangle has height $y=sin(theta)$ and base $1$ so its area is $$Area(small triangle)=frac{1}{2}y=frac{1}{2}sin(theta)$$
    Now we can use the sandwich theorem as $$Area(big triangle)geq Area(sector)geq Area(small triangle)$$



    using the equations we worked out above this becomes



    $$frac{tan(theta)}{2}geqfrac{theta}{2}geqfrac{sin(theta)}{2}$$ now multipliying through by two and using the fact that $$tan(theta)=frac{sin(theta)}{cos(theta)}$$ we get that
    $$frac{sin(theta)}{cos(theta)}geqthetageqsin(theta)$$
    taking reciprocals changes the inequalities so we have that
    $$frac{cos(theta)}{sin(theta)}leqfrac{1}{theta}leqfrac{1}{sin(theta)}$$
    now finally multiplying through by $sin(theta)$ we get $$cos(theta)leqfrac{sin(theta)}{theta}leq1$$
    Now $$lim limits_{theta to 0}cos(theta)=1$$ and$$lim limits_{theta to 0}1=1$$



    so by the sandwhich theorem $$lim limits_{theta to 0}frac{sin(theta)}{theta}=1$$ also. QED






    share|cite|improve this answer









    $endgroup$





















      2












      $begingroup$

      We may use a linear approximation for this limit. The following expression comes straight out of the equation of a line that is $y=mx+b$. Using the fact that $m=dfrac{d}{dx}[f(x)]$ $$f(x) approx f(a) + f'(a)(x-a)$$In this case $f(x)=sin x$ and $a=0$ $implies sin x approx f(0)+cos 0 (x-0)=x longleftrightarrow sin x approx x$. The following graph better illustrates this tangent line approximation.





      Since, $sin x approx x implies displaystyle lim_{x to 0} dfrac{sin x}{x}=1$. Quod Erat Demonstrandum






      share|cite|improve this answer









      $endgroup$









      • 2




        $begingroup$
        This is not a proof unless you can justify all these approximation steps.
        $endgroup$
        – FUZxxl
        Dec 4 '18 at 12:33










      • $begingroup$
        To find that $frac{mathrm{d}}{mathrm{d}x}sin(x)=cos(x)$ for the Taylor Series usually requires knowing the limit we are trying to find. In fact, $$left.frac{mathrm{d}}{mathrm{d}x}sin(x)right|_{x=0}=lim_{xto0}frac{sin(x)-sin(0)}{x-0}=lim_{xto0}frac{sin(x)}{x}$$
        $endgroup$
        – robjohn
        Dec 7 '18 at 20:20





















      1












      $begingroup$

      Here is a proof to those familiar with power series.



      The definition of $sin(x)$ is



      $$sin(x) = sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}h^{2k+1}$$



      Therefore we get



      $$begin{align} lim_{x to 0} frac{sin(x)}{x} &= lim_{x to 0} frac{sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}x^{2k+1}}{x} \&= lim_{x to 0} sum_{k=0}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 + lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 end{align}$$



      where we have used the fact that the power series $sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k}$ has radius of convergence $R=infty$ and therefore is continuous on $mathbb R$. This allows us to take the limit inside and we get



      $$lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} = sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} 0^{2k} = 0$$






      share|cite|improve this answer









      $endgroup$













      • $begingroup$
        Compare to this answer and this answer, which both use power series.
        $endgroup$
        – robjohn
        Jan 14 '18 at 19:48






      • 1




        $begingroup$
        The premise of the question is not to use power series.
        $endgroup$
        – FUZxxl
        Dec 4 '18 at 12:34



















      1












      $begingroup$

      We can also use Euler's formula to prove the limit:



      $$e^{ix} = cos x + isin x$$



      $$lim_{xto 0}dfrac{sin x}{x} = implies lim_{xto 0} dfrac{e^{ix}- e^{-ix}}{2i x}$$



      $$= lim_{xto 0} dfrac{e^{2ix}-1}{2ix}timesdfrac 1{ e^{ix} }= 1 times 1 = 1$$



      since:



      $lim_{f(x)to 0}dfrac{e^{f(x)}-1}{f(x)} = 1$






      share|cite|improve this answer









      $endgroup$












        protected by Community Mar 13 '15 at 19:29



        Thank you for your interest in this question.
        Because it has attracted low-quality or spam answers that had to be removed, posting an answer now requires 10 reputation on this site (the association bonus does not count).



        Would you like to answer one of these unanswered questions instead?














        22 Answers
        22






        active

        oldest

        votes








        22 Answers
        22






        active

        oldest

        votes









        active

        oldest

        votes






        active

        oldest

        votes









        479












        $begingroup$

        sinc and tanc at 0



        The area of $triangle ABC$ is $frac{1}{2}sin(x)$. The area of the colored wedge is $frac{1}{2}x$, and the area of $triangle ABD$ is $frac{1}{2}tan(x)$. By inclusion, we get
        $$
        frac{1}{2}tan(x)gefrac{1}{2}xgefrac{1}{2}sin(x)tag{1}
        $$
        Dividing $(1)$ by $frac{1}{2}sin(x)$ and taking reciprocals, we get
        $$
        cos(x)lefrac{sin(x)}{x}le1tag{2}
        $$
        Since $frac{sin(x)}{x}$ and $cos(x)$ are even functions, $(2)$ is valid for any non-zero $x$ between $-frac{pi}{2}$ and $frac{pi}{2}$. Furthermore, since $cos(x)$ is continuous near $0$ and $cos(0) = 1$, we get that
        $$
        lim_{xto0}frac{sin(x)}{x}=1tag{3}
        $$
        Also, dividing $(2)$ by $cos(x)$, we get that
        $$
        1lefrac{tan(x)}{x}lesec(x)tag{4}
        $$
        Since $sec(x)$ is continuous near $0$ and $sec(0) = 1$, we get that
        $$
        lim_{xto0}frac{tan(x)}{x}=1tag{5}
        $$






        share|cite|improve this answer











        $endgroup$









        • 71




          $begingroup$
          Do my homework! Note that $(1)$ says that for $0le xlefrac{pi}{2}$ we have $0lesin(x)le x$; therefore, $limlimits_{xto0^+}sin(x)=0$. Then $cos(x)=1-2sin^2(x/2)$ should finish the job.
          $endgroup$
          – robjohn
          Oct 23 '11 at 18:37








        • 3




          $begingroup$
          Thank you very much. I know that proverb, but I really wasn't able to find that out on my own.
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 18:41






        • 46




          $begingroup$
          From your comment, I wasn't expecting that you could find it on your own, but "Read the question!" seemed a bit rough around the edges.
          $endgroup$
          – robjohn
          Oct 23 '11 at 19:04






        • 22




          $begingroup$
          Sorry for that.
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 19:35






        • 6




          $begingroup$
          I think that your justification is slightly circular :). In an introductory calculus course, $(sin x,cos x)$ is probably defined to be the point we reach after traveling $x$ units counterclockwise along the unit circle from $(1,0)$.
          $endgroup$
          – Mike F
          Oct 23 '11 at 20:54


















        479












        $begingroup$

        sinc and tanc at 0



        The area of $triangle ABC$ is $frac{1}{2}sin(x)$. The area of the colored wedge is $frac{1}{2}x$, and the area of $triangle ABD$ is $frac{1}{2}tan(x)$. By inclusion, we get
        $$
        frac{1}{2}tan(x)gefrac{1}{2}xgefrac{1}{2}sin(x)tag{1}
        $$
        Dividing $(1)$ by $frac{1}{2}sin(x)$ and taking reciprocals, we get
        $$
        cos(x)lefrac{sin(x)}{x}le1tag{2}
        $$
        Since $frac{sin(x)}{x}$ and $cos(x)$ are even functions, $(2)$ is valid for any non-zero $x$ between $-frac{pi}{2}$ and $frac{pi}{2}$. Furthermore, since $cos(x)$ is continuous near $0$ and $cos(0) = 1$, we get that
        $$
        lim_{xto0}frac{sin(x)}{x}=1tag{3}
        $$
        Also, dividing $(2)$ by $cos(x)$, we get that
        $$
        1lefrac{tan(x)}{x}lesec(x)tag{4}
        $$
        Since $sec(x)$ is continuous near $0$ and $sec(0) = 1$, we get that
        $$
        lim_{xto0}frac{tan(x)}{x}=1tag{5}
        $$






        share|cite|improve this answer











        $endgroup$









        • 71




          $begingroup$
          Do my homework! Note that $(1)$ says that for $0le xlefrac{pi}{2}$ we have $0lesin(x)le x$; therefore, $limlimits_{xto0^+}sin(x)=0$. Then $cos(x)=1-2sin^2(x/2)$ should finish the job.
          $endgroup$
          – robjohn
          Oct 23 '11 at 18:37








        • 3




          $begingroup$
          Thank you very much. I know that proverb, but I really wasn't able to find that out on my own.
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 18:41






        • 46




          $begingroup$
          From your comment, I wasn't expecting that you could find it on your own, but "Read the question!" seemed a bit rough around the edges.
          $endgroup$
          – robjohn
          Oct 23 '11 at 19:04






        • 22




          $begingroup$
          Sorry for that.
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 19:35






        • 6




          $begingroup$
          I think that your justification is slightly circular :). In an introductory calculus course, $(sin x,cos x)$ is probably defined to be the point we reach after traveling $x$ units counterclockwise along the unit circle from $(1,0)$.
          $endgroup$
          – Mike F
          Oct 23 '11 at 20:54
















        479












        479








        479





        $begingroup$

        sinc and tanc at 0



        The area of $triangle ABC$ is $frac{1}{2}sin(x)$. The area of the colored wedge is $frac{1}{2}x$, and the area of $triangle ABD$ is $frac{1}{2}tan(x)$. By inclusion, we get
        $$
        frac{1}{2}tan(x)gefrac{1}{2}xgefrac{1}{2}sin(x)tag{1}
        $$
        Dividing $(1)$ by $frac{1}{2}sin(x)$ and taking reciprocals, we get
        $$
        cos(x)lefrac{sin(x)}{x}le1tag{2}
        $$
        Since $frac{sin(x)}{x}$ and $cos(x)$ are even functions, $(2)$ is valid for any non-zero $x$ between $-frac{pi}{2}$ and $frac{pi}{2}$. Furthermore, since $cos(x)$ is continuous near $0$ and $cos(0) = 1$, we get that
        $$
        lim_{xto0}frac{sin(x)}{x}=1tag{3}
        $$
        Also, dividing $(2)$ by $cos(x)$, we get that
        $$
        1lefrac{tan(x)}{x}lesec(x)tag{4}
        $$
        Since $sec(x)$ is continuous near $0$ and $sec(0) = 1$, we get that
        $$
        lim_{xto0}frac{tan(x)}{x}=1tag{5}
        $$






        share|cite|improve this answer











        $endgroup$



        sinc and tanc at 0



        The area of $triangle ABC$ is $frac{1}{2}sin(x)$. The area of the colored wedge is $frac{1}{2}x$, and the area of $triangle ABD$ is $frac{1}{2}tan(x)$. By inclusion, we get
        $$
        frac{1}{2}tan(x)gefrac{1}{2}xgefrac{1}{2}sin(x)tag{1}
        $$
        Dividing $(1)$ by $frac{1}{2}sin(x)$ and taking reciprocals, we get
        $$
        cos(x)lefrac{sin(x)}{x}le1tag{2}
        $$
        Since $frac{sin(x)}{x}$ and $cos(x)$ are even functions, $(2)$ is valid for any non-zero $x$ between $-frac{pi}{2}$ and $frac{pi}{2}$. Furthermore, since $cos(x)$ is continuous near $0$ and $cos(0) = 1$, we get that
        $$
        lim_{xto0}frac{sin(x)}{x}=1tag{3}
        $$
        Also, dividing $(2)$ by $cos(x)$, we get that
        $$
        1lefrac{tan(x)}{x}lesec(x)tag{4}
        $$
        Since $sec(x)$ is continuous near $0$ and $sec(0) = 1$, we get that
        $$
        lim_{xto0}frac{tan(x)}{x}=1tag{5}
        $$







        share|cite|improve this answer














        share|cite|improve this answer



        share|cite|improve this answer








        edited Jun 19 '13 at 14:37

























        answered Oct 23 '11 at 17:26









        robjohnrobjohn

        268k27308634




        268k27308634








        • 71




          $begingroup$
          Do my homework! Note that $(1)$ says that for $0le xlefrac{pi}{2}$ we have $0lesin(x)le x$; therefore, $limlimits_{xto0^+}sin(x)=0$. Then $cos(x)=1-2sin^2(x/2)$ should finish the job.
          $endgroup$
          – robjohn
          Oct 23 '11 at 18:37








        • 3




          $begingroup$
          Thank you very much. I know that proverb, but I really wasn't able to find that out on my own.
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 18:41






        • 46




          $begingroup$
          From your comment, I wasn't expecting that you could find it on your own, but "Read the question!" seemed a bit rough around the edges.
          $endgroup$
          – robjohn
          Oct 23 '11 at 19:04






        • 22




          $begingroup$
          Sorry for that.
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 19:35






        • 6




          $begingroup$
          I think that your justification is slightly circular :). In an introductory calculus course, $(sin x,cos x)$ is probably defined to be the point we reach after traveling $x$ units counterclockwise along the unit circle from $(1,0)$.
          $endgroup$
          – Mike F
          Oct 23 '11 at 20:54
















        • 71




          $begingroup$
          Do my homework! Note that $(1)$ says that for $0le xlefrac{pi}{2}$ we have $0lesin(x)le x$; therefore, $limlimits_{xto0^+}sin(x)=0$. Then $cos(x)=1-2sin^2(x/2)$ should finish the job.
          $endgroup$
          – robjohn
          Oct 23 '11 at 18:37








        • 3




          $begingroup$
          Thank you very much. I know that proverb, but I really wasn't able to find that out on my own.
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 18:41






        • 46




          $begingroup$
          From your comment, I wasn't expecting that you could find it on your own, but "Read the question!" seemed a bit rough around the edges.
          $endgroup$
          – robjohn
          Oct 23 '11 at 19:04






        • 22




          $begingroup$
          Sorry for that.
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 19:35






        • 6




          $begingroup$
          I think that your justification is slightly circular :). In an introductory calculus course, $(sin x,cos x)$ is probably defined to be the point we reach after traveling $x$ units counterclockwise along the unit circle from $(1,0)$.
          $endgroup$
          – Mike F
          Oct 23 '11 at 20:54










        71




        71




        $begingroup$
        Do my homework! Note that $(1)$ says that for $0le xlefrac{pi}{2}$ we have $0lesin(x)le x$; therefore, $limlimits_{xto0^+}sin(x)=0$. Then $cos(x)=1-2sin^2(x/2)$ should finish the job.
        $endgroup$
        – robjohn
        Oct 23 '11 at 18:37






        $begingroup$
        Do my homework! Note that $(1)$ says that for $0le xlefrac{pi}{2}$ we have $0lesin(x)le x$; therefore, $limlimits_{xto0^+}sin(x)=0$. Then $cos(x)=1-2sin^2(x/2)$ should finish the job.
        $endgroup$
        – robjohn
        Oct 23 '11 at 18:37






        3




        3




        $begingroup$
        Thank you very much. I know that proverb, but I really wasn't able to find that out on my own.
        $endgroup$
        – FUZxxl
        Oct 23 '11 at 18:41




        $begingroup$
        Thank you very much. I know that proverb, but I really wasn't able to find that out on my own.
        $endgroup$
        – FUZxxl
        Oct 23 '11 at 18:41




        46




        46




        $begingroup$
        From your comment, I wasn't expecting that you could find it on your own, but "Read the question!" seemed a bit rough around the edges.
        $endgroup$
        – robjohn
        Oct 23 '11 at 19:04




        $begingroup$
        From your comment, I wasn't expecting that you could find it on your own, but "Read the question!" seemed a bit rough around the edges.
        $endgroup$
        – robjohn
        Oct 23 '11 at 19:04




        22




        22




        $begingroup$
        Sorry for that.
        $endgroup$
        – FUZxxl
        Oct 23 '11 at 19:35




        $begingroup$
        Sorry for that.
        $endgroup$
        – FUZxxl
        Oct 23 '11 at 19:35




        6




        6




        $begingroup$
        I think that your justification is slightly circular :). In an introductory calculus course, $(sin x,cos x)$ is probably defined to be the point we reach after traveling $x$ units counterclockwise along the unit circle from $(1,0)$.
        $endgroup$
        – Mike F
        Oct 23 '11 at 20:54






        $begingroup$
        I think that your justification is slightly circular :). In an introductory calculus course, $(sin x,cos x)$ is probably defined to be the point we reach after traveling $x$ units counterclockwise along the unit circle from $(1,0)$.
        $endgroup$
        – Mike F
        Oct 23 '11 at 20:54













        119












        $begingroup$

        You should first prove that for $x > 0$ small that $sin x < x < tan x$. Then, dividing by $x$ you get
        $$
        { sin x over x} < 1
        $$
        and rearranging $1 < {tan x over x} = {sin x over x cos x }$
        $$
        cos x < {sin x over x}.
        $$
        Taking $x rightarrow 0^+$ you apply the squeeze theorem. For $x < 0$ and small use that $sin(-x) = -sin x$ so that $${sin(-x) over -x} = {sin x over x}.$$
        As far as why the first inequality I said is true, you can do this completely from triangles but I don't know how to draw the pictures here.



        diagram






        share|cite|improve this answer











        $endgroup$









        • 2




          $begingroup$
          But how to prove that $sin x<x<tan x$?
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 16:31










        • $begingroup$
          It is in the picture. The definition of radians makes the picture above true. Maybe that is worth mentioning: this limit explicitly depends on "$x$" being measured in radians.
          $endgroup$
          – tkr
          Oct 23 '11 at 16:33








        • 2




          $begingroup$
          This is a strange picture! Normally you want the $tan(theta)$ side to be parallel to the $sin(theta)$ side.
          $endgroup$
          – user641
          Oct 23 '11 at 17:36






        • 9




          $begingroup$
          If you make $tan(theta)$ parallel then you need to make the points $S$ and $Q$ the same. For whatever reason, this is the proof I like the most because it relates the tangent line at the point on the circle to the value we call "tangent". To each his own...
          $endgroup$
          – tkr
          Oct 23 '11 at 18:22








        • 2




          $begingroup$
          Even with this picture, you could use the area inclusion principle and argue that area of triangle OPQ < Area of sector OPQ < Area of triangle OPS. This translates into $sin theta < theta < tan theta$ which is the first line of this answer and hence, nothing strange about it.
          $endgroup$
          – Deepak Gupta
          Sep 2 '15 at 21:05
















        119












        $begingroup$

        You should first prove that for $x > 0$ small that $sin x < x < tan x$. Then, dividing by $x$ you get
        $$
        { sin x over x} < 1
        $$
        and rearranging $1 < {tan x over x} = {sin x over x cos x }$
        $$
        cos x < {sin x over x}.
        $$
        Taking $x rightarrow 0^+$ you apply the squeeze theorem. For $x < 0$ and small use that $sin(-x) = -sin x$ so that $${sin(-x) over -x} = {sin x over x}.$$
        As far as why the first inequality I said is true, you can do this completely from triangles but I don't know how to draw the pictures here.



        diagram






        share|cite|improve this answer











        $endgroup$









        • 2




          $begingroup$
          But how to prove that $sin x<x<tan x$?
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 16:31










        • $begingroup$
          It is in the picture. The definition of radians makes the picture above true. Maybe that is worth mentioning: this limit explicitly depends on "$x$" being measured in radians.
          $endgroup$
          – tkr
          Oct 23 '11 at 16:33








        • 2




          $begingroup$
          This is a strange picture! Normally you want the $tan(theta)$ side to be parallel to the $sin(theta)$ side.
          $endgroup$
          – user641
          Oct 23 '11 at 17:36






        • 9




          $begingroup$
          If you make $tan(theta)$ parallel then you need to make the points $S$ and $Q$ the same. For whatever reason, this is the proof I like the most because it relates the tangent line at the point on the circle to the value we call "tangent". To each his own...
          $endgroup$
          – tkr
          Oct 23 '11 at 18:22








        • 2




          $begingroup$
          Even with this picture, you could use the area inclusion principle and argue that area of triangle OPQ < Area of sector OPQ < Area of triangle OPS. This translates into $sin theta < theta < tan theta$ which is the first line of this answer and hence, nothing strange about it.
          $endgroup$
          – Deepak Gupta
          Sep 2 '15 at 21:05














        119












        119








        119





        $begingroup$

        You should first prove that for $x > 0$ small that $sin x < x < tan x$. Then, dividing by $x$ you get
        $$
        { sin x over x} < 1
        $$
        and rearranging $1 < {tan x over x} = {sin x over x cos x }$
        $$
        cos x < {sin x over x}.
        $$
        Taking $x rightarrow 0^+$ you apply the squeeze theorem. For $x < 0$ and small use that $sin(-x) = -sin x$ so that $${sin(-x) over -x} = {sin x over x}.$$
        As far as why the first inequality I said is true, you can do this completely from triangles but I don't know how to draw the pictures here.



        diagram






        share|cite|improve this answer











        $endgroup$



        You should first prove that for $x > 0$ small that $sin x < x < tan x$. Then, dividing by $x$ you get
        $$
        { sin x over x} < 1
        $$
        and rearranging $1 < {tan x over x} = {sin x over x cos x }$
        $$
        cos x < {sin x over x}.
        $$
        Taking $x rightarrow 0^+$ you apply the squeeze theorem. For $x < 0$ and small use that $sin(-x) = -sin x$ so that $${sin(-x) over -x} = {sin x over x}.$$
        As far as why the first inequality I said is true, you can do this completely from triangles but I don't know how to draw the pictures here.



        diagram







        share|cite|improve this answer














        share|cite|improve this answer



        share|cite|improve this answer








        edited Nov 20 '17 at 10:10









        Domates

        464415




        464415










        answered Oct 23 '11 at 16:28









        tkrtkr

        1,64811010




        1,64811010








        • 2




          $begingroup$
          But how to prove that $sin x<x<tan x$?
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 16:31










        • $begingroup$
          It is in the picture. The definition of radians makes the picture above true. Maybe that is worth mentioning: this limit explicitly depends on "$x$" being measured in radians.
          $endgroup$
          – tkr
          Oct 23 '11 at 16:33








        • 2




          $begingroup$
          This is a strange picture! Normally you want the $tan(theta)$ side to be parallel to the $sin(theta)$ side.
          $endgroup$
          – user641
          Oct 23 '11 at 17:36






        • 9




          $begingroup$
          If you make $tan(theta)$ parallel then you need to make the points $S$ and $Q$ the same. For whatever reason, this is the proof I like the most because it relates the tangent line at the point on the circle to the value we call "tangent". To each his own...
          $endgroup$
          – tkr
          Oct 23 '11 at 18:22








        • 2




          $begingroup$
          Even with this picture, you could use the area inclusion principle and argue that area of triangle OPQ < Area of sector OPQ < Area of triangle OPS. This translates into $sin theta < theta < tan theta$ which is the first line of this answer and hence, nothing strange about it.
          $endgroup$
          – Deepak Gupta
          Sep 2 '15 at 21:05














        • 2




          $begingroup$
          But how to prove that $sin x<x<tan x$?
          $endgroup$
          – FUZxxl
          Oct 23 '11 at 16:31










        • $begingroup$
          It is in the picture. The definition of radians makes the picture above true. Maybe that is worth mentioning: this limit explicitly depends on "$x$" being measured in radians.
          $endgroup$
          – tkr
          Oct 23 '11 at 16:33








        • 2




          $begingroup$
          This is a strange picture! Normally you want the $tan(theta)$ side to be parallel to the $sin(theta)$ side.
          $endgroup$
          – user641
          Oct 23 '11 at 17:36






        • 9




          $begingroup$
          If you make $tan(theta)$ parallel then you need to make the points $S$ and $Q$ the same. For whatever reason, this is the proof I like the most because it relates the tangent line at the point on the circle to the value we call "tangent". To each his own...
          $endgroup$
          – tkr
          Oct 23 '11 at 18:22








        • 2




          $begingroup$
          Even with this picture, you could use the area inclusion principle and argue that area of triangle OPQ < Area of sector OPQ < Area of triangle OPS. This translates into $sin theta < theta < tan theta$ which is the first line of this answer and hence, nothing strange about it.
          $endgroup$
          – Deepak Gupta
          Sep 2 '15 at 21:05








        2




        2




        $begingroup$
        But how to prove that $sin x<x<tan x$?
        $endgroup$
        – FUZxxl
        Oct 23 '11 at 16:31




        $begingroup$
        But how to prove that $sin x<x<tan x$?
        $endgroup$
        – FUZxxl
        Oct 23 '11 at 16:31












        $begingroup$
        It is in the picture. The definition of radians makes the picture above true. Maybe that is worth mentioning: this limit explicitly depends on "$x$" being measured in radians.
        $endgroup$
        – tkr
        Oct 23 '11 at 16:33






        $begingroup$
        It is in the picture. The definition of radians makes the picture above true. Maybe that is worth mentioning: this limit explicitly depends on "$x$" being measured in radians.
        $endgroup$
        – tkr
        Oct 23 '11 at 16:33






        2




        2




        $begingroup$
        This is a strange picture! Normally you want the $tan(theta)$ side to be parallel to the $sin(theta)$ side.
        $endgroup$
        – user641
        Oct 23 '11 at 17:36




        $begingroup$
        This is a strange picture! Normally you want the $tan(theta)$ side to be parallel to the $sin(theta)$ side.
        $endgroup$
        – user641
        Oct 23 '11 at 17:36




        9




        9




        $begingroup$
        If you make $tan(theta)$ parallel then you need to make the points $S$ and $Q$ the same. For whatever reason, this is the proof I like the most because it relates the tangent line at the point on the circle to the value we call "tangent". To each his own...
        $endgroup$
        – tkr
        Oct 23 '11 at 18:22






        $begingroup$
        If you make $tan(theta)$ parallel then you need to make the points $S$ and $Q$ the same. For whatever reason, this is the proof I like the most because it relates the tangent line at the point on the circle to the value we call "tangent". To each his own...
        $endgroup$
        – tkr
        Oct 23 '11 at 18:22






        2




        2




        $begingroup$
        Even with this picture, you could use the area inclusion principle and argue that area of triangle OPQ < Area of sector OPQ < Area of triangle OPS. This translates into $sin theta < theta < tan theta$ which is the first line of this answer and hence, nothing strange about it.
        $endgroup$
        – Deepak Gupta
        Sep 2 '15 at 21:05




        $begingroup$
        Even with this picture, you could use the area inclusion principle and argue that area of triangle OPQ < Area of sector OPQ < Area of triangle OPS. This translates into $sin theta < theta < tan theta$ which is the first line of this answer and hence, nothing strange about it.
        $endgroup$
        – Deepak Gupta
        Sep 2 '15 at 21:05











        88












        $begingroup$

        Usually calculus textbooks do this using geometric arguments followed by squeezing.



        Here's an Euler-esque way of looking at it---not a "proof" as that term is usually understood today, but still worth knowing about.



        Let $theta$ be the length of an arc along the circle of unit radius centered at $(0,0)$, from the point $(1,0)$ in a counterclockwise direction to some point $(costheta,sintheta)$ on the circle. Then of course $sintheta$ is the height of the latter point above the $x$-axis. Now imagine what happens if $theta$ is an infinitely small positive number. Then the arc is just an infinitely short vertical line, and the height of the endpoint above the $x$-axis is just the length of the arc. I.e. when $theta$ is an infinitely small number, then $sintheta$ is the same as $theta$. It follows that when $theta$ is an infinitely small nonzero number, then $dfrac{sintheta}{theta}=1$.



        That is how Euler viewed the matter. See his book on differential calculus.






        share|cite|improve this answer











        $endgroup$









        • 1




          $begingroup$
          I disagree with the recent edit to my answer and I have reverted to the previous version. "Infinitesmal" means "infinitely small". $qquad$
          $endgroup$
          – Michael Hardy
          Jan 21 '16 at 19:12










        • $begingroup$
          You read Euler's book ? Was it very difficult to read because of the notation and language of the time ?
          $endgroup$
          – user230452
          Feb 27 '16 at 4:03






        • 1




          $begingroup$
          @user230452 : Just some parts of it. I wouldn't say the differences in language and notation were the challenging part. $qquad$
          $endgroup$
          – Michael Hardy
          Feb 27 '16 at 18:18










        • $begingroup$
          Wouldn't one rather say that $frac{sin theta}{theta}$ is infinitely close to 1 if $theta$ is an infinitely small nonzero number?
          $endgroup$
          – Sven
          Apr 14 '17 at 13:05






        • 2




          $begingroup$
          @Sven : That is indeed how it's done in Robinson's "nonstandard analysis". There is another approach to rigorous infinitesimals in which it would be done the way Euler did it, saying that if $theta$ is infinitely small then $dfrac{sintheta}theta = 1.$ It's called "smooth infinitesimal analysis". In that approach, the square of an infinitesimal is $0$, so we have $$ frac{sintheta} theta = frac{theta - dfrac{theta^3} 6 + dfrac{theta^5}{120} - cdots cdots} theta = 1. $$ en.wikipedia.org/wiki/Smooth_infinitesimal_analysis $qquad$
          $endgroup$
          – Michael Hardy
          Apr 14 '17 at 18:48


















        88












        $begingroup$

        Usually calculus textbooks do this using geometric arguments followed by squeezing.



        Here's an Euler-esque way of looking at it---not a "proof" as that term is usually understood today, but still worth knowing about.



        Let $theta$ be the length of an arc along the circle of unit radius centered at $(0,0)$, from the point $(1,0)$ in a counterclockwise direction to some point $(costheta,sintheta)$ on the circle. Then of course $sintheta$ is the height of the latter point above the $x$-axis. Now imagine what happens if $theta$ is an infinitely small positive number. Then the arc is just an infinitely short vertical line, and the height of the endpoint above the $x$-axis is just the length of the arc. I.e. when $theta$ is an infinitely small number, then $sintheta$ is the same as $theta$. It follows that when $theta$ is an infinitely small nonzero number, then $dfrac{sintheta}{theta}=1$.



        That is how Euler viewed the matter. See his book on differential calculus.






        share|cite|improve this answer











        $endgroup$









        • 1




          $begingroup$
          I disagree with the recent edit to my answer and I have reverted to the previous version. "Infinitesmal" means "infinitely small". $qquad$
          $endgroup$
          – Michael Hardy
          Jan 21 '16 at 19:12










        • $begingroup$
          You read Euler's book ? Was it very difficult to read because of the notation and language of the time ?
          $endgroup$
          – user230452
          Feb 27 '16 at 4:03






        • 1




          $begingroup$
          @user230452 : Just some parts of it. I wouldn't say the differences in language and notation were the challenging part. $qquad$
          $endgroup$
          – Michael Hardy
          Feb 27 '16 at 18:18










        • $begingroup$
          Wouldn't one rather say that $frac{sin theta}{theta}$ is infinitely close to 1 if $theta$ is an infinitely small nonzero number?
          $endgroup$
          – Sven
          Apr 14 '17 at 13:05






        • 2




          $begingroup$
          @Sven : That is indeed how it's done in Robinson's "nonstandard analysis". There is another approach to rigorous infinitesimals in which it would be done the way Euler did it, saying that if $theta$ is infinitely small then $dfrac{sintheta}theta = 1.$ It's called "smooth infinitesimal analysis". In that approach, the square of an infinitesimal is $0$, so we have $$ frac{sintheta} theta = frac{theta - dfrac{theta^3} 6 + dfrac{theta^5}{120} - cdots cdots} theta = 1. $$ en.wikipedia.org/wiki/Smooth_infinitesimal_analysis $qquad$
          $endgroup$
          – Michael Hardy
          Apr 14 '17 at 18:48
















        88












        88








        88





        $begingroup$

        Usually calculus textbooks do this using geometric arguments followed by squeezing.



        Here's an Euler-esque way of looking at it---not a "proof" as that term is usually understood today, but still worth knowing about.



        Let $theta$ be the length of an arc along the circle of unit radius centered at $(0,0)$, from the point $(1,0)$ in a counterclockwise direction to some point $(costheta,sintheta)$ on the circle. Then of course $sintheta$ is the height of the latter point above the $x$-axis. Now imagine what happens if $theta$ is an infinitely small positive number. Then the arc is just an infinitely short vertical line, and the height of the endpoint above the $x$-axis is just the length of the arc. I.e. when $theta$ is an infinitely small number, then $sintheta$ is the same as $theta$. It follows that when $theta$ is an infinitely small nonzero number, then $dfrac{sintheta}{theta}=1$.



        That is how Euler viewed the matter. See his book on differential calculus.






        share|cite|improve this answer











        $endgroup$



        Usually calculus textbooks do this using geometric arguments followed by squeezing.



        Here's an Euler-esque way of looking at it---not a "proof" as that term is usually understood today, but still worth knowing about.



        Let $theta$ be the length of an arc along the circle of unit radius centered at $(0,0)$, from the point $(1,0)$ in a counterclockwise direction to some point $(costheta,sintheta)$ on the circle. Then of course $sintheta$ is the height of the latter point above the $x$-axis. Now imagine what happens if $theta$ is an infinitely small positive number. Then the arc is just an infinitely short vertical line, and the height of the endpoint above the $x$-axis is just the length of the arc. I.e. when $theta$ is an infinitely small number, then $sintheta$ is the same as $theta$. It follows that when $theta$ is an infinitely small nonzero number, then $dfrac{sintheta}{theta}=1$.



        That is how Euler viewed the matter. See his book on differential calculus.







        share|cite|improve this answer














        share|cite|improve this answer



        share|cite|improve this answer








        edited Jan 21 '16 at 19:11

























        answered Oct 23 '11 at 17:21









        Michael HardyMichael Hardy

        1




        1








        • 1




          $begingroup$
          I disagree with the recent edit to my answer and I have reverted to the previous version. "Infinitesmal" means "infinitely small". $qquad$
          $endgroup$
          – Michael Hardy
          Jan 21 '16 at 19:12










        • $begingroup$
          You read Euler's book ? Was it very difficult to read because of the notation and language of the time ?
          $endgroup$
          – user230452
          Feb 27 '16 at 4:03






        • 1




          $begingroup$
          @user230452 : Just some parts of it. I wouldn't say the differences in language and notation were the challenging part. $qquad$
          $endgroup$
          – Michael Hardy
          Feb 27 '16 at 18:18










        • $begingroup$
          Wouldn't one rather say that $frac{sin theta}{theta}$ is infinitely close to 1 if $theta$ is an infinitely small nonzero number?
          $endgroup$
          – Sven
          Apr 14 '17 at 13:05






        • 2




          $begingroup$
          @Sven : That is indeed how it's done in Robinson's "nonstandard analysis". There is another approach to rigorous infinitesimals in which it would be done the way Euler did it, saying that if $theta$ is infinitely small then $dfrac{sintheta}theta = 1.$ It's called "smooth infinitesimal analysis". In that approach, the square of an infinitesimal is $0$, so we have $$ frac{sintheta} theta = frac{theta - dfrac{theta^3} 6 + dfrac{theta^5}{120} - cdots cdots} theta = 1. $$ en.wikipedia.org/wiki/Smooth_infinitesimal_analysis $qquad$
          $endgroup$
          – Michael Hardy
          Apr 14 '17 at 18:48
















        • 1




          $begingroup$
          I disagree with the recent edit to my answer and I have reverted to the previous version. "Infinitesmal" means "infinitely small". $qquad$
          $endgroup$
          – Michael Hardy
          Jan 21 '16 at 19:12










        • $begingroup$
          You read Euler's book ? Was it very difficult to read because of the notation and language of the time ?
          $endgroup$
          – user230452
          Feb 27 '16 at 4:03






        • 1




          $begingroup$
          @user230452 : Just some parts of it. I wouldn't say the differences in language and notation were the challenging part. $qquad$
          $endgroup$
          – Michael Hardy
          Feb 27 '16 at 18:18










        • $begingroup$
          Wouldn't one rather say that $frac{sin theta}{theta}$ is infinitely close to 1 if $theta$ is an infinitely small nonzero number?
          $endgroup$
          – Sven
          Apr 14 '17 at 13:05






        • 2




          $begingroup$
          @Sven : That is indeed how it's done in Robinson's "nonstandard analysis". There is another approach to rigorous infinitesimals in which it would be done the way Euler did it, saying that if $theta$ is infinitely small then $dfrac{sintheta}theta = 1.$ It's called "smooth infinitesimal analysis". In that approach, the square of an infinitesimal is $0$, so we have $$ frac{sintheta} theta = frac{theta - dfrac{theta^3} 6 + dfrac{theta^5}{120} - cdots cdots} theta = 1. $$ en.wikipedia.org/wiki/Smooth_infinitesimal_analysis $qquad$
          $endgroup$
          – Michael Hardy
          Apr 14 '17 at 18:48










        1




        1




        $begingroup$
        I disagree with the recent edit to my answer and I have reverted to the previous version. "Infinitesmal" means "infinitely small". $qquad$
        $endgroup$
        – Michael Hardy
        Jan 21 '16 at 19:12




        $begingroup$
        I disagree with the recent edit to my answer and I have reverted to the previous version. "Infinitesmal" means "infinitely small". $qquad$
        $endgroup$
        – Michael Hardy
        Jan 21 '16 at 19:12












        $begingroup$
        You read Euler's book ? Was it very difficult to read because of the notation and language of the time ?
        $endgroup$
        – user230452
        Feb 27 '16 at 4:03




        $begingroup$
        You read Euler's book ? Was it very difficult to read because of the notation and language of the time ?
        $endgroup$
        – user230452
        Feb 27 '16 at 4:03




        1




        1




        $begingroup$
        @user230452 : Just some parts of it. I wouldn't say the differences in language and notation were the challenging part. $qquad$
        $endgroup$
        – Michael Hardy
        Feb 27 '16 at 18:18




        $begingroup$
        @user230452 : Just some parts of it. I wouldn't say the differences in language and notation were the challenging part. $qquad$
        $endgroup$
        – Michael Hardy
        Feb 27 '16 at 18:18












        $begingroup$
        Wouldn't one rather say that $frac{sin theta}{theta}$ is infinitely close to 1 if $theta$ is an infinitely small nonzero number?
        $endgroup$
        – Sven
        Apr 14 '17 at 13:05




        $begingroup$
        Wouldn't one rather say that $frac{sin theta}{theta}$ is infinitely close to 1 if $theta$ is an infinitely small nonzero number?
        $endgroup$
        – Sven
        Apr 14 '17 at 13:05




        2




        2




        $begingroup$
        @Sven : That is indeed how it's done in Robinson's "nonstandard analysis". There is another approach to rigorous infinitesimals in which it would be done the way Euler did it, saying that if $theta$ is infinitely small then $dfrac{sintheta}theta = 1.$ It's called "smooth infinitesimal analysis". In that approach, the square of an infinitesimal is $0$, so we have $$ frac{sintheta} theta = frac{theta - dfrac{theta^3} 6 + dfrac{theta^5}{120} - cdots cdots} theta = 1. $$ en.wikipedia.org/wiki/Smooth_infinitesimal_analysis $qquad$
        $endgroup$
        – Michael Hardy
        Apr 14 '17 at 18:48






        $begingroup$
        @Sven : That is indeed how it's done in Robinson's "nonstandard analysis". There is another approach to rigorous infinitesimals in which it would be done the way Euler did it, saying that if $theta$ is infinitely small then $dfrac{sintheta}theta = 1.$ It's called "smooth infinitesimal analysis". In that approach, the square of an infinitesimal is $0$, so we have $$ frac{sintheta} theta = frac{theta - dfrac{theta^3} 6 + dfrac{theta^5}{120} - cdots cdots} theta = 1. $$ en.wikipedia.org/wiki/Smooth_infinitesimal_analysis $qquad$
        $endgroup$
        – Michael Hardy
        Apr 14 '17 at 18:48













        73












        $begingroup$

        Look at this link:



        http://fatosmatematicos.blogspot.com/2010/08/provas-sem-palavras-parte-20.html



        Here is the picture I copied from that blog:



        Copy of the picture from the Fatos Matematicos blog






        share|cite|improve this answer











        $endgroup$













        • $begingroup$
          +1, nice site - What is the length BC?
          $endgroup$
          – NoChance
          Oct 24 '11 at 5:34






        • 11




          $begingroup$
          What is the argument for showing that $theta cos thetale sin theta$? The picture isn't immediately convincing.
          $endgroup$
          – Mark Viola
          Apr 15 '15 at 5:31










        • $begingroup$
          @Dr. MV: if you rescale the circle in $cos theta$ units, then the original $sin theta$ is a scaled $tan theta$ which is always longer than its arc.
          $endgroup$
          – Mattsteel
          May 30 '16 at 20:38
















        73












        $begingroup$

        Look at this link:



        http://fatosmatematicos.blogspot.com/2010/08/provas-sem-palavras-parte-20.html



        Here is the picture I copied from that blog:



        Copy of the picture from the Fatos Matematicos blog






        share|cite|improve this answer











        $endgroup$













        • $begingroup$
          +1, nice site - What is the length BC?
          $endgroup$
          – NoChance
          Oct 24 '11 at 5:34






        • 11




          $begingroup$
          What is the argument for showing that $theta cos thetale sin theta$? The picture isn't immediately convincing.
          $endgroup$
          – Mark Viola
          Apr 15 '15 at 5:31










        • $begingroup$
          @Dr. MV: if you rescale the circle in $cos theta$ units, then the original $sin theta$ is a scaled $tan theta$ which is always longer than its arc.
          $endgroup$
          – Mattsteel
          May 30 '16 at 20:38














        73












        73








        73





        $begingroup$

        Look at this link:



        http://fatosmatematicos.blogspot.com/2010/08/provas-sem-palavras-parte-20.html



        Here is the picture I copied from that blog:



        Copy of the picture from the Fatos Matematicos blog






        share|cite|improve this answer











        $endgroup$



        Look at this link:



        http://fatosmatematicos.blogspot.com/2010/08/provas-sem-palavras-parte-20.html



        Here is the picture I copied from that blog:



        Copy of the picture from the Fatos Matematicos blog







        share|cite|improve this answer














        share|cite|improve this answer



        share|cite|improve this answer








        edited Feb 17 '18 at 17:17









        elliot

        197




        197










        answered Oct 23 '11 at 21:15









        Paulo SérgioPaulo Sérgio

        87664




        87664












        • $begingroup$
          +1, nice site - What is the length BC?
          $endgroup$
          – NoChance
          Oct 24 '11 at 5:34






        • 11




          $begingroup$
          What is the argument for showing that $theta cos thetale sin theta$? The picture isn't immediately convincing.
          $endgroup$
          – Mark Viola
          Apr 15 '15 at 5:31










        • $begingroup$
          @Dr. MV: if you rescale the circle in $cos theta$ units, then the original $sin theta$ is a scaled $tan theta$ which is always longer than its arc.
          $endgroup$
          – Mattsteel
          May 30 '16 at 20:38


















        • $begingroup$
          +1, nice site - What is the length BC?
          $endgroup$
          – NoChance
          Oct 24 '11 at 5:34






        • 11




          $begingroup$
          What is the argument for showing that $theta cos thetale sin theta$? The picture isn't immediately convincing.
          $endgroup$
          – Mark Viola
          Apr 15 '15 at 5:31










        • $begingroup$
          @Dr. MV: if you rescale the circle in $cos theta$ units, then the original $sin theta$ is a scaled $tan theta$ which is always longer than its arc.
          $endgroup$
          – Mattsteel
          May 30 '16 at 20:38
















        $begingroup$
        +1, nice site - What is the length BC?
        $endgroup$
        – NoChance
        Oct 24 '11 at 5:34




        $begingroup$
        +1, nice site - What is the length BC?
        $endgroup$
        – NoChance
        Oct 24 '11 at 5:34




        11




        11




        $begingroup$
        What is the argument for showing that $theta cos thetale sin theta$? The picture isn't immediately convincing.
        $endgroup$
        – Mark Viola
        Apr 15 '15 at 5:31




        $begingroup$
        What is the argument for showing that $theta cos thetale sin theta$? The picture isn't immediately convincing.
        $endgroup$
        – Mark Viola
        Apr 15 '15 at 5:31












        $begingroup$
        @Dr. MV: if you rescale the circle in $cos theta$ units, then the original $sin theta$ is a scaled $tan theta$ which is always longer than its arc.
        $endgroup$
        – Mattsteel
        May 30 '16 at 20:38




        $begingroup$
        @Dr. MV: if you rescale the circle in $cos theta$ units, then the original $sin theta$ is a scaled $tan theta$ which is always longer than its arc.
        $endgroup$
        – Mattsteel
        May 30 '16 at 20:38











        49












        $begingroup$

        Here you may see an elementary approach that starts from a very interesting result, see this problem. All you need is a bit of imagination. When we take $lim_{nrightarrowinfty} frac{nsin(frac{pi}{n})}{1+sin(frac{pi}{n})}$ we may notice that we have infinitely many circles surrounding the unit circle with infinitely small diameters that lastly perfectly approximate the length of the unit circle when having it there infinity times . Therefore when multiplying n by the radius under the limit to infinity we get π. Let's denote $frac{pi}{n}$ by x.



        $$lim_{xrightarrow0}frac{pisin(x)}{x(1+sin(x))}={pi}Rightarrowlim_{xrightarrow0}frac{sin(x)}{x(1+sin(x))}=1Rightarrowlim_{xrightarrow0}frac{sin(x)}{x}=1$$



        The proof is complete.






        share|cite|improve this answer











        $endgroup$









        • 1




          $begingroup$
          this assumes the a priori knowledge of the existence of the limit $lim_{xto 0}frac{sin(x)}{x}$, how would you go about proving it before hand? (+1) by the way for the original and alternative proof
          $endgroup$
          – user153330
          Feb 24 '16 at 14:07
















        49












        $begingroup$

        Here you may see an elementary approach that starts from a very interesting result, see this problem. All you need is a bit of imagination. When we take $lim_{nrightarrowinfty} frac{nsin(frac{pi}{n})}{1+sin(frac{pi}{n})}$ we may notice that we have infinitely many circles surrounding the unit circle with infinitely small diameters that lastly perfectly approximate the length of the unit circle when having it there infinity times . Therefore when multiplying n by the radius under the limit to infinity we get π. Let's denote $frac{pi}{n}$ by x.



        $$lim_{xrightarrow0}frac{pisin(x)}{x(1+sin(x))}={pi}Rightarrowlim_{xrightarrow0}frac{sin(x)}{x(1+sin(x))}=1Rightarrowlim_{xrightarrow0}frac{sin(x)}{x}=1$$



        The proof is complete.






        share|cite|improve this answer











        $endgroup$









        • 1




          $begingroup$
          this assumes the a priori knowledge of the existence of the limit $lim_{xto 0}frac{sin(x)}{x}$, how would you go about proving it before hand? (+1) by the way for the original and alternative proof
          $endgroup$
          – user153330
          Feb 24 '16 at 14:07














        49












        49








        49





        $begingroup$

        Here you may see an elementary approach that starts from a very interesting result, see this problem. All you need is a bit of imagination. When we take $lim_{nrightarrowinfty} frac{nsin(frac{pi}{n})}{1+sin(frac{pi}{n})}$ we may notice that we have infinitely many circles surrounding the unit circle with infinitely small diameters that lastly perfectly approximate the length of the unit circle when having it there infinity times . Therefore when multiplying n by the radius under the limit to infinity we get π. Let's denote $frac{pi}{n}$ by x.



        $$lim_{xrightarrow0}frac{pisin(x)}{x(1+sin(x))}={pi}Rightarrowlim_{xrightarrow0}frac{sin(x)}{x(1+sin(x))}=1Rightarrowlim_{xrightarrow0}frac{sin(x)}{x}=1$$



        The proof is complete.






        share|cite|improve this answer











        $endgroup$



        Here you may see an elementary approach that starts from a very interesting result, see this problem. All you need is a bit of imagination. When we take $lim_{nrightarrowinfty} frac{nsin(frac{pi}{n})}{1+sin(frac{pi}{n})}$ we may notice that we have infinitely many circles surrounding the unit circle with infinitely small diameters that lastly perfectly approximate the length of the unit circle when having it there infinity times . Therefore when multiplying n by the radius under the limit to infinity we get π. Let's denote $frac{pi}{n}$ by x.



        $$lim_{xrightarrow0}frac{pisin(x)}{x(1+sin(x))}={pi}Rightarrowlim_{xrightarrow0}frac{sin(x)}{x(1+sin(x))}=1Rightarrowlim_{xrightarrow0}frac{sin(x)}{x}=1$$



        The proof is complete.







        share|cite|improve this answer














        share|cite|improve this answer



        share|cite|improve this answer








        edited Apr 13 '17 at 12:20









        Community

        1




        1










        answered Jun 7 '12 at 7:03









        user 1357113user 1357113

        22.4k878227




        22.4k878227








        • 1




          $begingroup$
          this assumes the a priori knowledge of the existence of the limit $lim_{xto 0}frac{sin(x)}{x}$, how would you go about proving it before hand? (+1) by the way for the original and alternative proof
          $endgroup$
          – user153330
          Feb 24 '16 at 14:07














        • 1




          $begingroup$
          this assumes the a priori knowledge of the existence of the limit $lim_{xto 0}frac{sin(x)}{x}$, how would you go about proving it before hand? (+1) by the way for the original and alternative proof
          $endgroup$
          – user153330
          Feb 24 '16 at 14:07








        1




        1




        $begingroup$
        this assumes the a priori knowledge of the existence of the limit $lim_{xto 0}frac{sin(x)}{x}$, how would you go about proving it before hand? (+1) by the way for the original and alternative proof
        $endgroup$
        – user153330
        Feb 24 '16 at 14:07




        $begingroup$
        this assumes the a priori knowledge of the existence of the limit $lim_{xto 0}frac{sin(x)}{x}$, how would you go about proving it before hand? (+1) by the way for the original and alternative proof
        $endgroup$
        – user153330
        Feb 24 '16 at 14:07











        49












        $begingroup$

        I claim that for $0<x<pi/2$ that the following holds
        $$sin x lt x lt tan x$$
        Figure 1

        In the diagram, we let $OC=OA=1$. In other words, $Arc:CA=x$ is an arc of a unit circle. The shortest distance from point $C$ to line $AO$ is line $CE=sin x$ (because $CEperp OA$). Another path from point $C$ to line $OA$ is arc $CA$ (which is longer than CE because it is not the shortest path). So we have at the very least
        $$sin x lt x$$
        Now we need to show that line $BA=tan x gt x$.

        Lines $AD$ and $CD$ are both tangent to arc $CA$. $CD+DA$ is longer than arc $CA$ because the set of points bound by sector $OCA$ is a subset of the set of points bound by quadrilateral $OCDA$, both of which are convex sets. This means that the perimeter of quadrilateral $OCDA$ must be longer than the perimeter of the sector $OCA$ (as per Archimedes, On the Sphere and Cylinder Book I). But both the sector and the quadrilateral both have sides $OC$ and $OA$, so we have
        $$CA=x<DC+DA$$
        But $BD>CD$ because it is the hypotenuse in $triangle BCD$ we have
        $$tan x = BA = BD+DAgt CD+DA gt CA=x gt sin x$$



        So we have
        $$sin x lt x lt tan x$$
        $$frac{sin x}{x} lt 1 lt frac{tan x}{x}=frac{sin x}{x}cdotsec x$$
        From this we can extract
        $$frac{sin x}{x} lt 1$$
        and
        $$1 lt frac{sin x}{x}cdotsec x$$
        $$cos x lt frac{sin x}{x}$$
        Putting these inequalities back together we have
        $$cos x lt frac{sin x}{x} lt 1$$



        Because $displaystylelim_{xto 0}cos x = 1$, by the squeeze theorem we have
        $$lim_{xto 0}frac{sin x}{x}=1$$






        share|cite|improve this answer









        $endgroup$









        • 1




          $begingroup$
          The reason that I chose an arc length proof is because most derivations that I've seen of the area of a circle (see mathopenref.com/circleareaderive.html) assume that $displaystylelim_{xto 0}frac{sin x}{x}=1$. Using an area based proof, to me, seems like putting the cart before the horse.
          $endgroup$
          – John Joy
          Nov 9 '14 at 15:14










        • $begingroup$
          How did Archimedes avoid assuming $frac{chord}{arc}rightarrow 1$ as $arc$ tends to zero in his argument about perimeters that you referred to? I asked the following question because this issue has kept haunting me for a while.
          $endgroup$
          – String
          Feb 5 '15 at 9:34






        • 2




          $begingroup$
          Actually, what Archimedes assumed was that if 2 curves (both concave in the same direction) have the same endpoints and one of the curves is strictly between a line segment with the same end points and the other curve, then the length of the enclosed curve is shorter than the length of enclosing curve. See axiom #2 archive.org/stream/worksofarchimede00arch#page/4/mode/2up
          $endgroup$
          – John Joy
          Feb 5 '15 at 14:17








        • 2




          $begingroup$
          OK, so he essentially axiomatizes the idea that among all curves connecting endpoints $A$ and $B$ satisfying some kind of regularity condition - concavity in the same direction, we have this inequality based on enclosure. This makes $arc(CA)<CD+DA$ in your figure merely equivalent to the assumption itself, but it makes sense and is very nicely put by Archimedes! Thank you for pointing me to that!
          $endgroup$
          – String
          Feb 5 '15 at 14:52






        • 2




          $begingroup$
          If you don't have a copy of Archimedes's book, you can approximate sector OCA with a polygon inside OCDA and use cut-the-knot.org/m/Geometry/PerimetersOfTwoConvexPolygons.shtml
          $endgroup$
          – CopyPasteIt
          Jun 6 '17 at 16:06
















        49












        $begingroup$

        I claim that for $0<x<pi/2$ that the following holds
        $$sin x lt x lt tan x$$
        Figure 1

        In the diagram, we let $OC=OA=1$. In other words, $Arc:CA=x$ is an arc of a unit circle. The shortest distance from point $C$ to line $AO$ is line $CE=sin x$ (because $CEperp OA$). Another path from point $C$ to line $OA$ is arc $CA$ (which is longer than CE because it is not the shortest path). So we have at the very least
        $$sin x lt x$$
        Now we need to show that line $BA=tan x gt x$.

        Lines $AD$ and $CD$ are both tangent to arc $CA$. $CD+DA$ is longer than arc $CA$ because the set of points bound by sector $OCA$ is a subset of the set of points bound by quadrilateral $OCDA$, both of which are convex sets. This means that the perimeter of quadrilateral $OCDA$ must be longer than the perimeter of the sector $OCA$ (as per Archimedes, On the Sphere and Cylinder Book I). But both the sector and the quadrilateral both have sides $OC$ and $OA$, so we have
        $$CA=x<DC+DA$$
        But $BD>CD$ because it is the hypotenuse in $triangle BCD$ we have
        $$tan x = BA = BD+DAgt CD+DA gt CA=x gt sin x$$



        So we have
        $$sin x lt x lt tan x$$
        $$frac{sin x}{x} lt 1 lt frac{tan x}{x}=frac{sin x}{x}cdotsec x$$
        From this we can extract
        $$frac{sin x}{x} lt 1$$
        and
        $$1 lt frac{sin x}{x}cdotsec x$$
        $$cos x lt frac{sin x}{x}$$
        Putting these inequalities back together we have
        $$cos x lt frac{sin x}{x} lt 1$$



        Because $displaystylelim_{xto 0}cos x = 1$, by the squeeze theorem we have
        $$lim_{xto 0}frac{sin x}{x}=1$$






        share|cite|improve this answer









        $endgroup$









        • 1




          $begingroup$
          The reason that I chose an arc length proof is because most derivations that I've seen of the area of a circle (see mathopenref.com/circleareaderive.html) assume that $displaystylelim_{xto 0}frac{sin x}{x}=1$. Using an area based proof, to me, seems like putting the cart before the horse.
          $endgroup$
          – John Joy
          Nov 9 '14 at 15:14










        • $begingroup$
          How did Archimedes avoid assuming $frac{chord}{arc}rightarrow 1$ as $arc$ tends to zero in his argument about perimeters that you referred to? I asked the following question because this issue has kept haunting me for a while.
          $endgroup$
          – String
          Feb 5 '15 at 9:34






        • 2




          $begingroup$
          Actually, what Archimedes assumed was that if 2 curves (both concave in the same direction) have the same endpoints and one of the curves is strictly between a line segment with the same end points and the other curve, then the length of the enclosed curve is shorter than the length of enclosing curve. See axiom #2 archive.org/stream/worksofarchimede00arch#page/4/mode/2up
          $endgroup$
          – John Joy
          Feb 5 '15 at 14:17








        • 2




          $begingroup$
          OK, so he essentially axiomatizes the idea that among all curves connecting endpoints $A$ and $B$ satisfying some kind of regularity condition - concavity in the same direction, we have this inequality based on enclosure. This makes $arc(CA)<CD+DA$ in your figure merely equivalent to the assumption itself, but it makes sense and is very nicely put by Archimedes! Thank you for pointing me to that!
          $endgroup$
          – String
          Feb 5 '15 at 14:52






        • 2




          $begingroup$
          If you don't have a copy of Archimedes's book, you can approximate sector OCA with a polygon inside OCDA and use cut-the-knot.org/m/Geometry/PerimetersOfTwoConvexPolygons.shtml
          $endgroup$
          – CopyPasteIt
          Jun 6 '17 at 16:06














        49












        49








        49





        $begingroup$

        I claim that for $0<x<pi/2$ that the following holds
        $$sin x lt x lt tan x$$
        Figure 1

        In the diagram, we let $OC=OA=1$. In other words, $Arc:CA=x$ is an arc of a unit circle. The shortest distance from point $C$ to line $AO$ is line $CE=sin x$ (because $CEperp OA$). Another path from point $C$ to line $OA$ is arc $CA$ (which is longer than CE because it is not the shortest path). So we have at the very least
        $$sin x lt x$$
        Now we need to show that line $BA=tan x gt x$.

        Lines $AD$ and $CD$ are both tangent to arc $CA$. $CD+DA$ is longer than arc $CA$ because the set of points bound by sector $OCA$ is a subset of the set of points bound by quadrilateral $OCDA$, both of which are convex sets. This means that the perimeter of quadrilateral $OCDA$ must be longer than the perimeter of the sector $OCA$ (as per Archimedes, On the Sphere and Cylinder Book I). But both the sector and the quadrilateral both have sides $OC$ and $OA$, so we have
        $$CA=x<DC+DA$$
        But $BD>CD$ because it is the hypotenuse in $triangle BCD$ we have
        $$tan x = BA = BD+DAgt CD+DA gt CA=x gt sin x$$



        So we have
        $$sin x lt x lt tan x$$
        $$frac{sin x}{x} lt 1 lt frac{tan x}{x}=frac{sin x}{x}cdotsec x$$
        From this we can extract
        $$frac{sin x}{x} lt 1$$
        and
        $$1 lt frac{sin x}{x}cdotsec x$$
        $$cos x lt frac{sin x}{x}$$
        Putting these inequalities back together we have
        $$cos x lt frac{sin x}{x} lt 1$$



        Because $displaystylelim_{xto 0}cos x = 1$, by the squeeze theorem we have
        $$lim_{xto 0}frac{sin x}{x}=1$$






        share|cite|improve this answer









        $endgroup$



        I claim that for $0<x<pi/2$ that the following holds
        $$sin x lt x lt tan x$$
        Figure 1

        In the diagram, we let $OC=OA=1$. In other words, $Arc:CA=x$ is an arc of a unit circle. The shortest distance from point $C$ to line $AO$ is line $CE=sin x$ (because $CEperp OA$). Another path from point $C$ to line $OA$ is arc $CA$ (which is longer than CE because it is not the shortest path). So we have at the very least
        $$sin x lt x$$
        Now we need to show that line $BA=tan x gt x$.

        Lines $AD$ and $CD$ are both tangent to arc $CA$. $CD+DA$ is longer than arc $CA$ because the set of points bound by sector $OCA$ is a subset of the set of points bound by quadrilateral $OCDA$, both of which are convex sets. This means that the perimeter of quadrilateral $OCDA$ must be longer than the perimeter of the sector $OCA$ (as per Archimedes, On the Sphere and Cylinder Book I). But both the sector and the quadrilateral both have sides $OC$ and $OA$, so we have
        $$CA=x<DC+DA$$
        But $BD>CD$ because it is the hypotenuse in $triangle BCD$ we have
        $$tan x = BA = BD+DAgt CD+DA gt CA=x gt sin x$$



        So we have
        $$sin x lt x lt tan x$$
        $$frac{sin x}{x} lt 1 lt frac{tan x}{x}=frac{sin x}{x}cdotsec x$$
        From this we can extract
        $$frac{sin x}{x} lt 1$$
        and
        $$1 lt frac{sin x}{x}cdotsec x$$
        $$cos x lt frac{sin x}{x}$$
        Putting these inequalities back together we have
        $$cos x lt frac{sin x}{x} lt 1$$



        Because $displaystylelim_{xto 0}cos x = 1$, by the squeeze theorem we have
        $$lim_{xto 0}frac{sin x}{x}=1$$







        share|cite|improve this answer












        share|cite|improve this answer



        share|cite|improve this answer










        answered Sep 7 '14 at 21:44









        John JoyJohn Joy

        6,31811727




        6,31811727








        • 1




          $begingroup$
          The reason that I chose an arc length proof is because most derivations that I've seen of the area of a circle (see mathopenref.com/circleareaderive.html) assume that $displaystylelim_{xto 0}frac{sin x}{x}=1$. Using an area based proof, to me, seems like putting the cart before the horse.
          $endgroup$
          – John Joy
          Nov 9 '14 at 15:14










        • $begingroup$
          How did Archimedes avoid assuming $frac{chord}{arc}rightarrow 1$ as $arc$ tends to zero in his argument about perimeters that you referred to? I asked the following question because this issue has kept haunting me for a while.
          $endgroup$
          – String
          Feb 5 '15 at 9:34






        • 2




          $begingroup$
          Actually, what Archimedes assumed was that if 2 curves (both concave in the same direction) have the same endpoints and one of the curves is strictly between a line segment with the same end points and the other curve, then the length of the enclosed curve is shorter than the length of enclosing curve. See axiom #2 archive.org/stream/worksofarchimede00arch#page/4/mode/2up
          $endgroup$
          – John Joy
          Feb 5 '15 at 14:17








        • 2




          $begingroup$
          OK, so he essentially axiomatizes the idea that among all curves connecting endpoints $A$ and $B$ satisfying some kind of regularity condition - concavity in the same direction, we have this inequality based on enclosure. This makes $arc(CA)<CD+DA$ in your figure merely equivalent to the assumption itself, but it makes sense and is very nicely put by Archimedes! Thank you for pointing me to that!
          $endgroup$
          – String
          Feb 5 '15 at 14:52






        • 2




          $begingroup$
          If you don't have a copy of Archimedes's book, you can approximate sector OCA with a polygon inside OCDA and use cut-the-knot.org/m/Geometry/PerimetersOfTwoConvexPolygons.shtml
          $endgroup$
          – CopyPasteIt
          Jun 6 '17 at 16:06














        • 1




          $begingroup$
          The reason that I chose an arc length proof is because most derivations that I've seen of the area of a circle (see mathopenref.com/circleareaderive.html) assume that $displaystylelim_{xto 0}frac{sin x}{x}=1$. Using an area based proof, to me, seems like putting the cart before the horse.
          $endgroup$
          – John Joy
          Nov 9 '14 at 15:14










        • $begingroup$
          How did Archimedes avoid assuming $frac{chord}{arc}rightarrow 1$ as $arc$ tends to zero in his argument about perimeters that you referred to? I asked the following question because this issue has kept haunting me for a while.
          $endgroup$
          – String
          Feb 5 '15 at 9:34






        • 2




          $begingroup$
          Actually, what Archimedes assumed was that if 2 curves (both concave in the same direction) have the same endpoints and one of the curves is strictly between a line segment with the same end points and the other curve, then the length of the enclosed curve is shorter than the length of enclosing curve. See axiom #2 archive.org/stream/worksofarchimede00arch#page/4/mode/2up
          $endgroup$
          – John Joy
          Feb 5 '15 at 14:17








        • 2




          $begingroup$
          OK, so he essentially axiomatizes the idea that among all curves connecting endpoints $A$ and $B$ satisfying some kind of regularity condition - concavity in the same direction, we have this inequality based on enclosure. This makes $arc(CA)<CD+DA$ in your figure merely equivalent to the assumption itself, but it makes sense and is very nicely put by Archimedes! Thank you for pointing me to that!
          $endgroup$
          – String
          Feb 5 '15 at 14:52






        • 2




          $begingroup$
          If you don't have a copy of Archimedes's book, you can approximate sector OCA with a polygon inside OCDA and use cut-the-knot.org/m/Geometry/PerimetersOfTwoConvexPolygons.shtml
          $endgroup$
          – CopyPasteIt
          Jun 6 '17 at 16:06








        1




        1




        $begingroup$
        The reason that I chose an arc length proof is because most derivations that I've seen of the area of a circle (see mathopenref.com/circleareaderive.html) assume that $displaystylelim_{xto 0}frac{sin x}{x}=1$. Using an area based proof, to me, seems like putting the cart before the horse.
        $endgroup$
        – John Joy
        Nov 9 '14 at 15:14




        $begingroup$
        The reason that I chose an arc length proof is because most derivations that I've seen of the area of a circle (see mathopenref.com/circleareaderive.html) assume that $displaystylelim_{xto 0}frac{sin x}{x}=1$. Using an area based proof, to me, seems like putting the cart before the horse.
        $endgroup$
        – John Joy
        Nov 9 '14 at 15:14












        $begingroup$
        How did Archimedes avoid assuming $frac{chord}{arc}rightarrow 1$ as $arc$ tends to zero in his argument about perimeters that you referred to? I asked the following question because this issue has kept haunting me for a while.
        $endgroup$
        – String
        Feb 5 '15 at 9:34




        $begingroup$
        How did Archimedes avoid assuming $frac{chord}{arc}rightarrow 1$ as $arc$ tends to zero in his argument about perimeters that you referred to? I asked the following question because this issue has kept haunting me for a while.
        $endgroup$
        – String
        Feb 5 '15 at 9:34




        2




        2




        $begingroup$
        Actually, what Archimedes assumed was that if 2 curves (both concave in the same direction) have the same endpoints and one of the curves is strictly between a line segment with the same end points and the other curve, then the length of the enclosed curve is shorter than the length of enclosing curve. See axiom #2 archive.org/stream/worksofarchimede00arch#page/4/mode/2up
        $endgroup$
        – John Joy
        Feb 5 '15 at 14:17






        $begingroup$
        Actually, what Archimedes assumed was that if 2 curves (both concave in the same direction) have the same endpoints and one of the curves is strictly between a line segment with the same end points and the other curve, then the length of the enclosed curve is shorter than the length of enclosing curve. See axiom #2 archive.org/stream/worksofarchimede00arch#page/4/mode/2up
        $endgroup$
        – John Joy
        Feb 5 '15 at 14:17






        2




        2




        $begingroup$
        OK, so he essentially axiomatizes the idea that among all curves connecting endpoints $A$ and $B$ satisfying some kind of regularity condition - concavity in the same direction, we have this inequality based on enclosure. This makes $arc(CA)<CD+DA$ in your figure merely equivalent to the assumption itself, but it makes sense and is very nicely put by Archimedes! Thank you for pointing me to that!
        $endgroup$
        – String
        Feb 5 '15 at 14:52




        $begingroup$
        OK, so he essentially axiomatizes the idea that among all curves connecting endpoints $A$ and $B$ satisfying some kind of regularity condition - concavity in the same direction, we have this inequality based on enclosure. This makes $arc(CA)<CD+DA$ in your figure merely equivalent to the assumption itself, but it makes sense and is very nicely put by Archimedes! Thank you for pointing me to that!
        $endgroup$
        – String
        Feb 5 '15 at 14:52




        2




        2




        $begingroup$
        If you don't have a copy of Archimedes's book, you can approximate sector OCA with a polygon inside OCDA and use cut-the-knot.org/m/Geometry/PerimetersOfTwoConvexPolygons.shtml
        $endgroup$
        – CopyPasteIt
        Jun 6 '17 at 16:06




        $begingroup$
        If you don't have a copy of Archimedes's book, you can approximate sector OCA with a polygon inside OCDA and use cut-the-knot.org/m/Geometry/PerimetersOfTwoConvexPolygons.shtml
        $endgroup$
        – CopyPasteIt
        Jun 6 '17 at 16:06











        46












        $begingroup$

        I am not sure if it counts as proof, but I have seen this done by a High Schooler.
        enter image description here



        In the given picture above,
        $displaystyle 2n text{ EJ} = 2nR sinleft( frac{pi}{n } right ) = text{ perimeter of polygon }$.



        $displaystyle lim_{nto infty }2nR sinleft( frac{pi}{n } right ) = lim_{nto infty } (text{ perimeter of polygon }) = 2 pi R implies lim_{nto infty}frac{sinleft( frac{pi}{n } right )}{left( frac{pi}{n } right )} = 1$ and let $frac{pi}{n} = x$.






        share|cite|improve this answer









        $endgroup$









        • 7




          $begingroup$
          This method is usually used to prove that the perimeter of a circle is $2pi R$ using the fact $limlimits_{xto 0}frac{sin x}{x}=1$
          $endgroup$
          – BPP
          Jul 20 '13 at 19:15






        • 3




          $begingroup$
          Santosh, how does one prove that the perimeter of the polygon converges to the $ 2pi$ ?
          $endgroup$
          – Imago
          Feb 5 '16 at 19:45










        • $begingroup$
          @Imago math.stackexchange.com/questions/720935/…
          $endgroup$
          – user301988
          Jun 6 '16 at 3:32
















        46












        $begingroup$

        I am not sure if it counts as proof, but I have seen this done by a High Schooler.
        enter image description here



        In the given picture above,
        $displaystyle 2n text{ EJ} = 2nR sinleft( frac{pi}{n } right ) = text{ perimeter of polygon }$.



        $displaystyle lim_{nto infty }2nR sinleft( frac{pi}{n } right ) = lim_{nto infty } (text{ perimeter of polygon }) = 2 pi R implies lim_{nto infty}frac{sinleft( frac{pi}{n } right )}{left( frac{pi}{n } right )} = 1$ and let $frac{pi}{n} = x$.






        share|cite|improve this answer









        $endgroup$









        • 7




          $begingroup$
          This method is usually used to prove that the perimeter of a circle is $2pi R$ using the fact $limlimits_{xto 0}frac{sin x}{x}=1$
          $endgroup$
          – BPP
          Jul 20 '13 at 19:15






        • 3




          $begingroup$
          Santosh, how does one prove that the perimeter of the polygon converges to the $ 2pi$ ?
          $endgroup$
          – Imago
          Feb 5 '16 at 19:45










        • $begingroup$
          @Imago math.stackexchange.com/questions/720935/…
          $endgroup$
          – user301988
          Jun 6 '16 at 3:32














        46












        46








        46





        $begingroup$

        I am not sure if it counts as proof, but I have seen this done by a High Schooler.
        enter image description here



        In the given picture above,
        $displaystyle 2n text{ EJ} = 2nR sinleft( frac{pi}{n } right ) = text{ perimeter of polygon }$.



        $displaystyle lim_{nto infty }2nR sinleft( frac{pi}{n } right ) = lim_{nto infty } (text{ perimeter of polygon }) = 2 pi R implies lim_{nto infty}frac{sinleft( frac{pi}{n } right )}{left( frac{pi}{n } right )} = 1$ and let $frac{pi}{n} = x$.






        share|cite|improve this answer









        $endgroup$



        I am not sure if it counts as proof, but I have seen this done by a High Schooler.
        enter image description here



        In the given picture above,
        $displaystyle 2n text{ EJ} = 2nR sinleft( frac{pi}{n } right ) = text{ perimeter of polygon }$.



        $displaystyle lim_{nto infty }2nR sinleft( frac{pi}{n } right ) = lim_{nto infty } (text{ perimeter of polygon }) = 2 pi R implies lim_{nto infty}frac{sinleft( frac{pi}{n } right )}{left( frac{pi}{n } right )} = 1$ and let $frac{pi}{n} = x$.







        share|cite|improve this answer












        share|cite|improve this answer



        share|cite|improve this answer










        answered May 16 '13 at 18:01









        Santosh LinkhaSantosh Linkha

        9,61422852




        9,61422852








        • 7




          $begingroup$
          This method is usually used to prove that the perimeter of a circle is $2pi R$ using the fact $limlimits_{xto 0}frac{sin x}{x}=1$
          $endgroup$
          – BPP
          Jul 20 '13 at 19:15






        • 3




          $begingroup$
          Santosh, how does one prove that the perimeter of the polygon converges to the $ 2pi$ ?
          $endgroup$
          – Imago
          Feb 5 '16 at 19:45










        • $begingroup$
          @Imago math.stackexchange.com/questions/720935/…
          $endgroup$
          – user301988
          Jun 6 '16 at 3:32














        • 7




          $begingroup$
          This method is usually used to prove that the perimeter of a circle is $2pi R$ using the fact $limlimits_{xto 0}frac{sin x}{x}=1$
          $endgroup$
          – BPP
          Jul 20 '13 at 19:15






        • 3




          $begingroup$
          Santosh, how does one prove that the perimeter of the polygon converges to the $ 2pi$ ?
          $endgroup$
          – Imago
          Feb 5 '16 at 19:45










        • $begingroup$
          @Imago math.stackexchange.com/questions/720935/…
          $endgroup$
          – user301988
          Jun 6 '16 at 3:32








        7




        7




        $begingroup$
        This method is usually used to prove that the perimeter of a circle is $2pi R$ using the fact $limlimits_{xto 0}frac{sin x}{x}=1$
        $endgroup$
        – BPP
        Jul 20 '13 at 19:15




        $begingroup$
        This method is usually used to prove that the perimeter of a circle is $2pi R$ using the fact $limlimits_{xto 0}frac{sin x}{x}=1$
        $endgroup$
        – BPP
        Jul 20 '13 at 19:15




        3




        3




        $begingroup$
        Santosh, how does one prove that the perimeter of the polygon converges to the $ 2pi$ ?
        $endgroup$
        – Imago
        Feb 5 '16 at 19:45




        $begingroup$
        Santosh, how does one prove that the perimeter of the polygon converges to the $ 2pi$ ?
        $endgroup$
        – Imago
        Feb 5 '16 at 19:45












        $begingroup$
        @Imago math.stackexchange.com/questions/720935/…
        $endgroup$
        – user301988
        Jun 6 '16 at 3:32




        $begingroup$
        @Imago math.stackexchange.com/questions/720935/…
        $endgroup$
        – user301988
        Jun 6 '16 at 3:32











        39












        $begingroup$

        Don't you feel strange about why most of the proofs are done with a figure? I've had this problem in the beginning, and realized after that this is due to the definition we use for the function $sin x$. Because the usual definition of $sin x$ we all study first in high schools depends on “classical geometry” and usually with a figure, you should depict out the figure and to make it clear.



        However, if you use other definitions of $sin x$ that are equivalent to the former, you'll find it more simple. For example,



        $$sin x = frac{x^1}{1!} - frac{x^3}{3!}+ frac{x^5}{5!} - cdots + cdots - cdots$$



        and hence



        $$frac{sin x}x = frac{x^0}{1!} - frac{x^2}{3!}+ frac{x^4}{5!} - cdots$$



        which obviously tends to $1$ as $x$ approaches 0.






        share|cite|improve this answer











        $endgroup$









        • 4




          $begingroup$
          Indeed, it's easy to see that this holds if one uses a series, but this question starts on the prerequisite that one does not use a series.
          $endgroup$
          – FUZxxl
          Mar 13 '15 at 17:54






        • 2




          $begingroup$
          And how do you know that the derivative of an infinte series is equal to the sum of the derivatives of each term?
          $endgroup$
          – steven gregory
          Apr 16 '16 at 4:32






        • 3




          $begingroup$
          @StevenGregory: Where does he use that? He only divides by $x$. What he does use is that $frac{1}{x}lim_{ntoinfty} sum_{k=0}^n a_k = lim_{ntoinfty}sum_{k=0}^n frac{a_k}{x}$
          $endgroup$
          – celtschk
          Sep 24 '16 at 9:36












        • $begingroup$
          @celschk starting with the infinite series as a 'definition' causes more problems than it solves. Now we have to prove that this sine behaves like the sine we learned in high school. This proof is compelling but it is not really a proof.
          $endgroup$
          – steven gregory
          Sep 24 '16 at 14:44
















        39












        $begingroup$

        Don't you feel strange about why most of the proofs are done with a figure? I've had this problem in the beginning, and realized after that this is due to the definition we use for the function $sin x$. Because the usual definition of $sin x$ we all study first in high schools depends on “classical geometry” and usually with a figure, you should depict out the figure and to make it clear.



        However, if you use other definitions of $sin x$ that are equivalent to the former, you'll find it more simple. For example,



        $$sin x = frac{x^1}{1!} - frac{x^3}{3!}+ frac{x^5}{5!} - cdots + cdots - cdots$$



        and hence



        $$frac{sin x}x = frac{x^0}{1!} - frac{x^2}{3!}+ frac{x^4}{5!} - cdots$$



        which obviously tends to $1$ as $x$ approaches 0.






        share|cite|improve this answer











        $endgroup$









        • 4




          $begingroup$
          Indeed, it's easy to see that this holds if one uses a series, but this question starts on the prerequisite that one does not use a series.
          $endgroup$
          – FUZxxl
          Mar 13 '15 at 17:54






        • 2




          $begingroup$
          And how do you know that the derivative of an infinte series is equal to the sum of the derivatives of each term?
          $endgroup$
          – steven gregory
          Apr 16 '16 at 4:32






        • 3




          $begingroup$
          @StevenGregory: Where does he use that? He only divides by $x$. What he does use is that $frac{1}{x}lim_{ntoinfty} sum_{k=0}^n a_k = lim_{ntoinfty}sum_{k=0}^n frac{a_k}{x}$
          $endgroup$
          – celtschk
          Sep 24 '16 at 9:36












        • $begingroup$
          @celschk starting with the infinite series as a 'definition' causes more problems than it solves. Now we have to prove that this sine behaves like the sine we learned in high school. This proof is compelling but it is not really a proof.
          $endgroup$
          – steven gregory
          Sep 24 '16 at 14:44














        39












        39








        39





        $begingroup$

        Don't you feel strange about why most of the proofs are done with a figure? I've had this problem in the beginning, and realized after that this is due to the definition we use for the function $sin x$. Because the usual definition of $sin x$ we all study first in high schools depends on “classical geometry” and usually with a figure, you should depict out the figure and to make it clear.



        However, if you use other definitions of $sin x$ that are equivalent to the former, you'll find it more simple. For example,



        $$sin x = frac{x^1}{1!} - frac{x^3}{3!}+ frac{x^5}{5!} - cdots + cdots - cdots$$



        and hence



        $$frac{sin x}x = frac{x^0}{1!} - frac{x^2}{3!}+ frac{x^4}{5!} - cdots$$



        which obviously tends to $1$ as $x$ approaches 0.






        share|cite|improve this answer











        $endgroup$



        Don't you feel strange about why most of the proofs are done with a figure? I've had this problem in the beginning, and realized after that this is due to the definition we use for the function $sin x$. Because the usual definition of $sin x$ we all study first in high schools depends on “classical geometry” and usually with a figure, you should depict out the figure and to make it clear.



        However, if you use other definitions of $sin x$ that are equivalent to the former, you'll find it more simple. For example,



        $$sin x = frac{x^1}{1!} - frac{x^3}{3!}+ frac{x^5}{5!} - cdots + cdots - cdots$$



        and hence



        $$frac{sin x}x = frac{x^0}{1!} - frac{x^2}{3!}+ frac{x^4}{5!} - cdots$$



        which obviously tends to $1$ as $x$ approaches 0.







        share|cite|improve this answer














        share|cite|improve this answer



        share|cite|improve this answer








        edited Mar 13 '15 at 17:57









        FUZxxl

        3,78052143




        3,78052143










        answered Mar 13 '15 at 17:06









        user223261user223261

        61155




        61155








        • 4




          $begingroup$
          Indeed, it's easy to see that this holds if one uses a series, but this question starts on the prerequisite that one does not use a series.
          $endgroup$
          – FUZxxl
          Mar 13 '15 at 17:54






        • 2




          $begingroup$
          And how do you know that the derivative of an infinte series is equal to the sum of the derivatives of each term?
          $endgroup$
          – steven gregory
          Apr 16 '16 at 4:32






        • 3




          $begingroup$
          @StevenGregory: Where does he use that? He only divides by $x$. What he does use is that $frac{1}{x}lim_{ntoinfty} sum_{k=0}^n a_k = lim_{ntoinfty}sum_{k=0}^n frac{a_k}{x}$
          $endgroup$
          – celtschk
          Sep 24 '16 at 9:36












        • $begingroup$
          @celschk starting with the infinite series as a 'definition' causes more problems than it solves. Now we have to prove that this sine behaves like the sine we learned in high school. This proof is compelling but it is not really a proof.
          $endgroup$
          – steven gregory
          Sep 24 '16 at 14:44














        • 4




          $begingroup$
          Indeed, it's easy to see that this holds if one uses a series, but this question starts on the prerequisite that one does not use a series.
          $endgroup$
          – FUZxxl
          Mar 13 '15 at 17:54






        • 2




          $begingroup$
          And how do you know that the derivative of an infinte series is equal to the sum of the derivatives of each term?
          $endgroup$
          – steven gregory
          Apr 16 '16 at 4:32






        • 3




          $begingroup$
          @StevenGregory: Where does he use that? He only divides by $x$. What he does use is that $frac{1}{x}lim_{ntoinfty} sum_{k=0}^n a_k = lim_{ntoinfty}sum_{k=0}^n frac{a_k}{x}$
          $endgroup$
          – celtschk
          Sep 24 '16 at 9:36












        • $begingroup$
          @celschk starting with the infinite series as a 'definition' causes more problems than it solves. Now we have to prove that this sine behaves like the sine we learned in high school. This proof is compelling but it is not really a proof.
          $endgroup$
          – steven gregory
          Sep 24 '16 at 14:44








        4




        4




        $begingroup$
        Indeed, it's easy to see that this holds if one uses a series, but this question starts on the prerequisite that one does not use a series.
        $endgroup$
        – FUZxxl
        Mar 13 '15 at 17:54




        $begingroup$
        Indeed, it's easy to see that this holds if one uses a series, but this question starts on the prerequisite that one does not use a series.
        $endgroup$
        – FUZxxl
        Mar 13 '15 at 17:54




        2




        2




        $begingroup$
        And how do you know that the derivative of an infinte series is equal to the sum of the derivatives of each term?
        $endgroup$
        – steven gregory
        Apr 16 '16 at 4:32




        $begingroup$
        And how do you know that the derivative of an infinte series is equal to the sum of the derivatives of each term?
        $endgroup$
        – steven gregory
        Apr 16 '16 at 4:32




        3




        3




        $begingroup$
        @StevenGregory: Where does he use that? He only divides by $x$. What he does use is that $frac{1}{x}lim_{ntoinfty} sum_{k=0}^n a_k = lim_{ntoinfty}sum_{k=0}^n frac{a_k}{x}$
        $endgroup$
        – celtschk
        Sep 24 '16 at 9:36






        $begingroup$
        @StevenGregory: Where does he use that? He only divides by $x$. What he does use is that $frac{1}{x}lim_{ntoinfty} sum_{k=0}^n a_k = lim_{ntoinfty}sum_{k=0}^n frac{a_k}{x}$
        $endgroup$
        – celtschk
        Sep 24 '16 at 9:36














        $begingroup$
        @celschk starting with the infinite series as a 'definition' causes more problems than it solves. Now we have to prove that this sine behaves like the sine we learned in high school. This proof is compelling but it is not really a proof.
        $endgroup$
        – steven gregory
        Sep 24 '16 at 14:44




        $begingroup$
        @celschk starting with the infinite series as a 'definition' causes more problems than it solves. Now we have to prove that this sine behaves like the sine we learned in high school. This proof is compelling but it is not really a proof.
        $endgroup$
        – steven gregory
        Sep 24 '16 at 14:44











        30












        $begingroup$

        Here's one more:
        $$
        lim_{x to 0} frac{sin x}{x}=lim_{x to 0} lim_{v to 0}frac{sin (x+v)-sin v}{x}\
        =lim_{v to 0} lim_{x to 0}frac{sin (x+v)-sin v}{x}=lim_{v to 0}sin'v=lim_{v to 0} cos v=1
        $$






        share|cite|improve this answer









        $endgroup$









        • 32




          $begingroup$
          Usually, this limit is used to compute the derivative of $sin(x)$.
          $endgroup$
          – robjohn
          Jul 20 '13 at 18:43
















        30












        $begingroup$

        Here's one more:
        $$
        lim_{x to 0} frac{sin x}{x}=lim_{x to 0} lim_{v to 0}frac{sin (x+v)-sin v}{x}\
        =lim_{v to 0} lim_{x to 0}frac{sin (x+v)-sin v}{x}=lim_{v to 0}sin'v=lim_{v to 0} cos v=1
        $$






        share|cite|improve this answer









        $endgroup$









        • 32




          $begingroup$
          Usually, this limit is used to compute the derivative of $sin(x)$.
          $endgroup$
          – robjohn
          Jul 20 '13 at 18:43














        30












        30








        30





        $begingroup$

        Here's one more:
        $$
        lim_{x to 0} frac{sin x}{x}=lim_{x to 0} lim_{v to 0}frac{sin (x+v)-sin v}{x}\
        =lim_{v to 0} lim_{x to 0}frac{sin (x+v)-sin v}{x}=lim_{v to 0}sin'v=lim_{v to 0} cos v=1
        $$






        share|cite|improve this answer









        $endgroup$



        Here's one more:
        $$
        lim_{x to 0} frac{sin x}{x}=lim_{x to 0} lim_{v to 0}frac{sin (x+v)-sin v}{x}\
        =lim_{v to 0} lim_{x to 0}frac{sin (x+v)-sin v}{x}=lim_{v to 0}sin'v=lim_{v to 0} cos v=1
        $$







        share|cite|improve this answer












        share|cite|improve this answer



        share|cite|improve this answer










        answered Jul 20 '13 at 18:37









        AlexAlex

        14.3k42134




        14.3k42134








        • 32




          $begingroup$
          Usually, this limit is used to compute the derivative of $sin(x)$.
          $endgroup$
          – robjohn
          Jul 20 '13 at 18:43














        • 32




          $begingroup$
          Usually, this limit is used to compute the derivative of $sin(x)$.
          $endgroup$
          – robjohn
          Jul 20 '13 at 18:43








        32




        32




        $begingroup$
        Usually, this limit is used to compute the derivative of $sin(x)$.
        $endgroup$
        – robjohn
        Jul 20 '13 at 18:43




        $begingroup$
        Usually, this limit is used to compute the derivative of $sin(x)$.
        $endgroup$
        – robjohn
        Jul 20 '13 at 18:43











        28












        $begingroup$

        It depends on your definition of the sine function. I would suggest checking out the geometric proof in ProofWiki.






        share|cite|improve this answer









        $endgroup$


















          28












          $begingroup$

          It depends on your definition of the sine function. I would suggest checking out the geometric proof in ProofWiki.






          share|cite|improve this answer









          $endgroup$
















            28












            28








            28





            $begingroup$

            It depends on your definition of the sine function. I would suggest checking out the geometric proof in ProofWiki.






            share|cite|improve this answer









            $endgroup$



            It depends on your definition of the sine function. I would suggest checking out the geometric proof in ProofWiki.







            share|cite|improve this answer












            share|cite|improve this answer



            share|cite|improve this answer










            answered Oct 23 '11 at 16:30









            Yuval FilmusYuval Filmus

            48.7k471145




            48.7k471145























                21












                $begingroup$

                The strategy is to find $frac{darcsin y}{dy}$ first. This can easily be done using the picture below.



                arcsin as area



                From the above picture, $arcsin y$ is twice the area of the orange bit. The area of the red bit is ${1 over 2}ysqrt{1-y^2}$. The area of the red bit plus the orange bit is $int_{0}^y sqrt{1-Y^2} dY$. So $$arcsin y = 2int_{0}^y sqrt{1-Y^2} dY - ysqrt{1-y^2}.$$ Differentiating with respect to $y$ gives $frac{darcsin y}{dy} = frac{1}{sqrt{1-y^2}}$. Using the theorem for the derivative of inverse functions yields $sin' theta = sqrt{1 - sin^2 theta} = cos theta$.



                (A similar thing can be done with the arc length definition of $arcsin$.)






                share|cite|improve this answer











                $endgroup$













                • $begingroup$
                  While this is indeed an interesting approach, integrals haven't been taught at the point where this limit is proved. Thank you for your answer though.
                  $endgroup$
                  – FUZxxl
                  Apr 11 '15 at 17:38
















                21












                $begingroup$

                The strategy is to find $frac{darcsin y}{dy}$ first. This can easily be done using the picture below.



                arcsin as area



                From the above picture, $arcsin y$ is twice the area of the orange bit. The area of the red bit is ${1 over 2}ysqrt{1-y^2}$. The area of the red bit plus the orange bit is $int_{0}^y sqrt{1-Y^2} dY$. So $$arcsin y = 2int_{0}^y sqrt{1-Y^2} dY - ysqrt{1-y^2}.$$ Differentiating with respect to $y$ gives $frac{darcsin y}{dy} = frac{1}{sqrt{1-y^2}}$. Using the theorem for the derivative of inverse functions yields $sin' theta = sqrt{1 - sin^2 theta} = cos theta$.



                (A similar thing can be done with the arc length definition of $arcsin$.)






                share|cite|improve this answer











                $endgroup$













                • $begingroup$
                  While this is indeed an interesting approach, integrals haven't been taught at the point where this limit is proved. Thank you for your answer though.
                  $endgroup$
                  – FUZxxl
                  Apr 11 '15 at 17:38














                21












                21








                21





                $begingroup$

                The strategy is to find $frac{darcsin y}{dy}$ first. This can easily be done using the picture below.



                arcsin as area



                From the above picture, $arcsin y$ is twice the area of the orange bit. The area of the red bit is ${1 over 2}ysqrt{1-y^2}$. The area of the red bit plus the orange bit is $int_{0}^y sqrt{1-Y^2} dY$. So $$arcsin y = 2int_{0}^y sqrt{1-Y^2} dY - ysqrt{1-y^2}.$$ Differentiating with respect to $y$ gives $frac{darcsin y}{dy} = frac{1}{sqrt{1-y^2}}$. Using the theorem for the derivative of inverse functions yields $sin' theta = sqrt{1 - sin^2 theta} = cos theta$.



                (A similar thing can be done with the arc length definition of $arcsin$.)






                share|cite|improve this answer











                $endgroup$



                The strategy is to find $frac{darcsin y}{dy}$ first. This can easily be done using the picture below.



                arcsin as area



                From the above picture, $arcsin y$ is twice the area of the orange bit. The area of the red bit is ${1 over 2}ysqrt{1-y^2}$. The area of the red bit plus the orange bit is $int_{0}^y sqrt{1-Y^2} dY$. So $$arcsin y = 2int_{0}^y sqrt{1-Y^2} dY - ysqrt{1-y^2}.$$ Differentiating with respect to $y$ gives $frac{darcsin y}{dy} = frac{1}{sqrt{1-y^2}}$. Using the theorem for the derivative of inverse functions yields $sin' theta = sqrt{1 - sin^2 theta} = cos theta$.



                (A similar thing can be done with the arc length definition of $arcsin$.)







                share|cite|improve this answer














                share|cite|improve this answer



                share|cite|improve this answer








                edited Mar 7 '18 at 16:46

























                answered Apr 11 '15 at 17:26









                man on laptopman on laptop

                5,74611438




                5,74611438












                • $begingroup$
                  While this is indeed an interesting approach, integrals haven't been taught at the point where this limit is proved. Thank you for your answer though.
                  $endgroup$
                  – FUZxxl
                  Apr 11 '15 at 17:38


















                • $begingroup$
                  While this is indeed an interesting approach, integrals haven't been taught at the point where this limit is proved. Thank you for your answer though.
                  $endgroup$
                  – FUZxxl
                  Apr 11 '15 at 17:38
















                $begingroup$
                While this is indeed an interesting approach, integrals haven't been taught at the point where this limit is proved. Thank you for your answer though.
                $endgroup$
                – FUZxxl
                Apr 11 '15 at 17:38




                $begingroup$
                While this is indeed an interesting approach, integrals haven't been taught at the point where this limit is proved. Thank you for your answer though.
                $endgroup$
                – FUZxxl
                Apr 11 '15 at 17:38











                20












                $begingroup$

                Let $f:{yinmathbb{R}:yneq 0}tomathbb{R}$ be a function defined by $f(x):=dfrac{sin x}{x}$ for all $xin {yinmathbb{R}:yneq 0}$.



                We have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$ if and only if for every $varepsilon>0$, there exists a $delta>0$ such that $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



                Let $varepsilon>0$ be an arbitrary real number.



                Note that $sin x=displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n+1}}{(2n+1)!}$.



                If $x neq 0$, we have $dfrac{sin x}{x}=$$displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}=1+$$displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}$.



                We thus have



                $|f(x)-1|=left|dfrac{sin x}{x}-1right|=left|displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}right|leq left|displaystylesum_{n=1}^{infty} dfrac{x^{2n}}{(2n+1)!}right|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|$



                Therefore we have



                $|f(x)-1|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|leq displaystyle sum_{n=1}^{infty} |x^{2n}|=sum_{n=1}^{infty}|x^2|^n$



                If $0<|x|<1$, then $0<|x^2|<1$, and the infinite series $displaystylesum_{n=1}^{infty}|x^2|^n$ converges to $dfrac{x^2}{1-x^2}$.



                Choose $delta:=sqrt{dfrac{varepsilon}{1+varepsilon}}$.
                Then $0<|x-0|<delta$ implies that $0<|x|<$$sqrt{dfrac{varepsilon}{1+varepsilon}}<1$, and hence $x^2<varepsilon-varepsilon x^2$. But $x^2<varepsilon-varepsilon x^2$ implies that $dfrac{x^2}{1-x^2}<varepsilon$.



                We therefore have $sum_{n=1}^{infty}|x^2|^n<varepsilon$ whenever $0<|x-0|<delta$. But since $|f(x)-1|leqdisplaystylesum_{n=1}^{infty}|x^2|^n$, we have $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



                Since $varepsilon$ was arbitrary, we have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$.






                share|cite|improve this answer









                $endgroup$













                • $begingroup$
                  The premise of the question was not to use power series.
                  $endgroup$
                  – FUZxxl
                  Dec 4 '18 at 12:35
















                20












                $begingroup$

                Let $f:{yinmathbb{R}:yneq 0}tomathbb{R}$ be a function defined by $f(x):=dfrac{sin x}{x}$ for all $xin {yinmathbb{R}:yneq 0}$.



                We have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$ if and only if for every $varepsilon>0$, there exists a $delta>0$ such that $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



                Let $varepsilon>0$ be an arbitrary real number.



                Note that $sin x=displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n+1}}{(2n+1)!}$.



                If $x neq 0$, we have $dfrac{sin x}{x}=$$displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}=1+$$displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}$.



                We thus have



                $|f(x)-1|=left|dfrac{sin x}{x}-1right|=left|displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}right|leq left|displaystylesum_{n=1}^{infty} dfrac{x^{2n}}{(2n+1)!}right|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|$



                Therefore we have



                $|f(x)-1|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|leq displaystyle sum_{n=1}^{infty} |x^{2n}|=sum_{n=1}^{infty}|x^2|^n$



                If $0<|x|<1$, then $0<|x^2|<1$, and the infinite series $displaystylesum_{n=1}^{infty}|x^2|^n$ converges to $dfrac{x^2}{1-x^2}$.



                Choose $delta:=sqrt{dfrac{varepsilon}{1+varepsilon}}$.
                Then $0<|x-0|<delta$ implies that $0<|x|<$$sqrt{dfrac{varepsilon}{1+varepsilon}}<1$, and hence $x^2<varepsilon-varepsilon x^2$. But $x^2<varepsilon-varepsilon x^2$ implies that $dfrac{x^2}{1-x^2}<varepsilon$.



                We therefore have $sum_{n=1}^{infty}|x^2|^n<varepsilon$ whenever $0<|x-0|<delta$. But since $|f(x)-1|leqdisplaystylesum_{n=1}^{infty}|x^2|^n$, we have $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



                Since $varepsilon$ was arbitrary, we have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$.






                share|cite|improve this answer









                $endgroup$













                • $begingroup$
                  The premise of the question was not to use power series.
                  $endgroup$
                  – FUZxxl
                  Dec 4 '18 at 12:35














                20












                20








                20





                $begingroup$

                Let $f:{yinmathbb{R}:yneq 0}tomathbb{R}$ be a function defined by $f(x):=dfrac{sin x}{x}$ for all $xin {yinmathbb{R}:yneq 0}$.



                We have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$ if and only if for every $varepsilon>0$, there exists a $delta>0$ such that $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



                Let $varepsilon>0$ be an arbitrary real number.



                Note that $sin x=displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n+1}}{(2n+1)!}$.



                If $x neq 0$, we have $dfrac{sin x}{x}=$$displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}=1+$$displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}$.



                We thus have



                $|f(x)-1|=left|dfrac{sin x}{x}-1right|=left|displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}right|leq left|displaystylesum_{n=1}^{infty} dfrac{x^{2n}}{(2n+1)!}right|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|$



                Therefore we have



                $|f(x)-1|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|leq displaystyle sum_{n=1}^{infty} |x^{2n}|=sum_{n=1}^{infty}|x^2|^n$



                If $0<|x|<1$, then $0<|x^2|<1$, and the infinite series $displaystylesum_{n=1}^{infty}|x^2|^n$ converges to $dfrac{x^2}{1-x^2}$.



                Choose $delta:=sqrt{dfrac{varepsilon}{1+varepsilon}}$.
                Then $0<|x-0|<delta$ implies that $0<|x|<$$sqrt{dfrac{varepsilon}{1+varepsilon}}<1$, and hence $x^2<varepsilon-varepsilon x^2$. But $x^2<varepsilon-varepsilon x^2$ implies that $dfrac{x^2}{1-x^2}<varepsilon$.



                We therefore have $sum_{n=1}^{infty}|x^2|^n<varepsilon$ whenever $0<|x-0|<delta$. But since $|f(x)-1|leqdisplaystylesum_{n=1}^{infty}|x^2|^n$, we have $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



                Since $varepsilon$ was arbitrary, we have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$.






                share|cite|improve this answer









                $endgroup$



                Let $f:{yinmathbb{R}:yneq 0}tomathbb{R}$ be a function defined by $f(x):=dfrac{sin x}{x}$ for all $xin {yinmathbb{R}:yneq 0}$.



                We have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$ if and only if for every $varepsilon>0$, there exists a $delta>0$ such that $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



                Let $varepsilon>0$ be an arbitrary real number.



                Note that $sin x=displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n+1}}{(2n+1)!}$.



                If $x neq 0$, we have $dfrac{sin x}{x}=$$displaystyle sum_{n=0}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}=1+$$displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}$.



                We thus have



                $|f(x)-1|=left|dfrac{sin x}{x}-1right|=left|displaystyle sum_{n=1}^{infty}(-1)^ndfrac{x^{2n}}{(2n+1)!}right|leq left|displaystylesum_{n=1}^{infty} dfrac{x^{2n}}{(2n+1)!}right|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|$



                Therefore we have



                $|f(x)-1|leq displaystylesum_{n=1}^{infty} left|dfrac{x^{2n}}{(2n+1)!}right|leq displaystyle sum_{n=1}^{infty} |x^{2n}|=sum_{n=1}^{infty}|x^2|^n$



                If $0<|x|<1$, then $0<|x^2|<1$, and the infinite series $displaystylesum_{n=1}^{infty}|x^2|^n$ converges to $dfrac{x^2}{1-x^2}$.



                Choose $delta:=sqrt{dfrac{varepsilon}{1+varepsilon}}$.
                Then $0<|x-0|<delta$ implies that $0<|x|<$$sqrt{dfrac{varepsilon}{1+varepsilon}}<1$, and hence $x^2<varepsilon-varepsilon x^2$. But $x^2<varepsilon-varepsilon x^2$ implies that $dfrac{x^2}{1-x^2}<varepsilon$.



                We therefore have $sum_{n=1}^{infty}|x^2|^n<varepsilon$ whenever $0<|x-0|<delta$. But since $|f(x)-1|leqdisplaystylesum_{n=1}^{infty}|x^2|^n$, we have $|f(x)-1|<varepsilon$ whenever $0<|x-0|<delta$.



                Since $varepsilon$ was arbitrary, we have $displaystylelim_{x to 0}dfrac{sin x}{x}=1$.







                share|cite|improve this answer












                share|cite|improve this answer



                share|cite|improve this answer










                answered Oct 10 '16 at 14:55









                Supreeth NarasimhaswamySupreeth Narasimhaswamy

                648514




                648514












                • $begingroup$
                  The premise of the question was not to use power series.
                  $endgroup$
                  – FUZxxl
                  Dec 4 '18 at 12:35


















                • $begingroup$
                  The premise of the question was not to use power series.
                  $endgroup$
                  – FUZxxl
                  Dec 4 '18 at 12:35
















                $begingroup$
                The premise of the question was not to use power series.
                $endgroup$
                – FUZxxl
                Dec 4 '18 at 12:35




                $begingroup$
                The premise of the question was not to use power series.
                $endgroup$
                – FUZxxl
                Dec 4 '18 at 12:35











                17












                $begingroup$


                Here is a different approach that uses the integral definition of the arcsine function. We will deduce the limit of interest without appeal to geometry or differential calculus.



                Instead, we only rely on elementary analysis of continuous functions and their inverses along with simple properties of the Riemann integral. To that end, we now proceed.






                We define the sine function, $sin(x)$, as the inverse function of the function $f(x)$ given by




                $$bbox[5px,border:2px solid #C0A000]{f(x)=int_0^x frac{1}{sqrt{1-t^2}},dt }tag 1$$




                for $|x|< 1$.




                NOTE: It can be shown that the sine function defined as the inverse of $f(x)$ given in $(1)$ has all of the familiar properties that characterize the circular function $sin(x)$.




                It is straightforward to show that since $frac{1}{sqrt{1-t^2}}$ is positive and continuous for $tin (-1,1)$, $f(x)$ is continuous and strictly increasing for $xin (-1,1)$ with $displaystylelim_{xto 0}f(x)=f(0)=0$.



                Therefore, since $f$ is continuous and strictly increasing, its inverse function, $sin(x)$, exists and is also continuous and strictly increasing with $displaystyle lim_{xto 0}sin(x)=sin(0)=0$.





                From $(1)$, we have the bounds (SEE HERE)




                $$bbox[5px,border:2px solid #C0A000]{1 le frac{f(x)}xle frac{1}{sqrt{1-x^2}}} tag 2$$




                for $xin (-1,1)$, whence applying the squeeze theorem to $(2)$ yields



                $$lim_{xto 0}frac{f(x)}{x}=1 tag 3$$





                Finally, let $y=f(x)$ so that $x=sin(y)$. As $xto 0$, $yto 0$ and we can write $(3)$ as



                $$lim_{yto 0}frac{y}{sin(y)}=1$$



                from which we have




                $$bbox[5px,border:2px solid #C0A000]{lim_{yto 0}frac{sin(y)}{y}=1}$$




                as was to be shown!






                NOTE:



                We can deduce the following set of useful inequalities from $(2)$. We let $x=sin(theta)$ and restrict $x$ so that $xin [0,1)$. In addition, we define new functions, $cos(theta)=sqrt{1-sin^2(theta)}$ and $tan(theta)=sin(theta)/cos(theta)$.



                Then, we have from $(2)$



                $$bbox[5px,border:2px solid #C0A000]{ycos(y)le sin(y)le yle tan(y)} $$



                which are the familiar inequalities often introduced in an introductory geometry or trigonometry course.







                share|cite|improve this answer











                $endgroup$









                • 1




                  $begingroup$
                  This is similar to my answer. You're formalizing the arc-length definition of $sin$ in the following three steps. 1) Defining $arcsin$ geometrically 2) Expressing that informal definition using integration 3) Defining $sin$ as the inverse of $arcsin$. Also, where you used the sqeeuze theorem, you could equally have used the Fundamental Theorem of Calculus and the Inverse Function Theorem. Then you'd get my answer
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 10:16








                • 1




                  $begingroup$
                  Another difference is you used the arc-length definition and I used the area definition
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 10:19






                • 1




                  $begingroup$
                  @ogogmad I don't use any geometrical argument. A function $f$ is defined without reference to the arc length interpretation. Both $f$ and its inverse has properties. It can be shown that the inverse function is the well-known sine function. And while one can apply the FOC instead of the squeeze theorem, there is no advantage in doing so. So, why on earth are you leaving these comments?
                  $endgroup$
                  – Mark Viola
                  Feb 5 '18 at 14:50






                • 1




                  $begingroup$
                  Because I'm pointing out that your abstract definitions make geometric sense, whether or not your realize it. I'm also pointing out that your squeeze argument is just a special case of FOC
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 16:43






                • 1




                  $begingroup$
                  In fact, all of the arguments above are morally the same, AFAICT. The squeeze arguments are just unpacking the proof of FOC on a special case. The FOC itself is proved using a very similar squeeze argument to the one people are using. The areas of the sectors are simply the integrals that are being differentiated. In tkr's argument, he's instead differentiating an arclength integral, which gives the same result anyway. There's the remaining feature that our two answers uses $arcsin$ while everyone else's uses $sin$; they may be implicitly unpacking the proof of the inverse function theorem.
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 17:01
















                17












                $begingroup$


                Here is a different approach that uses the integral definition of the arcsine function. We will deduce the limit of interest without appeal to geometry or differential calculus.



                Instead, we only rely on elementary analysis of continuous functions and their inverses along with simple properties of the Riemann integral. To that end, we now proceed.






                We define the sine function, $sin(x)$, as the inverse function of the function $f(x)$ given by




                $$bbox[5px,border:2px solid #C0A000]{f(x)=int_0^x frac{1}{sqrt{1-t^2}},dt }tag 1$$




                for $|x|< 1$.




                NOTE: It can be shown that the sine function defined as the inverse of $f(x)$ given in $(1)$ has all of the familiar properties that characterize the circular function $sin(x)$.




                It is straightforward to show that since $frac{1}{sqrt{1-t^2}}$ is positive and continuous for $tin (-1,1)$, $f(x)$ is continuous and strictly increasing for $xin (-1,1)$ with $displaystylelim_{xto 0}f(x)=f(0)=0$.



                Therefore, since $f$ is continuous and strictly increasing, its inverse function, $sin(x)$, exists and is also continuous and strictly increasing with $displaystyle lim_{xto 0}sin(x)=sin(0)=0$.





                From $(1)$, we have the bounds (SEE HERE)




                $$bbox[5px,border:2px solid #C0A000]{1 le frac{f(x)}xle frac{1}{sqrt{1-x^2}}} tag 2$$




                for $xin (-1,1)$, whence applying the squeeze theorem to $(2)$ yields



                $$lim_{xto 0}frac{f(x)}{x}=1 tag 3$$





                Finally, let $y=f(x)$ so that $x=sin(y)$. As $xto 0$, $yto 0$ and we can write $(3)$ as



                $$lim_{yto 0}frac{y}{sin(y)}=1$$



                from which we have




                $$bbox[5px,border:2px solid #C0A000]{lim_{yto 0}frac{sin(y)}{y}=1}$$




                as was to be shown!






                NOTE:



                We can deduce the following set of useful inequalities from $(2)$. We let $x=sin(theta)$ and restrict $x$ so that $xin [0,1)$. In addition, we define new functions, $cos(theta)=sqrt{1-sin^2(theta)}$ and $tan(theta)=sin(theta)/cos(theta)$.



                Then, we have from $(2)$



                $$bbox[5px,border:2px solid #C0A000]{ycos(y)le sin(y)le yle tan(y)} $$



                which are the familiar inequalities often introduced in an introductory geometry or trigonometry course.







                share|cite|improve this answer











                $endgroup$









                • 1




                  $begingroup$
                  This is similar to my answer. You're formalizing the arc-length definition of $sin$ in the following three steps. 1) Defining $arcsin$ geometrically 2) Expressing that informal definition using integration 3) Defining $sin$ as the inverse of $arcsin$. Also, where you used the sqeeuze theorem, you could equally have used the Fundamental Theorem of Calculus and the Inverse Function Theorem. Then you'd get my answer
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 10:16








                • 1




                  $begingroup$
                  Another difference is you used the arc-length definition and I used the area definition
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 10:19






                • 1




                  $begingroup$
                  @ogogmad I don't use any geometrical argument. A function $f$ is defined without reference to the arc length interpretation. Both $f$ and its inverse has properties. It can be shown that the inverse function is the well-known sine function. And while one can apply the FOC instead of the squeeze theorem, there is no advantage in doing so. So, why on earth are you leaving these comments?
                  $endgroup$
                  – Mark Viola
                  Feb 5 '18 at 14:50






                • 1




                  $begingroup$
                  Because I'm pointing out that your abstract definitions make geometric sense, whether or not your realize it. I'm also pointing out that your squeeze argument is just a special case of FOC
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 16:43






                • 1




                  $begingroup$
                  In fact, all of the arguments above are morally the same, AFAICT. The squeeze arguments are just unpacking the proof of FOC on a special case. The FOC itself is proved using a very similar squeeze argument to the one people are using. The areas of the sectors are simply the integrals that are being differentiated. In tkr's argument, he's instead differentiating an arclength integral, which gives the same result anyway. There's the remaining feature that our two answers uses $arcsin$ while everyone else's uses $sin$; they may be implicitly unpacking the proof of the inverse function theorem.
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 17:01














                17












                17








                17





                $begingroup$


                Here is a different approach that uses the integral definition of the arcsine function. We will deduce the limit of interest without appeal to geometry or differential calculus.



                Instead, we only rely on elementary analysis of continuous functions and their inverses along with simple properties of the Riemann integral. To that end, we now proceed.






                We define the sine function, $sin(x)$, as the inverse function of the function $f(x)$ given by




                $$bbox[5px,border:2px solid #C0A000]{f(x)=int_0^x frac{1}{sqrt{1-t^2}},dt }tag 1$$




                for $|x|< 1$.




                NOTE: It can be shown that the sine function defined as the inverse of $f(x)$ given in $(1)$ has all of the familiar properties that characterize the circular function $sin(x)$.




                It is straightforward to show that since $frac{1}{sqrt{1-t^2}}$ is positive and continuous for $tin (-1,1)$, $f(x)$ is continuous and strictly increasing for $xin (-1,1)$ with $displaystylelim_{xto 0}f(x)=f(0)=0$.



                Therefore, since $f$ is continuous and strictly increasing, its inverse function, $sin(x)$, exists and is also continuous and strictly increasing with $displaystyle lim_{xto 0}sin(x)=sin(0)=0$.





                From $(1)$, we have the bounds (SEE HERE)




                $$bbox[5px,border:2px solid #C0A000]{1 le frac{f(x)}xle frac{1}{sqrt{1-x^2}}} tag 2$$




                for $xin (-1,1)$, whence applying the squeeze theorem to $(2)$ yields



                $$lim_{xto 0}frac{f(x)}{x}=1 tag 3$$





                Finally, let $y=f(x)$ so that $x=sin(y)$. As $xto 0$, $yto 0$ and we can write $(3)$ as



                $$lim_{yto 0}frac{y}{sin(y)}=1$$



                from which we have




                $$bbox[5px,border:2px solid #C0A000]{lim_{yto 0}frac{sin(y)}{y}=1}$$




                as was to be shown!






                NOTE:



                We can deduce the following set of useful inequalities from $(2)$. We let $x=sin(theta)$ and restrict $x$ so that $xin [0,1)$. In addition, we define new functions, $cos(theta)=sqrt{1-sin^2(theta)}$ and $tan(theta)=sin(theta)/cos(theta)$.



                Then, we have from $(2)$



                $$bbox[5px,border:2px solid #C0A000]{ycos(y)le sin(y)le yle tan(y)} $$



                which are the familiar inequalities often introduced in an introductory geometry or trigonometry course.







                share|cite|improve this answer











                $endgroup$




                Here is a different approach that uses the integral definition of the arcsine function. We will deduce the limit of interest without appeal to geometry or differential calculus.



                Instead, we only rely on elementary analysis of continuous functions and their inverses along with simple properties of the Riemann integral. To that end, we now proceed.






                We define the sine function, $sin(x)$, as the inverse function of the function $f(x)$ given by




                $$bbox[5px,border:2px solid #C0A000]{f(x)=int_0^x frac{1}{sqrt{1-t^2}},dt }tag 1$$




                for $|x|< 1$.




                NOTE: It can be shown that the sine function defined as the inverse of $f(x)$ given in $(1)$ has all of the familiar properties that characterize the circular function $sin(x)$.




                It is straightforward to show that since $frac{1}{sqrt{1-t^2}}$ is positive and continuous for $tin (-1,1)$, $f(x)$ is continuous and strictly increasing for $xin (-1,1)$ with $displaystylelim_{xto 0}f(x)=f(0)=0$.



                Therefore, since $f$ is continuous and strictly increasing, its inverse function, $sin(x)$, exists and is also continuous and strictly increasing with $displaystyle lim_{xto 0}sin(x)=sin(0)=0$.





                From $(1)$, we have the bounds (SEE HERE)




                $$bbox[5px,border:2px solid #C0A000]{1 le frac{f(x)}xle frac{1}{sqrt{1-x^2}}} tag 2$$




                for $xin (-1,1)$, whence applying the squeeze theorem to $(2)$ yields



                $$lim_{xto 0}frac{f(x)}{x}=1 tag 3$$





                Finally, let $y=f(x)$ so that $x=sin(y)$. As $xto 0$, $yto 0$ and we can write $(3)$ as



                $$lim_{yto 0}frac{y}{sin(y)}=1$$



                from which we have




                $$bbox[5px,border:2px solid #C0A000]{lim_{yto 0}frac{sin(y)}{y}=1}$$




                as was to be shown!






                NOTE:



                We can deduce the following set of useful inequalities from $(2)$. We let $x=sin(theta)$ and restrict $x$ so that $xin [0,1)$. In addition, we define new functions, $cos(theta)=sqrt{1-sin^2(theta)}$ and $tan(theta)=sin(theta)/cos(theta)$.



                Then, we have from $(2)$



                $$bbox[5px,border:2px solid #C0A000]{ycos(y)le sin(y)le yle tan(y)} $$



                which are the familiar inequalities often introduced in an introductory geometry or trigonometry course.








                share|cite|improve this answer














                share|cite|improve this answer



                share|cite|improve this answer








                edited Dec 6 '16 at 18:53

























                answered Dec 6 '16 at 18:47









                Mark ViolaMark Viola

                132k1277174




                132k1277174








                • 1




                  $begingroup$
                  This is similar to my answer. You're formalizing the arc-length definition of $sin$ in the following three steps. 1) Defining $arcsin$ geometrically 2) Expressing that informal definition using integration 3) Defining $sin$ as the inverse of $arcsin$. Also, where you used the sqeeuze theorem, you could equally have used the Fundamental Theorem of Calculus and the Inverse Function Theorem. Then you'd get my answer
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 10:16








                • 1




                  $begingroup$
                  Another difference is you used the arc-length definition and I used the area definition
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 10:19






                • 1




                  $begingroup$
                  @ogogmad I don't use any geometrical argument. A function $f$ is defined without reference to the arc length interpretation. Both $f$ and its inverse has properties. It can be shown that the inverse function is the well-known sine function. And while one can apply the FOC instead of the squeeze theorem, there is no advantage in doing so. So, why on earth are you leaving these comments?
                  $endgroup$
                  – Mark Viola
                  Feb 5 '18 at 14:50






                • 1




                  $begingroup$
                  Because I'm pointing out that your abstract definitions make geometric sense, whether or not your realize it. I'm also pointing out that your squeeze argument is just a special case of FOC
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 16:43






                • 1




                  $begingroup$
                  In fact, all of the arguments above are morally the same, AFAICT. The squeeze arguments are just unpacking the proof of FOC on a special case. The FOC itself is proved using a very similar squeeze argument to the one people are using. The areas of the sectors are simply the integrals that are being differentiated. In tkr's argument, he's instead differentiating an arclength integral, which gives the same result anyway. There's the remaining feature that our two answers uses $arcsin$ while everyone else's uses $sin$; they may be implicitly unpacking the proof of the inverse function theorem.
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 17:01














                • 1




                  $begingroup$
                  This is similar to my answer. You're formalizing the arc-length definition of $sin$ in the following three steps. 1) Defining $arcsin$ geometrically 2) Expressing that informal definition using integration 3) Defining $sin$ as the inverse of $arcsin$. Also, where you used the sqeeuze theorem, you could equally have used the Fundamental Theorem of Calculus and the Inverse Function Theorem. Then you'd get my answer
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 10:16








                • 1




                  $begingroup$
                  Another difference is you used the arc-length definition and I used the area definition
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 10:19






                • 1




                  $begingroup$
                  @ogogmad I don't use any geometrical argument. A function $f$ is defined without reference to the arc length interpretation. Both $f$ and its inverse has properties. It can be shown that the inverse function is the well-known sine function. And while one can apply the FOC instead of the squeeze theorem, there is no advantage in doing so. So, why on earth are you leaving these comments?
                  $endgroup$
                  – Mark Viola
                  Feb 5 '18 at 14:50






                • 1




                  $begingroup$
                  Because I'm pointing out that your abstract definitions make geometric sense, whether or not your realize it. I'm also pointing out that your squeeze argument is just a special case of FOC
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 16:43






                • 1




                  $begingroup$
                  In fact, all of the arguments above are morally the same, AFAICT. The squeeze arguments are just unpacking the proof of FOC on a special case. The FOC itself is proved using a very similar squeeze argument to the one people are using. The areas of the sectors are simply the integrals that are being differentiated. In tkr's argument, he's instead differentiating an arclength integral, which gives the same result anyway. There's the remaining feature that our two answers uses $arcsin$ while everyone else's uses $sin$; they may be implicitly unpacking the proof of the inverse function theorem.
                  $endgroup$
                  – man on laptop
                  Feb 5 '18 at 17:01








                1




                1




                $begingroup$
                This is similar to my answer. You're formalizing the arc-length definition of $sin$ in the following three steps. 1) Defining $arcsin$ geometrically 2) Expressing that informal definition using integration 3) Defining $sin$ as the inverse of $arcsin$. Also, where you used the sqeeuze theorem, you could equally have used the Fundamental Theorem of Calculus and the Inverse Function Theorem. Then you'd get my answer
                $endgroup$
                – man on laptop
                Feb 5 '18 at 10:16






                $begingroup$
                This is similar to my answer. You're formalizing the arc-length definition of $sin$ in the following three steps. 1) Defining $arcsin$ geometrically 2) Expressing that informal definition using integration 3) Defining $sin$ as the inverse of $arcsin$. Also, where you used the sqeeuze theorem, you could equally have used the Fundamental Theorem of Calculus and the Inverse Function Theorem. Then you'd get my answer
                $endgroup$
                – man on laptop
                Feb 5 '18 at 10:16






                1




                1




                $begingroup$
                Another difference is you used the arc-length definition and I used the area definition
                $endgroup$
                – man on laptop
                Feb 5 '18 at 10:19




                $begingroup$
                Another difference is you used the arc-length definition and I used the area definition
                $endgroup$
                – man on laptop
                Feb 5 '18 at 10:19




                1




                1




                $begingroup$
                @ogogmad I don't use any geometrical argument. A function $f$ is defined without reference to the arc length interpretation. Both $f$ and its inverse has properties. It can be shown that the inverse function is the well-known sine function. And while one can apply the FOC instead of the squeeze theorem, there is no advantage in doing so. So, why on earth are you leaving these comments?
                $endgroup$
                – Mark Viola
                Feb 5 '18 at 14:50




                $begingroup$
                @ogogmad I don't use any geometrical argument. A function $f$ is defined without reference to the arc length interpretation. Both $f$ and its inverse has properties. It can be shown that the inverse function is the well-known sine function. And while one can apply the FOC instead of the squeeze theorem, there is no advantage in doing so. So, why on earth are you leaving these comments?
                $endgroup$
                – Mark Viola
                Feb 5 '18 at 14:50




                1




                1




                $begingroup$
                Because I'm pointing out that your abstract definitions make geometric sense, whether or not your realize it. I'm also pointing out that your squeeze argument is just a special case of FOC
                $endgroup$
                – man on laptop
                Feb 5 '18 at 16:43




                $begingroup$
                Because I'm pointing out that your abstract definitions make geometric sense, whether or not your realize it. I'm also pointing out that your squeeze argument is just a special case of FOC
                $endgroup$
                – man on laptop
                Feb 5 '18 at 16:43




                1




                1




                $begingroup$
                In fact, all of the arguments above are morally the same, AFAICT. The squeeze arguments are just unpacking the proof of FOC on a special case. The FOC itself is proved using a very similar squeeze argument to the one people are using. The areas of the sectors are simply the integrals that are being differentiated. In tkr's argument, he's instead differentiating an arclength integral, which gives the same result anyway. There's the remaining feature that our two answers uses $arcsin$ while everyone else's uses $sin$; they may be implicitly unpacking the proof of the inverse function theorem.
                $endgroup$
                – man on laptop
                Feb 5 '18 at 17:01




                $begingroup$
                In fact, all of the arguments above are morally the same, AFAICT. The squeeze arguments are just unpacking the proof of FOC on a special case. The FOC itself is proved using a very similar squeeze argument to the one people are using. The areas of the sectors are simply the integrals that are being differentiated. In tkr's argument, he's instead differentiating an arclength integral, which gives the same result anyway. There's the remaining feature that our two answers uses $arcsin$ while everyone else's uses $sin$; they may be implicitly unpacking the proof of the inverse function theorem.
                $endgroup$
                – man on laptop
                Feb 5 '18 at 17:01











                15












                $begingroup$

                Because $sin x$ has zeroes at $x=n pi$ for arbitrary integer $n$ including $x=0$, you can use Vieta's Theorem $sin x = A(cdots(x+2 pi)(x+pi)x(x-pi)(x-2 pi)cdots)$ with a constant $A$. Because $sin(frac{pi}{2})=1$ this constant can be determined by the equation: $$1=A(cdots(frac{pi}{2}+2 pi)(frac{pi}{2}+pi)frac{pi}{2}(frac{pi}{2}-pi)(frac{pi}{2}-2 pi)cdots).$$



                Now, in the Expression $f(x):= frac{sin(x)}{x}$ the $x$ cancels such that $$f(x)=A(cdots(x+2 pi)(x+pi)(x-pi)(x-2 pi)cdots),$$ hence: $$lim_{x rightarrow 0} f(x)=A(cdots(2 pi) cdot picdot(- pi)cdot(-2 pi)cdots) = A prod_{k=1}^infty (-k^2 pi^2).$$



                $frac{1}{A} = frac{pi}{2} prod_{k=1}^infty ((frac{pi}{2})^2-k^2 pi^2)$. The proof is completed when the Wallis product is used.






                share|cite|improve this answer











                $endgroup$









                • 5




                  $begingroup$
                  None of the infinite products in your answer converge.
                  $endgroup$
                  – Antonio Vargas
                  Oct 15 '15 at 1:30






                • 2




                  $begingroup$
                  Ontop of Antonio Vargas comment, Vieta's theorem only works for finite number of zeros. You need the Weierstrass-Factorization Theorem or Hadamard-Factorization Theorem. Btw you also need to verify that $sin(x)$ does not have any complex zeros.
                  $endgroup$
                  – MrYouMath
                  Jun 10 '16 at 12:52
















                15












                $begingroup$

                Because $sin x$ has zeroes at $x=n pi$ for arbitrary integer $n$ including $x=0$, you can use Vieta's Theorem $sin x = A(cdots(x+2 pi)(x+pi)x(x-pi)(x-2 pi)cdots)$ with a constant $A$. Because $sin(frac{pi}{2})=1$ this constant can be determined by the equation: $$1=A(cdots(frac{pi}{2}+2 pi)(frac{pi}{2}+pi)frac{pi}{2}(frac{pi}{2}-pi)(frac{pi}{2}-2 pi)cdots).$$



                Now, in the Expression $f(x):= frac{sin(x)}{x}$ the $x$ cancels such that $$f(x)=A(cdots(x+2 pi)(x+pi)(x-pi)(x-2 pi)cdots),$$ hence: $$lim_{x rightarrow 0} f(x)=A(cdots(2 pi) cdot picdot(- pi)cdot(-2 pi)cdots) = A prod_{k=1}^infty (-k^2 pi^2).$$



                $frac{1}{A} = frac{pi}{2} prod_{k=1}^infty ((frac{pi}{2})^2-k^2 pi^2)$. The proof is completed when the Wallis product is used.






                share|cite|improve this answer











                $endgroup$









                • 5




                  $begingroup$
                  None of the infinite products in your answer converge.
                  $endgroup$
                  – Antonio Vargas
                  Oct 15 '15 at 1:30






                • 2




                  $begingroup$
                  Ontop of Antonio Vargas comment, Vieta's theorem only works for finite number of zeros. You need the Weierstrass-Factorization Theorem or Hadamard-Factorization Theorem. Btw you also need to verify that $sin(x)$ does not have any complex zeros.
                  $endgroup$
                  – MrYouMath
                  Jun 10 '16 at 12:52














                15












                15








                15





                $begingroup$

                Because $sin x$ has zeroes at $x=n pi$ for arbitrary integer $n$ including $x=0$, you can use Vieta's Theorem $sin x = A(cdots(x+2 pi)(x+pi)x(x-pi)(x-2 pi)cdots)$ with a constant $A$. Because $sin(frac{pi}{2})=1$ this constant can be determined by the equation: $$1=A(cdots(frac{pi}{2}+2 pi)(frac{pi}{2}+pi)frac{pi}{2}(frac{pi}{2}-pi)(frac{pi}{2}-2 pi)cdots).$$



                Now, in the Expression $f(x):= frac{sin(x)}{x}$ the $x$ cancels such that $$f(x)=A(cdots(x+2 pi)(x+pi)(x-pi)(x-2 pi)cdots),$$ hence: $$lim_{x rightarrow 0} f(x)=A(cdots(2 pi) cdot picdot(- pi)cdot(-2 pi)cdots) = A prod_{k=1}^infty (-k^2 pi^2).$$



                $frac{1}{A} = frac{pi}{2} prod_{k=1}^infty ((frac{pi}{2})^2-k^2 pi^2)$. The proof is completed when the Wallis product is used.






                share|cite|improve this answer











                $endgroup$



                Because $sin x$ has zeroes at $x=n pi$ for arbitrary integer $n$ including $x=0$, you can use Vieta's Theorem $sin x = A(cdots(x+2 pi)(x+pi)x(x-pi)(x-2 pi)cdots)$ with a constant $A$. Because $sin(frac{pi}{2})=1$ this constant can be determined by the equation: $$1=A(cdots(frac{pi}{2}+2 pi)(frac{pi}{2}+pi)frac{pi}{2}(frac{pi}{2}-pi)(frac{pi}{2}-2 pi)cdots).$$



                Now, in the Expression $f(x):= frac{sin(x)}{x}$ the $x$ cancels such that $$f(x)=A(cdots(x+2 pi)(x+pi)(x-pi)(x-2 pi)cdots),$$ hence: $$lim_{x rightarrow 0} f(x)=A(cdots(2 pi) cdot picdot(- pi)cdot(-2 pi)cdots) = A prod_{k=1}^infty (-k^2 pi^2).$$



                $frac{1}{A} = frac{pi}{2} prod_{k=1}^infty ((frac{pi}{2})^2-k^2 pi^2)$. The proof is completed when the Wallis product is used.







                share|cite|improve this answer














                share|cite|improve this answer



                share|cite|improve this answer








                edited Jun 18 '16 at 1:38









                Michael Hardy

                1




                1










                answered Mar 4 '15 at 20:15









                kryomaximkryomaxim

                2,3571825




                2,3571825








                • 5




                  $begingroup$
                  None of the infinite products in your answer converge.
                  $endgroup$
                  – Antonio Vargas
                  Oct 15 '15 at 1:30






                • 2




                  $begingroup$
                  Ontop of Antonio Vargas comment, Vieta's theorem only works for finite number of zeros. You need the Weierstrass-Factorization Theorem or Hadamard-Factorization Theorem. Btw you also need to verify that $sin(x)$ does not have any complex zeros.
                  $endgroup$
                  – MrYouMath
                  Jun 10 '16 at 12:52














                • 5




                  $begingroup$
                  None of the infinite products in your answer converge.
                  $endgroup$
                  – Antonio Vargas
                  Oct 15 '15 at 1:30






                • 2




                  $begingroup$
                  Ontop of Antonio Vargas comment, Vieta's theorem only works for finite number of zeros. You need the Weierstrass-Factorization Theorem or Hadamard-Factorization Theorem. Btw you also need to verify that $sin(x)$ does not have any complex zeros.
                  $endgroup$
                  – MrYouMath
                  Jun 10 '16 at 12:52








                5




                5




                $begingroup$
                None of the infinite products in your answer converge.
                $endgroup$
                – Antonio Vargas
                Oct 15 '15 at 1:30




                $begingroup$
                None of the infinite products in your answer converge.
                $endgroup$
                – Antonio Vargas
                Oct 15 '15 at 1:30




                2




                2




                $begingroup$
                Ontop of Antonio Vargas comment, Vieta's theorem only works for finite number of zeros. You need the Weierstrass-Factorization Theorem or Hadamard-Factorization Theorem. Btw you also need to verify that $sin(x)$ does not have any complex zeros.
                $endgroup$
                – MrYouMath
                Jun 10 '16 at 12:52




                $begingroup$
                Ontop of Antonio Vargas comment, Vieta's theorem only works for finite number of zeros. You need the Weierstrass-Factorization Theorem or Hadamard-Factorization Theorem. Btw you also need to verify that $sin(x)$ does not have any complex zeros.
                $endgroup$
                – MrYouMath
                Jun 10 '16 at 12:52











                14












                $begingroup$

                Usual proofs can be circular, but there is a simple way for proving such inequality.



                Let $theta$ be an acute angle and let $O,A,B,C,D,C'$ as in the following diagram:
                enter image description here



                We may show that:



                $$ CD stackrel{(1)}{ geq };stackrel{largefrown}{CB}; stackrel{(2)}{geq } CB,stackrel{(3)}{geq} AB $$



                $(1)$: The quadrilateral $OCDC'$ and the circle sector delimited by $O,C,C'$ are two convex sets. Since the circle sector is a subset of the quadrilateral, the perimeter of the circle sector is less than the perimeter of the quadrilateral.



                $(2)$: the $CB$ segment is the shortest path between $B$ and $C$.



                $(3)$ $CAB$ is a right triangle, hence $CBgeq AB$ by the Pythagorean theorem.



                In terms of $theta$ we get:
                $$ tantheta geq theta geq 2sinfrac{theta}{2} geq sintheta $$
                for any $thetainleft[0,frac{pi}{2}right)$. Since the involved functions are odd functions the reverse inequality holds over $left(-frac{pi}{2},0right]$, and $lim_{thetato 0}frac{sintheta}{theta}=1$ follows by squeezing.





                A slightly different approach might be the following one: let us assume $thetainleft(0,frac{pi}{2}right)$.
                By $(2)$ and $(3)$ we have
                $$ theta geq 2sinfrac{theta}{2}geq sintheta $$
                hence the sequence ${a_n}_{ngeq 0}$ defined by $a_n = 2^n sinfrac{theta}{2^n}$ is increasing and bounded by $theta$. Any increasing and bounded sequence is convergent, and we actually have $lim_{nto +infty}a_n=theta$ since $stackrel{largefrown}{BC}$ is a rectifiable curve and for every $ngeq 1$ the $a_n$ term is the length of a polygonal approximation of $stackrel{largefrown}{BC}$ through $2^{n-1}$ equal segments. In particular



                $$ forall thetainleft(0,frac{pi}{2}right), qquad lim_{nto +infty}frac{sinleft(frac{theta}{2^n}right)}{frac{theta}{2^n}} = 1 $$
                and this grants that if the limit $lim_{xto 0}frac{sin x}{x}$ exists, it is $1$. By $sin xleq x$ we get $limsup_{xto 0}frac{sin x}{x}leq 1$, hence it is enough to show that $liminf_{xto 0}frac{sin x}{x}geq 1$. We already know that for any $x$ close enough to the origin the sequence $frac{sin x}{x},frac{sin(x/2)}{x/2},frac{sin(x/4)}{x/4},ldots$ is convergent to $1$, hence we are done.



                Long story short: $lim_{xto 0}frac{sin x}{x}=1$ follows from the fact that a circle is a rectifiable curve, and a circle is a rectifiable curve because it is the boundary of a convex, bounded subset of $mathbb{R}^2$. The convexity of the disk follows from the triangle inequality: a disk is a closed ball for the euclidean distance.





                $(1)$ relies on this powerful Lemma:




                Lemma. If $A,B$ are convex bounded sets in $mathbb{R}^2$ and $Asubsetneq B$, the perimeter of $A$ is less than the perimeter of $B$.




                Proof: by boundedness and convexity, $partial A$ and $partial B$ are rectifiable, with lengths $L(A)=mu(partial A),,L(B)=mu(partial B)$. Always by convexity, there is some chord in $B$ that does not meet the interior of $A$ (a tangent to $partial A$ at a smooth point does the job, for instance). Assume that such chord has endpoints $B_1, B_2 in partial B$ and perform a cut along $B_1 B_2$: both the area and the perimeter of $B$ decrease, but $B$ remains a bounded convex set enclosing $A$. Since $A$ can be approximated through a sequence of consecutive cuts, $L(A)<L(B)$ follows.






                share|cite|improve this answer











                $endgroup$









                • 1




                  $begingroup$
                  Late to the party, but I think this is the non-circular (no puns intended) way of thinking that has been missing so far. Nice job!
                  $endgroup$
                  – String
                  Jun 13 '17 at 9:05






                • 1




                  $begingroup$
                  Up voting! I think Archimedes used a weaker form of your Lemma to get the circumference of a circle.
                  $endgroup$
                  – CopyPasteIt
                  Jun 21 '17 at 23:49










                • $begingroup$
                  Great answer (+1)
                  $endgroup$
                  – MathLover
                  Jan 25 '18 at 23:07
















                14












                $begingroup$

                Usual proofs can be circular, but there is a simple way for proving such inequality.



                Let $theta$ be an acute angle and let $O,A,B,C,D,C'$ as in the following diagram:
                enter image description here



                We may show that:



                $$ CD stackrel{(1)}{ geq };stackrel{largefrown}{CB}; stackrel{(2)}{geq } CB,stackrel{(3)}{geq} AB $$



                $(1)$: The quadrilateral $OCDC'$ and the circle sector delimited by $O,C,C'$ are two convex sets. Since the circle sector is a subset of the quadrilateral, the perimeter of the circle sector is less than the perimeter of the quadrilateral.



                $(2)$: the $CB$ segment is the shortest path between $B$ and $C$.



                $(3)$ $CAB$ is a right triangle, hence $CBgeq AB$ by the Pythagorean theorem.



                In terms of $theta$ we get:
                $$ tantheta geq theta geq 2sinfrac{theta}{2} geq sintheta $$
                for any $thetainleft[0,frac{pi}{2}right)$. Since the involved functions are odd functions the reverse inequality holds over $left(-frac{pi}{2},0right]$, and $lim_{thetato 0}frac{sintheta}{theta}=1$ follows by squeezing.





                A slightly different approach might be the following one: let us assume $thetainleft(0,frac{pi}{2}right)$.
                By $(2)$ and $(3)$ we have
                $$ theta geq 2sinfrac{theta}{2}geq sintheta $$
                hence the sequence ${a_n}_{ngeq 0}$ defined by $a_n = 2^n sinfrac{theta}{2^n}$ is increasing and bounded by $theta$. Any increasing and bounded sequence is convergent, and we actually have $lim_{nto +infty}a_n=theta$ since $stackrel{largefrown}{BC}$ is a rectifiable curve and for every $ngeq 1$ the $a_n$ term is the length of a polygonal approximation of $stackrel{largefrown}{BC}$ through $2^{n-1}$ equal segments. In particular



                $$ forall thetainleft(0,frac{pi}{2}right), qquad lim_{nto +infty}frac{sinleft(frac{theta}{2^n}right)}{frac{theta}{2^n}} = 1 $$
                and this grants that if the limit $lim_{xto 0}frac{sin x}{x}$ exists, it is $1$. By $sin xleq x$ we get $limsup_{xto 0}frac{sin x}{x}leq 1$, hence it is enough to show that $liminf_{xto 0}frac{sin x}{x}geq 1$. We already know that for any $x$ close enough to the origin the sequence $frac{sin x}{x},frac{sin(x/2)}{x/2},frac{sin(x/4)}{x/4},ldots$ is convergent to $1$, hence we are done.



                Long story short: $lim_{xto 0}frac{sin x}{x}=1$ follows from the fact that a circle is a rectifiable curve, and a circle is a rectifiable curve because it is the boundary of a convex, bounded subset of $mathbb{R}^2$. The convexity of the disk follows from the triangle inequality: a disk is a closed ball for the euclidean distance.





                $(1)$ relies on this powerful Lemma:




                Lemma. If $A,B$ are convex bounded sets in $mathbb{R}^2$ and $Asubsetneq B$, the perimeter of $A$ is less than the perimeter of $B$.




                Proof: by boundedness and convexity, $partial A$ and $partial B$ are rectifiable, with lengths $L(A)=mu(partial A),,L(B)=mu(partial B)$. Always by convexity, there is some chord in $B$ that does not meet the interior of $A$ (a tangent to $partial A$ at a smooth point does the job, for instance). Assume that such chord has endpoints $B_1, B_2 in partial B$ and perform a cut along $B_1 B_2$: both the area and the perimeter of $B$ decrease, but $B$ remains a bounded convex set enclosing $A$. Since $A$ can be approximated through a sequence of consecutive cuts, $L(A)<L(B)$ follows.






                share|cite|improve this answer











                $endgroup$









                • 1




                  $begingroup$
                  Late to the party, but I think this is the non-circular (no puns intended) way of thinking that has been missing so far. Nice job!
                  $endgroup$
                  – String
                  Jun 13 '17 at 9:05






                • 1




                  $begingroup$
                  Up voting! I think Archimedes used a weaker form of your Lemma to get the circumference of a circle.
                  $endgroup$
                  – CopyPasteIt
                  Jun 21 '17 at 23:49










                • $begingroup$
                  Great answer (+1)
                  $endgroup$
                  – MathLover
                  Jan 25 '18 at 23:07














                14












                14








                14





                $begingroup$

                Usual proofs can be circular, but there is a simple way for proving such inequality.



                Let $theta$ be an acute angle and let $O,A,B,C,D,C'$ as in the following diagram:
                enter image description here



                We may show that:



                $$ CD stackrel{(1)}{ geq };stackrel{largefrown}{CB}; stackrel{(2)}{geq } CB,stackrel{(3)}{geq} AB $$



                $(1)$: The quadrilateral $OCDC'$ and the circle sector delimited by $O,C,C'$ are two convex sets. Since the circle sector is a subset of the quadrilateral, the perimeter of the circle sector is less than the perimeter of the quadrilateral.



                $(2)$: the $CB$ segment is the shortest path between $B$ and $C$.



                $(3)$ $CAB$ is a right triangle, hence $CBgeq AB$ by the Pythagorean theorem.



                In terms of $theta$ we get:
                $$ tantheta geq theta geq 2sinfrac{theta}{2} geq sintheta $$
                for any $thetainleft[0,frac{pi}{2}right)$. Since the involved functions are odd functions the reverse inequality holds over $left(-frac{pi}{2},0right]$, and $lim_{thetato 0}frac{sintheta}{theta}=1$ follows by squeezing.





                A slightly different approach might be the following one: let us assume $thetainleft(0,frac{pi}{2}right)$.
                By $(2)$ and $(3)$ we have
                $$ theta geq 2sinfrac{theta}{2}geq sintheta $$
                hence the sequence ${a_n}_{ngeq 0}$ defined by $a_n = 2^n sinfrac{theta}{2^n}$ is increasing and bounded by $theta$. Any increasing and bounded sequence is convergent, and we actually have $lim_{nto +infty}a_n=theta$ since $stackrel{largefrown}{BC}$ is a rectifiable curve and for every $ngeq 1$ the $a_n$ term is the length of a polygonal approximation of $stackrel{largefrown}{BC}$ through $2^{n-1}$ equal segments. In particular



                $$ forall thetainleft(0,frac{pi}{2}right), qquad lim_{nto +infty}frac{sinleft(frac{theta}{2^n}right)}{frac{theta}{2^n}} = 1 $$
                and this grants that if the limit $lim_{xto 0}frac{sin x}{x}$ exists, it is $1$. By $sin xleq x$ we get $limsup_{xto 0}frac{sin x}{x}leq 1$, hence it is enough to show that $liminf_{xto 0}frac{sin x}{x}geq 1$. We already know that for any $x$ close enough to the origin the sequence $frac{sin x}{x},frac{sin(x/2)}{x/2},frac{sin(x/4)}{x/4},ldots$ is convergent to $1$, hence we are done.



                Long story short: $lim_{xto 0}frac{sin x}{x}=1$ follows from the fact that a circle is a rectifiable curve, and a circle is a rectifiable curve because it is the boundary of a convex, bounded subset of $mathbb{R}^2$. The convexity of the disk follows from the triangle inequality: a disk is a closed ball for the euclidean distance.





                $(1)$ relies on this powerful Lemma:




                Lemma. If $A,B$ are convex bounded sets in $mathbb{R}^2$ and $Asubsetneq B$, the perimeter of $A$ is less than the perimeter of $B$.




                Proof: by boundedness and convexity, $partial A$ and $partial B$ are rectifiable, with lengths $L(A)=mu(partial A),,L(B)=mu(partial B)$. Always by convexity, there is some chord in $B$ that does not meet the interior of $A$ (a tangent to $partial A$ at a smooth point does the job, for instance). Assume that such chord has endpoints $B_1, B_2 in partial B$ and perform a cut along $B_1 B_2$: both the area and the perimeter of $B$ decrease, but $B$ remains a bounded convex set enclosing $A$. Since $A$ can be approximated through a sequence of consecutive cuts, $L(A)<L(B)$ follows.






                share|cite|improve this answer











                $endgroup$



                Usual proofs can be circular, but there is a simple way for proving such inequality.



                Let $theta$ be an acute angle and let $O,A,B,C,D,C'$ as in the following diagram:
                enter image description here



                We may show that:



                $$ CD stackrel{(1)}{ geq };stackrel{largefrown}{CB}; stackrel{(2)}{geq } CB,stackrel{(3)}{geq} AB $$



                $(1)$: The quadrilateral $OCDC'$ and the circle sector delimited by $O,C,C'$ are two convex sets. Since the circle sector is a subset of the quadrilateral, the perimeter of the circle sector is less than the perimeter of the quadrilateral.



                $(2)$: the $CB$ segment is the shortest path between $B$ and $C$.



                $(3)$ $CAB$ is a right triangle, hence $CBgeq AB$ by the Pythagorean theorem.



                In terms of $theta$ we get:
                $$ tantheta geq theta geq 2sinfrac{theta}{2} geq sintheta $$
                for any $thetainleft[0,frac{pi}{2}right)$. Since the involved functions are odd functions the reverse inequality holds over $left(-frac{pi}{2},0right]$, and $lim_{thetato 0}frac{sintheta}{theta}=1$ follows by squeezing.





                A slightly different approach might be the following one: let us assume $thetainleft(0,frac{pi}{2}right)$.
                By $(2)$ and $(3)$ we have
                $$ theta geq 2sinfrac{theta}{2}geq sintheta $$
                hence the sequence ${a_n}_{ngeq 0}$ defined by $a_n = 2^n sinfrac{theta}{2^n}$ is increasing and bounded by $theta$. Any increasing and bounded sequence is convergent, and we actually have $lim_{nto +infty}a_n=theta$ since $stackrel{largefrown}{BC}$ is a rectifiable curve and for every $ngeq 1$ the $a_n$ term is the length of a polygonal approximation of $stackrel{largefrown}{BC}$ through $2^{n-1}$ equal segments. In particular



                $$ forall thetainleft(0,frac{pi}{2}right), qquad lim_{nto +infty}frac{sinleft(frac{theta}{2^n}right)}{frac{theta}{2^n}} = 1 $$
                and this grants that if the limit $lim_{xto 0}frac{sin x}{x}$ exists, it is $1$. By $sin xleq x$ we get $limsup_{xto 0}frac{sin x}{x}leq 1$, hence it is enough to show that $liminf_{xto 0}frac{sin x}{x}geq 1$. We already know that for any $x$ close enough to the origin the sequence $frac{sin x}{x},frac{sin(x/2)}{x/2},frac{sin(x/4)}{x/4},ldots$ is convergent to $1$, hence we are done.



                Long story short: $lim_{xto 0}frac{sin x}{x}=1$ follows from the fact that a circle is a rectifiable curve, and a circle is a rectifiable curve because it is the boundary of a convex, bounded subset of $mathbb{R}^2$. The convexity of the disk follows from the triangle inequality: a disk is a closed ball for the euclidean distance.





                $(1)$ relies on this powerful Lemma:




                Lemma. If $A,B$ are convex bounded sets in $mathbb{R}^2$ and $Asubsetneq B$, the perimeter of $A$ is less than the perimeter of $B$.




                Proof: by boundedness and convexity, $partial A$ and $partial B$ are rectifiable, with lengths $L(A)=mu(partial A),,L(B)=mu(partial B)$. Always by convexity, there is some chord in $B$ that does not meet the interior of $A$ (a tangent to $partial A$ at a smooth point does the job, for instance). Assume that such chord has endpoints $B_1, B_2 in partial B$ and perform a cut along $B_1 B_2$: both the area and the perimeter of $B$ decrease, but $B$ remains a bounded convex set enclosing $A$. Since $A$ can be approximated through a sequence of consecutive cuts, $L(A)<L(B)$ follows.







                share|cite|improve this answer














                share|cite|improve this answer



                share|cite|improve this answer








                edited Jun 13 '17 at 6:35

























                answered Jun 13 '17 at 0:03









                Jack D'AurizioJack D'Aurizio

                290k33282664




                290k33282664








                • 1




                  $begingroup$
                  Late to the party, but I think this is the non-circular (no puns intended) way of thinking that has been missing so far. Nice job!
                  $endgroup$
                  – String
                  Jun 13 '17 at 9:05






                • 1




                  $begingroup$
                  Up voting! I think Archimedes used a weaker form of your Lemma to get the circumference of a circle.
                  $endgroup$
                  – CopyPasteIt
                  Jun 21 '17 at 23:49










                • $begingroup$
                  Great answer (+1)
                  $endgroup$
                  – MathLover
                  Jan 25 '18 at 23:07














                • 1




                  $begingroup$
                  Late to the party, but I think this is the non-circular (no puns intended) way of thinking that has been missing so far. Nice job!
                  $endgroup$
                  – String
                  Jun 13 '17 at 9:05






                • 1




                  $begingroup$
                  Up voting! I think Archimedes used a weaker form of your Lemma to get the circumference of a circle.
                  $endgroup$
                  – CopyPasteIt
                  Jun 21 '17 at 23:49










                • $begingroup$
                  Great answer (+1)
                  $endgroup$
                  – MathLover
                  Jan 25 '18 at 23:07








                1




                1




                $begingroup$
                Late to the party, but I think this is the non-circular (no puns intended) way of thinking that has been missing so far. Nice job!
                $endgroup$
                – String
                Jun 13 '17 at 9:05




                $begingroup$
                Late to the party, but I think this is the non-circular (no puns intended) way of thinking that has been missing so far. Nice job!
                $endgroup$
                – String
                Jun 13 '17 at 9:05




                1




                1




                $begingroup$
                Up voting! I think Archimedes used a weaker form of your Lemma to get the circumference of a circle.
                $endgroup$
                – CopyPasteIt
                Jun 21 '17 at 23:49




                $begingroup$
                Up voting! I think Archimedes used a weaker form of your Lemma to get the circumference of a circle.
                $endgroup$
                – CopyPasteIt
                Jun 21 '17 at 23:49












                $begingroup$
                Great answer (+1)
                $endgroup$
                – MathLover
                Jan 25 '18 at 23:07




                $begingroup$
                Great answer (+1)
                $endgroup$
                – MathLover
                Jan 25 '18 at 23:07











                12












                $begingroup$

                Let $sin(x)$ is defined as solution of $frac{d^2}{dx^2}textrm{f}(x)=-textrm{f}(x)$ with $mathrm f(0)=0,,frac{d}{dx}mathrm f(0)=C$ initial conditions, so exact solution is $mathrm f(x)=Ccdotsin(x)$.
                Define second derivative as
                $$
                begin{align*}
                frac{d^2}{dx^2}textrm{f}(x)=lim_{Delta xto 0}{frac{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}-frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}{Delta x}}&=\=lim_{Delta xto 0}{frac{mathrm f(x)-2cdot mathrm f(x-Delta x)+mathrm f(x-2cdotDelta x)}{Delta x^2}}
                end{align*}
                $$
                we can easy check this limit for any (?) functions. Similarly, we can define the first derivative for right, middle and left points:
                $$
                frac{d}{dx}textrm{f}(x)left{
                begin{aligned}
                &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}}
                \
                &=lim_{Delta xto 0}{frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}\
                &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-2cdotDelta x)}{2cdotDelta x}}
                end{aligned}
                right.
                $$
                Let's use the finite elements method assuming $Td=Delta x,,y_n=mathrm f(x),,y_{n-1}=mathrm f(x-Delta x),,y_{n-2}=mathrm f(x-2cdot Delta x)$
                Override differential equation as
                $$
                frac{y_n-2cdot y_{n-1}+y_{n-2}}{Td^2}=-y_n
                $$
                Now solve this implicit equation for $y_n$ to obtain explicit recurrence relation:
                $$
                y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                $$
                Using arbitrarily small but non-zero quantity Td we can plot exponentially decaying sampled sine function (because the poles are inside the unit circle of the transfer function corresponding to the given recurrence relation).
                Similarly we write three systems for the initial conditions:



                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_n-y_{n-1}}{Td}
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_{n-1}-y_{n-2}}{Td}
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_n-y_{n-2}}{2cdot Td}
                end{aligned}right.
                $$
                Solve this sequence of equations for $y_{n-1}$ and $y_{n-2}$:
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1-Td^2right)
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+Td^2right)\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1+Td^2right)
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+frac{Td^2}{2}right)\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}
                end{aligned}right.
                $$
                At zero point $y_n=mathrm f(0)=0$ and we can see linear dependence:
                $$
                begin{aligned}
                &y_{n-1} = -Ccdot Td\
                &y_{n-2}=-2cdot Ccdot Td
                end{aligned}
                $$
                for all three solutions. Replace back:
                $$
                begin{array}{l}
                mathrm f(0)&=0\
                mathrm f(0-Delta x) &= -Ccdot Delta x\
                mathrm f(0-2cdot Delta x) &= -2cdot Ccdot Delta x
                end{array}
                $$
                So all three $frac{d}{dx}mathrm f(0)$ limits is equal to $C$ at $x=0$ and in accordance with $mathrm f(x)=Ccdotsin(x)$ by definition we can write
                $$
                lim_{Delta xto 0}{frac{mathrm f(0)-mathrm f(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{0-(-C cdot Delta x)}{Delta x}}=C
                $$
                Thus
                $$
                lim_{Delta xto 0}{frac{sin(0)-Ccdotsin(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{Ccdotsin(Delta x)}{Delta x}}=Ccdotlim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=C
                $$
                and $lim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=1$






                share|cite|improve this answer









                $endgroup$













                • $begingroup$
                  While this is an exceptionally neat answer, it is not really appropriate for the level (high school) on which I asked the question, given that computing this limit is an introductory exercise into differentials. Plus we defined $sin$ geometrically (i.e. though the relationship between angles and sides in a right triangle), so your prerequisites must be established first.
                  $endgroup$
                  – FUZxxl
                  Aug 19 '16 at 20:40










                • $begingroup$
                  @FUZxxl do you realize that users with different levels and mathematical background may check this question? The answer given above might be helpful for them.
                  $endgroup$
                  – Xam
                  Sep 23 '17 at 3:38
















                12












                $begingroup$

                Let $sin(x)$ is defined as solution of $frac{d^2}{dx^2}textrm{f}(x)=-textrm{f}(x)$ with $mathrm f(0)=0,,frac{d}{dx}mathrm f(0)=C$ initial conditions, so exact solution is $mathrm f(x)=Ccdotsin(x)$.
                Define second derivative as
                $$
                begin{align*}
                frac{d^2}{dx^2}textrm{f}(x)=lim_{Delta xto 0}{frac{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}-frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}{Delta x}}&=\=lim_{Delta xto 0}{frac{mathrm f(x)-2cdot mathrm f(x-Delta x)+mathrm f(x-2cdotDelta x)}{Delta x^2}}
                end{align*}
                $$
                we can easy check this limit for any (?) functions. Similarly, we can define the first derivative for right, middle and left points:
                $$
                frac{d}{dx}textrm{f}(x)left{
                begin{aligned}
                &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}}
                \
                &=lim_{Delta xto 0}{frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}\
                &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-2cdotDelta x)}{2cdotDelta x}}
                end{aligned}
                right.
                $$
                Let's use the finite elements method assuming $Td=Delta x,,y_n=mathrm f(x),,y_{n-1}=mathrm f(x-Delta x),,y_{n-2}=mathrm f(x-2cdot Delta x)$
                Override differential equation as
                $$
                frac{y_n-2cdot y_{n-1}+y_{n-2}}{Td^2}=-y_n
                $$
                Now solve this implicit equation for $y_n$ to obtain explicit recurrence relation:
                $$
                y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                $$
                Using arbitrarily small but non-zero quantity Td we can plot exponentially decaying sampled sine function (because the poles are inside the unit circle of the transfer function corresponding to the given recurrence relation).
                Similarly we write three systems for the initial conditions:



                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_n-y_{n-1}}{Td}
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_{n-1}-y_{n-2}}{Td}
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_n-y_{n-2}}{2cdot Td}
                end{aligned}right.
                $$
                Solve this sequence of equations for $y_{n-1}$ and $y_{n-2}$:
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1-Td^2right)
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+Td^2right)\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1+Td^2right)
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+frac{Td^2}{2}right)\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}
                end{aligned}right.
                $$
                At zero point $y_n=mathrm f(0)=0$ and we can see linear dependence:
                $$
                begin{aligned}
                &y_{n-1} = -Ccdot Td\
                &y_{n-2}=-2cdot Ccdot Td
                end{aligned}
                $$
                for all three solutions. Replace back:
                $$
                begin{array}{l}
                mathrm f(0)&=0\
                mathrm f(0-Delta x) &= -Ccdot Delta x\
                mathrm f(0-2cdot Delta x) &= -2cdot Ccdot Delta x
                end{array}
                $$
                So all three $frac{d}{dx}mathrm f(0)$ limits is equal to $C$ at $x=0$ and in accordance with $mathrm f(x)=Ccdotsin(x)$ by definition we can write
                $$
                lim_{Delta xto 0}{frac{mathrm f(0)-mathrm f(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{0-(-C cdot Delta x)}{Delta x}}=C
                $$
                Thus
                $$
                lim_{Delta xto 0}{frac{sin(0)-Ccdotsin(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{Ccdotsin(Delta x)}{Delta x}}=Ccdotlim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=C
                $$
                and $lim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=1$






                share|cite|improve this answer









                $endgroup$













                • $begingroup$
                  While this is an exceptionally neat answer, it is not really appropriate for the level (high school) on which I asked the question, given that computing this limit is an introductory exercise into differentials. Plus we defined $sin$ geometrically (i.e. though the relationship between angles and sides in a right triangle), so your prerequisites must be established first.
                  $endgroup$
                  – FUZxxl
                  Aug 19 '16 at 20:40










                • $begingroup$
                  @FUZxxl do you realize that users with different levels and mathematical background may check this question? The answer given above might be helpful for them.
                  $endgroup$
                  – Xam
                  Sep 23 '17 at 3:38














                12












                12








                12





                $begingroup$

                Let $sin(x)$ is defined as solution of $frac{d^2}{dx^2}textrm{f}(x)=-textrm{f}(x)$ with $mathrm f(0)=0,,frac{d}{dx}mathrm f(0)=C$ initial conditions, so exact solution is $mathrm f(x)=Ccdotsin(x)$.
                Define second derivative as
                $$
                begin{align*}
                frac{d^2}{dx^2}textrm{f}(x)=lim_{Delta xto 0}{frac{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}-frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}{Delta x}}&=\=lim_{Delta xto 0}{frac{mathrm f(x)-2cdot mathrm f(x-Delta x)+mathrm f(x-2cdotDelta x)}{Delta x^2}}
                end{align*}
                $$
                we can easy check this limit for any (?) functions. Similarly, we can define the first derivative for right, middle and left points:
                $$
                frac{d}{dx}textrm{f}(x)left{
                begin{aligned}
                &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}}
                \
                &=lim_{Delta xto 0}{frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}\
                &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-2cdotDelta x)}{2cdotDelta x}}
                end{aligned}
                right.
                $$
                Let's use the finite elements method assuming $Td=Delta x,,y_n=mathrm f(x),,y_{n-1}=mathrm f(x-Delta x),,y_{n-2}=mathrm f(x-2cdot Delta x)$
                Override differential equation as
                $$
                frac{y_n-2cdot y_{n-1}+y_{n-2}}{Td^2}=-y_n
                $$
                Now solve this implicit equation for $y_n$ to obtain explicit recurrence relation:
                $$
                y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                $$
                Using arbitrarily small but non-zero quantity Td we can plot exponentially decaying sampled sine function (because the poles are inside the unit circle of the transfer function corresponding to the given recurrence relation).
                Similarly we write three systems for the initial conditions:



                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_n-y_{n-1}}{Td}
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_{n-1}-y_{n-2}}{Td}
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_n-y_{n-2}}{2cdot Td}
                end{aligned}right.
                $$
                Solve this sequence of equations for $y_{n-1}$ and $y_{n-2}$:
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1-Td^2right)
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+Td^2right)\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1+Td^2right)
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+frac{Td^2}{2}right)\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}
                end{aligned}right.
                $$
                At zero point $y_n=mathrm f(0)=0$ and we can see linear dependence:
                $$
                begin{aligned}
                &y_{n-1} = -Ccdot Td\
                &y_{n-2}=-2cdot Ccdot Td
                end{aligned}
                $$
                for all three solutions. Replace back:
                $$
                begin{array}{l}
                mathrm f(0)&=0\
                mathrm f(0-Delta x) &= -Ccdot Delta x\
                mathrm f(0-2cdot Delta x) &= -2cdot Ccdot Delta x
                end{array}
                $$
                So all three $frac{d}{dx}mathrm f(0)$ limits is equal to $C$ at $x=0$ and in accordance with $mathrm f(x)=Ccdotsin(x)$ by definition we can write
                $$
                lim_{Delta xto 0}{frac{mathrm f(0)-mathrm f(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{0-(-C cdot Delta x)}{Delta x}}=C
                $$
                Thus
                $$
                lim_{Delta xto 0}{frac{sin(0)-Ccdotsin(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{Ccdotsin(Delta x)}{Delta x}}=Ccdotlim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=C
                $$
                and $lim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=1$






                share|cite|improve this answer









                $endgroup$



                Let $sin(x)$ is defined as solution of $frac{d^2}{dx^2}textrm{f}(x)=-textrm{f}(x)$ with $mathrm f(0)=0,,frac{d}{dx}mathrm f(0)=C$ initial conditions, so exact solution is $mathrm f(x)=Ccdotsin(x)$.
                Define second derivative as
                $$
                begin{align*}
                frac{d^2}{dx^2}textrm{f}(x)=lim_{Delta xto 0}{frac{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}-frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}{Delta x}}&=\=lim_{Delta xto 0}{frac{mathrm f(x)-2cdot mathrm f(x-Delta x)+mathrm f(x-2cdotDelta x)}{Delta x^2}}
                end{align*}
                $$
                we can easy check this limit for any (?) functions. Similarly, we can define the first derivative for right, middle and left points:
                $$
                frac{d}{dx}textrm{f}(x)left{
                begin{aligned}
                &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-Delta x)}{Delta x}}
                \
                &=lim_{Delta xto 0}{frac{mathrm f(x-Delta x)-mathrm f(x-2cdotDelta x)}{Delta x}}\
                &=lim_{Delta xto 0}{frac{mathrm f(x)-mathrm f(x-2cdotDelta x)}{2cdotDelta x}}
                end{aligned}
                right.
                $$
                Let's use the finite elements method assuming $Td=Delta x,,y_n=mathrm f(x),,y_{n-1}=mathrm f(x-Delta x),,y_{n-2}=mathrm f(x-2cdot Delta x)$
                Override differential equation as
                $$
                frac{y_n-2cdot y_{n-1}+y_{n-2}}{Td^2}=-y_n
                $$
                Now solve this implicit equation for $y_n$ to obtain explicit recurrence relation:
                $$
                y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                $$
                Using arbitrarily small but non-zero quantity Td we can plot exponentially decaying sampled sine function (because the poles are inside the unit circle of the transfer function corresponding to the given recurrence relation).
                Similarly we write three systems for the initial conditions:



                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_n-y_{n-1}}{Td}
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_{n-1}-y_{n-2}}{Td}
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_n = frac{2cdot y_{n-1}-y_{n-2}}{1+Td^2}
                \
                &C=frac{y_n-y_{n-2}}{2cdot Td}
                end{aligned}right.
                $$
                Solve this sequence of equations for $y_{n-1}$ and $y_{n-2}$:
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1-Td^2right)
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+Td^2right)\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}cdotleft(1+Td^2right)
                end{aligned}right.
                $$
                $$
                left{
                begin{aligned}
                &y_{n-1} = -Ccdot Td + y_{n}cdotleft(1+frac{Td^2}{2}right)\
                &y_{n-2}=-2cdot Ccdot Td + y_{n}
                end{aligned}right.
                $$
                At zero point $y_n=mathrm f(0)=0$ and we can see linear dependence:
                $$
                begin{aligned}
                &y_{n-1} = -Ccdot Td\
                &y_{n-2}=-2cdot Ccdot Td
                end{aligned}
                $$
                for all three solutions. Replace back:
                $$
                begin{array}{l}
                mathrm f(0)&=0\
                mathrm f(0-Delta x) &= -Ccdot Delta x\
                mathrm f(0-2cdot Delta x) &= -2cdot Ccdot Delta x
                end{array}
                $$
                So all three $frac{d}{dx}mathrm f(0)$ limits is equal to $C$ at $x=0$ and in accordance with $mathrm f(x)=Ccdotsin(x)$ by definition we can write
                $$
                lim_{Delta xto 0}{frac{mathrm f(0)-mathrm f(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{0-(-C cdot Delta x)}{Delta x}}=C
                $$
                Thus
                $$
                lim_{Delta xto 0}{frac{sin(0)-Ccdotsin(0-Delta x)}{Delta x}}=lim_{Delta xto 0}{frac{Ccdotsin(Delta x)}{Delta x}}=Ccdotlim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=C
                $$
                and $lim_{Delta xto 0}{frac{sin(Delta x)}{Delta x}}=1$







                share|cite|improve this answer












                share|cite|improve this answer



                share|cite|improve this answer










                answered Aug 19 '16 at 20:14









                Timur ZhoraevTimur Zhoraev

                30436




                30436












                • $begingroup$
                  While this is an exceptionally neat answer, it is not really appropriate for the level (high school) on which I asked the question, given that computing this limit is an introductory exercise into differentials. Plus we defined $sin$ geometrically (i.e. though the relationship between angles and sides in a right triangle), so your prerequisites must be established first.
                  $endgroup$
                  – FUZxxl
                  Aug 19 '16 at 20:40










                • $begingroup$
                  @FUZxxl do you realize that users with different levels and mathematical background may check this question? The answer given above might be helpful for them.
                  $endgroup$
                  – Xam
                  Sep 23 '17 at 3:38


















                • $begingroup$
                  While this is an exceptionally neat answer, it is not really appropriate for the level (high school) on which I asked the question, given that computing this limit is an introductory exercise into differentials. Plus we defined $sin$ geometrically (i.e. though the relationship between angles and sides in a right triangle), so your prerequisites must be established first.
                  $endgroup$
                  – FUZxxl
                  Aug 19 '16 at 20:40










                • $begingroup$
                  @FUZxxl do you realize that users with different levels and mathematical background may check this question? The answer given above might be helpful for them.
                  $endgroup$
                  – Xam
                  Sep 23 '17 at 3:38
















                $begingroup$
                While this is an exceptionally neat answer, it is not really appropriate for the level (high school) on which I asked the question, given that computing this limit is an introductory exercise into differentials. Plus we defined $sin$ geometrically (i.e. though the relationship between angles and sides in a right triangle), so your prerequisites must be established first.
                $endgroup$
                – FUZxxl
                Aug 19 '16 at 20:40




                $begingroup$
                While this is an exceptionally neat answer, it is not really appropriate for the level (high school) on which I asked the question, given that computing this limit is an introductory exercise into differentials. Plus we defined $sin$ geometrically (i.e. though the relationship between angles and sides in a right triangle), so your prerequisites must be established first.
                $endgroup$
                – FUZxxl
                Aug 19 '16 at 20:40












                $begingroup$
                @FUZxxl do you realize that users with different levels and mathematical background may check this question? The answer given above might be helpful for them.
                $endgroup$
                – Xam
                Sep 23 '17 at 3:38




                $begingroup$
                @FUZxxl do you realize that users with different levels and mathematical background may check this question? The answer given above might be helpful for them.
                $endgroup$
                – Xam
                Sep 23 '17 at 3:38











                11












                $begingroup$

                Simple one is using sandwich theorem Which demonstrated earlier.In this method you have to show that $frac{sin x}{x} $ lies between other two functions. As $x longrightarrow 0$ both of them will tends to ONE.



                Then as in the case of sandwich (if both the bread part go to one stomach the middle portion will also go to the same stomach) $frac{sin x}{x}$ will go to ONE.



                You can use geogebra to see the visualization of this phenomena using geogebra.First you input $sin x$ and $x$ and observe that near to $0$ values of $sin x$ and $x$ are same.



                Secondly input $frac{sin x}{x}$ then observe function is approaching to $1$ as $x$ tends to $0$






                share|cite|improve this answer











                $endgroup$









                • 4




                  $begingroup$
                  This describes the Sandwich Theorem, but does not answer the question. At best, this should be a comment to the question,
                  $endgroup$
                  – robjohn
                  Dec 13 '14 at 13:31
















                11












                $begingroup$

                Simple one is using sandwich theorem Which demonstrated earlier.In this method you have to show that $frac{sin x}{x} $ lies between other two functions. As $x longrightarrow 0$ both of them will tends to ONE.



                Then as in the case of sandwich (if both the bread part go to one stomach the middle portion will also go to the same stomach) $frac{sin x}{x}$ will go to ONE.



                You can use geogebra to see the visualization of this phenomena using geogebra.First you input $sin x$ and $x$ and observe that near to $0$ values of $sin x$ and $x$ are same.



                Secondly input $frac{sin x}{x}$ then observe function is approaching to $1$ as $x$ tends to $0$






                share|cite|improve this answer











                $endgroup$









                • 4




                  $begingroup$
                  This describes the Sandwich Theorem, but does not answer the question. At best, this should be a comment to the question,
                  $endgroup$
                  – robjohn
                  Dec 13 '14 at 13:31














                11












                11








                11





                $begingroup$

                Simple one is using sandwich theorem Which demonstrated earlier.In this method you have to show that $frac{sin x}{x} $ lies between other two functions. As $x longrightarrow 0$ both of them will tends to ONE.



                Then as in the case of sandwich (if both the bread part go to one stomach the middle portion will also go to the same stomach) $frac{sin x}{x}$ will go to ONE.



                You can use geogebra to see the visualization of this phenomena using geogebra.First you input $sin x$ and $x$ and observe that near to $0$ values of $sin x$ and $x$ are same.



                Secondly input $frac{sin x}{x}$ then observe function is approaching to $1$ as $x$ tends to $0$






                share|cite|improve this answer











                $endgroup$



                Simple one is using sandwich theorem Which demonstrated earlier.In this method you have to show that $frac{sin x}{x} $ lies between other two functions. As $x longrightarrow 0$ both of them will tends to ONE.



                Then as in the case of sandwich (if both the bread part go to one stomach the middle portion will also go to the same stomach) $frac{sin x}{x}$ will go to ONE.



                You can use geogebra to see the visualization of this phenomena using geogebra.First you input $sin x$ and $x$ and observe that near to $0$ values of $sin x$ and $x$ are same.



                Secondly input $frac{sin x}{x}$ then observe function is approaching to $1$ as $x$ tends to $0$







                share|cite|improve this answer














                share|cite|improve this answer



                share|cite|improve this answer








                edited Dec 13 '14 at 10:00









                Davide Giraudo

                127k16151264




                127k16151264










                answered Sep 15 '14 at 2:56









                MadhuMadhu

                7291925




                7291925








                • 4




                  $begingroup$
                  This describes the Sandwich Theorem, but does not answer the question. At best, this should be a comment to the question,
                  $endgroup$
                  – robjohn
                  Dec 13 '14 at 13:31














                • 4




                  $begingroup$
                  This describes the Sandwich Theorem, but does not answer the question. At best, this should be a comment to the question,
                  $endgroup$
                  – robjohn
                  Dec 13 '14 at 13:31








                4




                4




                $begingroup$
                This describes the Sandwich Theorem, but does not answer the question. At best, this should be a comment to the question,
                $endgroup$
                – robjohn
                Dec 13 '14 at 13:31




                $begingroup$
                This describes the Sandwich Theorem, but does not answer the question. At best, this should be a comment to the question,
                $endgroup$
                – robjohn
                Dec 13 '14 at 13:31











                8












                $begingroup$

                Originally posted on the proofs without words post, here is a simple image that explains the derivative of $sin(x)$, which as we all know, is directly related to the limit at hand.



                enter image description here



                If one is not so convinced, take a look at the above picture and notice that if $upm h$ is in the first quadrant, then



                $$frac{sin(x+h)-sin(x)}h<cos(x)<frac{sin(x-h)-sin(x)}{-h}$$





                Notice that



                $$
                begin{align}frac{d}{dx}sin(x)&=lim_{hto0}frac{sin(x+h)-sin(x)}h\text{picture above}&=lim_{hto0}frac{sin(x)cos(h)+cos(x)sin(h)-sin(x)}h\cos(x)&=lim_{hto0}sin(x)frac{cos(h)-1}h+cos(x)frac{sin(h)}h\cos(0)&=lim_{hto0}frac{sin(h)}hend{align}
                $$






                share|cite|improve this answer











                $endgroup$









                • 2




                  $begingroup$
                  Note that your reason is circular: In order to prove the derivative of $sin$, we need to know the limit of $sin hover h.$ Attempting to prove the limit the other way round is counterproductive.
                  $endgroup$
                  – FUZxxl
                  Nov 16 '16 at 0:47






                • 1




                  $begingroup$
                  @FUZxxl no actually, the whole point of this was a geometric proof of the fact.
                  $endgroup$
                  – Simply Beautiful Art
                  Nov 16 '16 at 17:23










                • $begingroup$
                  @FUZxxl If you are still unsatisfied about it, there is a whole post on the topic.
                  $endgroup$
                  – Simply Beautiful Art
                  Nov 17 '16 at 0:44










                • $begingroup$
                  @FUZxxl Since you still seem unhappy, I've added a better "squeeze theorem" type thing for the derivative of $sin(x)$.
                  $endgroup$
                  – Simply Beautiful Art
                  Jan 5 '17 at 23:14
















                8












                $begingroup$

                Originally posted on the proofs without words post, here is a simple image that explains the derivative of $sin(x)$, which as we all know, is directly related to the limit at hand.



                enter image description here



                If one is not so convinced, take a look at the above picture and notice that if $upm h$ is in the first quadrant, then



                $$frac{sin(x+h)-sin(x)}h<cos(x)<frac{sin(x-h)-sin(x)}{-h}$$





                Notice that



                $$
                begin{align}frac{d}{dx}sin(x)&=lim_{hto0}frac{sin(x+h)-sin(x)}h\text{picture above}&=lim_{hto0}frac{sin(x)cos(h)+cos(x)sin(h)-sin(x)}h\cos(x)&=lim_{hto0}sin(x)frac{cos(h)-1}h+cos(x)frac{sin(h)}h\cos(0)&=lim_{hto0}frac{sin(h)}hend{align}
                $$






                share|cite|improve this answer











                $endgroup$









                • 2




                  $begingroup$
                  Note that your reason is circular: In order to prove the derivative of $sin$, we need to know the limit of $sin hover h.$ Attempting to prove the limit the other way round is counterproductive.
                  $endgroup$
                  – FUZxxl
                  Nov 16 '16 at 0:47






                • 1




                  $begingroup$
                  @FUZxxl no actually, the whole point of this was a geometric proof of the fact.
                  $endgroup$
                  – Simply Beautiful Art
                  Nov 16 '16 at 17:23










                • $begingroup$
                  @FUZxxl If you are still unsatisfied about it, there is a whole post on the topic.
                  $endgroup$
                  – Simply Beautiful Art
                  Nov 17 '16 at 0:44










                • $begingroup$
                  @FUZxxl Since you still seem unhappy, I've added a better "squeeze theorem" type thing for the derivative of $sin(x)$.
                  $endgroup$
                  – Simply Beautiful Art
                  Jan 5 '17 at 23:14














                8












                8








                8





                $begingroup$

                Originally posted on the proofs without words post, here is a simple image that explains the derivative of $sin(x)$, which as we all know, is directly related to the limit at hand.



                enter image description here



                If one is not so convinced, take a look at the above picture and notice that if $upm h$ is in the first quadrant, then



                $$frac{sin(x+h)-sin(x)}h<cos(x)<frac{sin(x-h)-sin(x)}{-h}$$





                Notice that



                $$
                begin{align}frac{d}{dx}sin(x)&=lim_{hto0}frac{sin(x+h)-sin(x)}h\text{picture above}&=lim_{hto0}frac{sin(x)cos(h)+cos(x)sin(h)-sin(x)}h\cos(x)&=lim_{hto0}sin(x)frac{cos(h)-1}h+cos(x)frac{sin(h)}h\cos(0)&=lim_{hto0}frac{sin(h)}hend{align}
                $$






                share|cite|improve this answer











                $endgroup$



                Originally posted on the proofs without words post, here is a simple image that explains the derivative of $sin(x)$, which as we all know, is directly related to the limit at hand.



                enter image description here



                If one is not so convinced, take a look at the above picture and notice that if $upm h$ is in the first quadrant, then



                $$frac{sin(x+h)-sin(x)}h<cos(x)<frac{sin(x-h)-sin(x)}{-h}$$





                Notice that



                $$
                begin{align}frac{d}{dx}sin(x)&=lim_{hto0}frac{sin(x+h)-sin(x)}h\text{picture above}&=lim_{hto0}frac{sin(x)cos(h)+cos(x)sin(h)-sin(x)}h\cos(x)&=lim_{hto0}sin(x)frac{cos(h)-1}h+cos(x)frac{sin(h)}h\cos(0)&=lim_{hto0}frac{sin(h)}hend{align}
                $$







                share|cite|improve this answer














                share|cite|improve this answer



                share|cite|improve this answer








                edited Apr 13 '17 at 12:58









                Community

                1




                1










                answered Nov 15 '16 at 22:52









                Simply Beautiful ArtSimply Beautiful Art

                50.6k579183




                50.6k579183








                • 2




                  $begingroup$
                  Note that your reason is circular: In order to prove the derivative of $sin$, we need to know the limit of $sin hover h.$ Attempting to prove the limit the other way round is counterproductive.
                  $endgroup$
                  – FUZxxl
                  Nov 16 '16 at 0:47






                • 1




                  $begingroup$
                  @FUZxxl no actually, the whole point of this was a geometric proof of the fact.
                  $endgroup$
                  – Simply Beautiful Art
                  Nov 16 '16 at 17:23










                • $begingroup$
                  @FUZxxl If you are still unsatisfied about it, there is a whole post on the topic.
                  $endgroup$
                  – Simply Beautiful Art
                  Nov 17 '16 at 0:44










                • $begingroup$
                  @FUZxxl Since you still seem unhappy, I've added a better "squeeze theorem" type thing for the derivative of $sin(x)$.
                  $endgroup$
                  – Simply Beautiful Art
                  Jan 5 '17 at 23:14














                • 2




                  $begingroup$
                  Note that your reason is circular: In order to prove the derivative of $sin$, we need to know the limit of $sin hover h.$ Attempting to prove the limit the other way round is counterproductive.
                  $endgroup$
                  – FUZxxl
                  Nov 16 '16 at 0:47






                • 1




                  $begingroup$
                  @FUZxxl no actually, the whole point of this was a geometric proof of the fact.
                  $endgroup$
                  – Simply Beautiful Art
                  Nov 16 '16 at 17:23










                • $begingroup$
                  @FUZxxl If you are still unsatisfied about it, there is a whole post on the topic.
                  $endgroup$
                  – Simply Beautiful Art
                  Nov 17 '16 at 0:44










                • $begingroup$
                  @FUZxxl Since you still seem unhappy, I've added a better "squeeze theorem" type thing for the derivative of $sin(x)$.
                  $endgroup$
                  – Simply Beautiful Art
                  Jan 5 '17 at 23:14








                2




                2




                $begingroup$
                Note that your reason is circular: In order to prove the derivative of $sin$, we need to know the limit of $sin hover h.$ Attempting to prove the limit the other way round is counterproductive.
                $endgroup$
                – FUZxxl
                Nov 16 '16 at 0:47




                $begingroup$
                Note that your reason is circular: In order to prove the derivative of $sin$, we need to know the limit of $sin hover h.$ Attempting to prove the limit the other way round is counterproductive.
                $endgroup$
                – FUZxxl
                Nov 16 '16 at 0:47




                1




                1




                $begingroup$
                @FUZxxl no actually, the whole point of this was a geometric proof of the fact.
                $endgroup$
                – Simply Beautiful Art
                Nov 16 '16 at 17:23




                $begingroup$
                @FUZxxl no actually, the whole point of this was a geometric proof of the fact.
                $endgroup$
                – Simply Beautiful Art
                Nov 16 '16 at 17:23












                $begingroup$
                @FUZxxl If you are still unsatisfied about it, there is a whole post on the topic.
                $endgroup$
                – Simply Beautiful Art
                Nov 17 '16 at 0:44




                $begingroup$
                @FUZxxl If you are still unsatisfied about it, there is a whole post on the topic.
                $endgroup$
                – Simply Beautiful Art
                Nov 17 '16 at 0:44












                $begingroup$
                @FUZxxl Since you still seem unhappy, I've added a better "squeeze theorem" type thing for the derivative of $sin(x)$.
                $endgroup$
                – Simply Beautiful Art
                Jan 5 '17 at 23:14




                $begingroup$
                @FUZxxl Since you still seem unhappy, I've added a better "squeeze theorem" type thing for the derivative of $sin(x)$.
                $endgroup$
                – Simply Beautiful Art
                Jan 5 '17 at 23:14











                4












                $begingroup$

                Here is another approach.



                enter image description here(1)
                picture 2(2)



                In the large triangle, $$tan(theta)=frac{opp}{adg}=frac{z}{1}=z$$ So the triangle has height $$z=tan(theta)$$ and base $1$
                so it's area is $$Area(big triangle)=frac{1}{2}z=frac{1}{2}tan(theta)$$



                Next the sector area as a fraction of the entire circle, the sector is (see the right hand side of picture (1))$$frac{theta}{2pi}$$ of the entire circle so it's area is



                $$Area(sector)=frac{theta}{2pi}*(pi)(1)^2=frac{theta}{2}$$
                The triangle within the sector has height $y$. But $y=frac{y}{1}=frac{opp}{hyp}=sin(theta)$ so the small triangle has height $y=sin(theta)$ and base $1$ so its area is $$Area(small triangle)=frac{1}{2}y=frac{1}{2}sin(theta)$$
                Now we can use the sandwich theorem as $$Area(big triangle)geq Area(sector)geq Area(small triangle)$$



                using the equations we worked out above this becomes



                $$frac{tan(theta)}{2}geqfrac{theta}{2}geqfrac{sin(theta)}{2}$$ now multipliying through by two and using the fact that $$tan(theta)=frac{sin(theta)}{cos(theta)}$$ we get that
                $$frac{sin(theta)}{cos(theta)}geqthetageqsin(theta)$$
                taking reciprocals changes the inequalities so we have that
                $$frac{cos(theta)}{sin(theta)}leqfrac{1}{theta}leqfrac{1}{sin(theta)}$$
                now finally multiplying through by $sin(theta)$ we get $$cos(theta)leqfrac{sin(theta)}{theta}leq1$$
                Now $$lim limits_{theta to 0}cos(theta)=1$$ and$$lim limits_{theta to 0}1=1$$



                so by the sandwhich theorem $$lim limits_{theta to 0}frac{sin(theta)}{theta}=1$$ also. QED






                share|cite|improve this answer









                $endgroup$


















                  4












                  $begingroup$

                  Here is another approach.



                  enter image description here(1)
                  picture 2(2)



                  In the large triangle, $$tan(theta)=frac{opp}{adg}=frac{z}{1}=z$$ So the triangle has height $$z=tan(theta)$$ and base $1$
                  so it's area is $$Area(big triangle)=frac{1}{2}z=frac{1}{2}tan(theta)$$



                  Next the sector area as a fraction of the entire circle, the sector is (see the right hand side of picture (1))$$frac{theta}{2pi}$$ of the entire circle so it's area is



                  $$Area(sector)=frac{theta}{2pi}*(pi)(1)^2=frac{theta}{2}$$
                  The triangle within the sector has height $y$. But $y=frac{y}{1}=frac{opp}{hyp}=sin(theta)$ so the small triangle has height $y=sin(theta)$ and base $1$ so its area is $$Area(small triangle)=frac{1}{2}y=frac{1}{2}sin(theta)$$
                  Now we can use the sandwich theorem as $$Area(big triangle)geq Area(sector)geq Area(small triangle)$$



                  using the equations we worked out above this becomes



                  $$frac{tan(theta)}{2}geqfrac{theta}{2}geqfrac{sin(theta)}{2}$$ now multipliying through by two and using the fact that $$tan(theta)=frac{sin(theta)}{cos(theta)}$$ we get that
                  $$frac{sin(theta)}{cos(theta)}geqthetageqsin(theta)$$
                  taking reciprocals changes the inequalities so we have that
                  $$frac{cos(theta)}{sin(theta)}leqfrac{1}{theta}leqfrac{1}{sin(theta)}$$
                  now finally multiplying through by $sin(theta)$ we get $$cos(theta)leqfrac{sin(theta)}{theta}leq1$$
                  Now $$lim limits_{theta to 0}cos(theta)=1$$ and$$lim limits_{theta to 0}1=1$$



                  so by the sandwhich theorem $$lim limits_{theta to 0}frac{sin(theta)}{theta}=1$$ also. QED






                  share|cite|improve this answer









                  $endgroup$
















                    4












                    4








                    4





                    $begingroup$

                    Here is another approach.



                    enter image description here(1)
                    picture 2(2)



                    In the large triangle, $$tan(theta)=frac{opp}{adg}=frac{z}{1}=z$$ So the triangle has height $$z=tan(theta)$$ and base $1$
                    so it's area is $$Area(big triangle)=frac{1}{2}z=frac{1}{2}tan(theta)$$



                    Next the sector area as a fraction of the entire circle, the sector is (see the right hand side of picture (1))$$frac{theta}{2pi}$$ of the entire circle so it's area is



                    $$Area(sector)=frac{theta}{2pi}*(pi)(1)^2=frac{theta}{2}$$
                    The triangle within the sector has height $y$. But $y=frac{y}{1}=frac{opp}{hyp}=sin(theta)$ so the small triangle has height $y=sin(theta)$ and base $1$ so its area is $$Area(small triangle)=frac{1}{2}y=frac{1}{2}sin(theta)$$
                    Now we can use the sandwich theorem as $$Area(big triangle)geq Area(sector)geq Area(small triangle)$$



                    using the equations we worked out above this becomes



                    $$frac{tan(theta)}{2}geqfrac{theta}{2}geqfrac{sin(theta)}{2}$$ now multipliying through by two and using the fact that $$tan(theta)=frac{sin(theta)}{cos(theta)}$$ we get that
                    $$frac{sin(theta)}{cos(theta)}geqthetageqsin(theta)$$
                    taking reciprocals changes the inequalities so we have that
                    $$frac{cos(theta)}{sin(theta)}leqfrac{1}{theta}leqfrac{1}{sin(theta)}$$
                    now finally multiplying through by $sin(theta)$ we get $$cos(theta)leqfrac{sin(theta)}{theta}leq1$$
                    Now $$lim limits_{theta to 0}cos(theta)=1$$ and$$lim limits_{theta to 0}1=1$$



                    so by the sandwhich theorem $$lim limits_{theta to 0}frac{sin(theta)}{theta}=1$$ also. QED






                    share|cite|improve this answer









                    $endgroup$



                    Here is another approach.



                    enter image description here(1)
                    picture 2(2)



                    In the large triangle, $$tan(theta)=frac{opp}{adg}=frac{z}{1}=z$$ So the triangle has height $$z=tan(theta)$$ and base $1$
                    so it's area is $$Area(big triangle)=frac{1}{2}z=frac{1}{2}tan(theta)$$



                    Next the sector area as a fraction of the entire circle, the sector is (see the right hand side of picture (1))$$frac{theta}{2pi}$$ of the entire circle so it's area is



                    $$Area(sector)=frac{theta}{2pi}*(pi)(1)^2=frac{theta}{2}$$
                    The triangle within the sector has height $y$. But $y=frac{y}{1}=frac{opp}{hyp}=sin(theta)$ so the small triangle has height $y=sin(theta)$ and base $1$ so its area is $$Area(small triangle)=frac{1}{2}y=frac{1}{2}sin(theta)$$
                    Now we can use the sandwich theorem as $$Area(big triangle)geq Area(sector)geq Area(small triangle)$$



                    using the equations we worked out above this becomes



                    $$frac{tan(theta)}{2}geqfrac{theta}{2}geqfrac{sin(theta)}{2}$$ now multipliying through by two and using the fact that $$tan(theta)=frac{sin(theta)}{cos(theta)}$$ we get that
                    $$frac{sin(theta)}{cos(theta)}geqthetageqsin(theta)$$
                    taking reciprocals changes the inequalities so we have that
                    $$frac{cos(theta)}{sin(theta)}leqfrac{1}{theta}leqfrac{1}{sin(theta)}$$
                    now finally multiplying through by $sin(theta)$ we get $$cos(theta)leqfrac{sin(theta)}{theta}leq1$$
                    Now $$lim limits_{theta to 0}cos(theta)=1$$ and$$lim limits_{theta to 0}1=1$$



                    so by the sandwhich theorem $$lim limits_{theta to 0}frac{sin(theta)}{theta}=1$$ also. QED







                    share|cite|improve this answer












                    share|cite|improve this answer



                    share|cite|improve this answer










                    answered Mar 30 '17 at 19:30







                    user395952






























                        2












                        $begingroup$

                        We may use a linear approximation for this limit. The following expression comes straight out of the equation of a line that is $y=mx+b$. Using the fact that $m=dfrac{d}{dx}[f(x)]$ $$f(x) approx f(a) + f'(a)(x-a)$$In this case $f(x)=sin x$ and $a=0$ $implies sin x approx f(0)+cos 0 (x-0)=x longleftrightarrow sin x approx x$. The following graph better illustrates this tangent line approximation.





                        Since, $sin x approx x implies displaystyle lim_{x to 0} dfrac{sin x}{x}=1$. Quod Erat Demonstrandum






                        share|cite|improve this answer









                        $endgroup$









                        • 2




                          $begingroup$
                          This is not a proof unless you can justify all these approximation steps.
                          $endgroup$
                          – FUZxxl
                          Dec 4 '18 at 12:33










                        • $begingroup$
                          To find that $frac{mathrm{d}}{mathrm{d}x}sin(x)=cos(x)$ for the Taylor Series usually requires knowing the limit we are trying to find. In fact, $$left.frac{mathrm{d}}{mathrm{d}x}sin(x)right|_{x=0}=lim_{xto0}frac{sin(x)-sin(0)}{x-0}=lim_{xto0}frac{sin(x)}{x}$$
                          $endgroup$
                          – robjohn
                          Dec 7 '18 at 20:20


















                        2












                        $begingroup$

                        We may use a linear approximation for this limit. The following expression comes straight out of the equation of a line that is $y=mx+b$. Using the fact that $m=dfrac{d}{dx}[f(x)]$ $$f(x) approx f(a) + f'(a)(x-a)$$In this case $f(x)=sin x$ and $a=0$ $implies sin x approx f(0)+cos 0 (x-0)=x longleftrightarrow sin x approx x$. The following graph better illustrates this tangent line approximation.





                        Since, $sin x approx x implies displaystyle lim_{x to 0} dfrac{sin x}{x}=1$. Quod Erat Demonstrandum






                        share|cite|improve this answer









                        $endgroup$









                        • 2




                          $begingroup$
                          This is not a proof unless you can justify all these approximation steps.
                          $endgroup$
                          – FUZxxl
                          Dec 4 '18 at 12:33










                        • $begingroup$
                          To find that $frac{mathrm{d}}{mathrm{d}x}sin(x)=cos(x)$ for the Taylor Series usually requires knowing the limit we are trying to find. In fact, $$left.frac{mathrm{d}}{mathrm{d}x}sin(x)right|_{x=0}=lim_{xto0}frac{sin(x)-sin(0)}{x-0}=lim_{xto0}frac{sin(x)}{x}$$
                          $endgroup$
                          – robjohn
                          Dec 7 '18 at 20:20
















                        2












                        2








                        2





                        $begingroup$

                        We may use a linear approximation for this limit. The following expression comes straight out of the equation of a line that is $y=mx+b$. Using the fact that $m=dfrac{d}{dx}[f(x)]$ $$f(x) approx f(a) + f'(a)(x-a)$$In this case $f(x)=sin x$ and $a=0$ $implies sin x approx f(0)+cos 0 (x-0)=x longleftrightarrow sin x approx x$. The following graph better illustrates this tangent line approximation.





                        Since, $sin x approx x implies displaystyle lim_{x to 0} dfrac{sin x}{x}=1$. Quod Erat Demonstrandum






                        share|cite|improve this answer









                        $endgroup$



                        We may use a linear approximation for this limit. The following expression comes straight out of the equation of a line that is $y=mx+b$. Using the fact that $m=dfrac{d}{dx}[f(x)]$ $$f(x) approx f(a) + f'(a)(x-a)$$In this case $f(x)=sin x$ and $a=0$ $implies sin x approx f(0)+cos 0 (x-0)=x longleftrightarrow sin x approx x$. The following graph better illustrates this tangent line approximation.





                        Since, $sin x approx x implies displaystyle lim_{x to 0} dfrac{sin x}{x}=1$. Quod Erat Demonstrandum







                        share|cite|improve this answer












                        share|cite|improve this answer



                        share|cite|improve this answer










                        answered Dec 4 '18 at 5:17









                        Paras KhoslaParas Khosla

                        1,075216




                        1,075216








                        • 2




                          $begingroup$
                          This is not a proof unless you can justify all these approximation steps.
                          $endgroup$
                          – FUZxxl
                          Dec 4 '18 at 12:33










                        • $begingroup$
                          To find that $frac{mathrm{d}}{mathrm{d}x}sin(x)=cos(x)$ for the Taylor Series usually requires knowing the limit we are trying to find. In fact, $$left.frac{mathrm{d}}{mathrm{d}x}sin(x)right|_{x=0}=lim_{xto0}frac{sin(x)-sin(0)}{x-0}=lim_{xto0}frac{sin(x)}{x}$$
                          $endgroup$
                          – robjohn
                          Dec 7 '18 at 20:20
















                        • 2




                          $begingroup$
                          This is not a proof unless you can justify all these approximation steps.
                          $endgroup$
                          – FUZxxl
                          Dec 4 '18 at 12:33










                        • $begingroup$
                          To find that $frac{mathrm{d}}{mathrm{d}x}sin(x)=cos(x)$ for the Taylor Series usually requires knowing the limit we are trying to find. In fact, $$left.frac{mathrm{d}}{mathrm{d}x}sin(x)right|_{x=0}=lim_{xto0}frac{sin(x)-sin(0)}{x-0}=lim_{xto0}frac{sin(x)}{x}$$
                          $endgroup$
                          – robjohn
                          Dec 7 '18 at 20:20










                        2




                        2




                        $begingroup$
                        This is not a proof unless you can justify all these approximation steps.
                        $endgroup$
                        – FUZxxl
                        Dec 4 '18 at 12:33




                        $begingroup$
                        This is not a proof unless you can justify all these approximation steps.
                        $endgroup$
                        – FUZxxl
                        Dec 4 '18 at 12:33












                        $begingroup$
                        To find that $frac{mathrm{d}}{mathrm{d}x}sin(x)=cos(x)$ for the Taylor Series usually requires knowing the limit we are trying to find. In fact, $$left.frac{mathrm{d}}{mathrm{d}x}sin(x)right|_{x=0}=lim_{xto0}frac{sin(x)-sin(0)}{x-0}=lim_{xto0}frac{sin(x)}{x}$$
                        $endgroup$
                        – robjohn
                        Dec 7 '18 at 20:20






                        $begingroup$
                        To find that $frac{mathrm{d}}{mathrm{d}x}sin(x)=cos(x)$ for the Taylor Series usually requires knowing the limit we are trying to find. In fact, $$left.frac{mathrm{d}}{mathrm{d}x}sin(x)right|_{x=0}=lim_{xto0}frac{sin(x)-sin(0)}{x-0}=lim_{xto0}frac{sin(x)}{x}$$
                        $endgroup$
                        – robjohn
                        Dec 7 '18 at 20:20













                        1












                        $begingroup$

                        Here is a proof to those familiar with power series.



                        The definition of $sin(x)$ is



                        $$sin(x) = sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}h^{2k+1}$$



                        Therefore we get



                        $$begin{align} lim_{x to 0} frac{sin(x)}{x} &= lim_{x to 0} frac{sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}x^{2k+1}}{x} \&= lim_{x to 0} sum_{k=0}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 + lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 end{align}$$



                        where we have used the fact that the power series $sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k}$ has radius of convergence $R=infty$ and therefore is continuous on $mathbb R$. This allows us to take the limit inside and we get



                        $$lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} = sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} 0^{2k} = 0$$






                        share|cite|improve this answer









                        $endgroup$













                        • $begingroup$
                          Compare to this answer and this answer, which both use power series.
                          $endgroup$
                          – robjohn
                          Jan 14 '18 at 19:48






                        • 1




                          $begingroup$
                          The premise of the question is not to use power series.
                          $endgroup$
                          – FUZxxl
                          Dec 4 '18 at 12:34
















                        1












                        $begingroup$

                        Here is a proof to those familiar with power series.



                        The definition of $sin(x)$ is



                        $$sin(x) = sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}h^{2k+1}$$



                        Therefore we get



                        $$begin{align} lim_{x to 0} frac{sin(x)}{x} &= lim_{x to 0} frac{sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}x^{2k+1}}{x} \&= lim_{x to 0} sum_{k=0}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 + lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 end{align}$$



                        where we have used the fact that the power series $sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k}$ has radius of convergence $R=infty$ and therefore is continuous on $mathbb R$. This allows us to take the limit inside and we get



                        $$lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} = sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} 0^{2k} = 0$$






                        share|cite|improve this answer









                        $endgroup$













                        • $begingroup$
                          Compare to this answer and this answer, which both use power series.
                          $endgroup$
                          – robjohn
                          Jan 14 '18 at 19:48






                        • 1




                          $begingroup$
                          The premise of the question is not to use power series.
                          $endgroup$
                          – FUZxxl
                          Dec 4 '18 at 12:34














                        1












                        1








                        1





                        $begingroup$

                        Here is a proof to those familiar with power series.



                        The definition of $sin(x)$ is



                        $$sin(x) = sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}h^{2k+1}$$



                        Therefore we get



                        $$begin{align} lim_{x to 0} frac{sin(x)}{x} &= lim_{x to 0} frac{sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}x^{2k+1}}{x} \&= lim_{x to 0} sum_{k=0}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 + lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 end{align}$$



                        where we have used the fact that the power series $sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k}$ has radius of convergence $R=infty$ and therefore is continuous on $mathbb R$. This allows us to take the limit inside and we get



                        $$lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} = sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} 0^{2k} = 0$$






                        share|cite|improve this answer









                        $endgroup$



                        Here is a proof to those familiar with power series.



                        The definition of $sin(x)$ is



                        $$sin(x) = sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}h^{2k+1}$$



                        Therefore we get



                        $$begin{align} lim_{x to 0} frac{sin(x)}{x} &= lim_{x to 0} frac{sum_{k=0}^infty frac{(-1)^k}{(2k+1)!}x^{2k+1}}{x} \&= lim_{x to 0} sum_{k=0}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 + lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} \&= 1 end{align}$$



                        where we have used the fact that the power series $sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k}$ has radius of convergence $R=infty$ and therefore is continuous on $mathbb R$. This allows us to take the limit inside and we get



                        $$lim_{x to 0} sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} x^{2k} = sum_{k=1}^infty frac{(-1)^k}{(2k+1)!} 0^{2k} = 0$$







                        share|cite|improve this answer












                        share|cite|improve this answer



                        share|cite|improve this answer










                        answered Jan 14 '18 at 18:08









                        philmcolephilmcole

                        841419




                        841419












                        • $begingroup$
                          Compare to this answer and this answer, which both use power series.
                          $endgroup$
                          – robjohn
                          Jan 14 '18 at 19:48






                        • 1




                          $begingroup$
                          The premise of the question is not to use power series.
                          $endgroup$
                          – FUZxxl
                          Dec 4 '18 at 12:34


















                        • $begingroup$
                          Compare to this answer and this answer, which both use power series.
                          $endgroup$
                          – robjohn
                          Jan 14 '18 at 19:48






                        • 1




                          $begingroup$
                          The premise of the question is not to use power series.
                          $endgroup$
                          – FUZxxl
                          Dec 4 '18 at 12:34
















                        $begingroup$
                        Compare to this answer and this answer, which both use power series.
                        $endgroup$
                        – robjohn
                        Jan 14 '18 at 19:48




                        $begingroup$
                        Compare to this answer and this answer, which both use power series.
                        $endgroup$
                        – robjohn
                        Jan 14 '18 at 19:48




                        1




                        1




                        $begingroup$
                        The premise of the question is not to use power series.
                        $endgroup$
                        – FUZxxl
                        Dec 4 '18 at 12:34




                        $begingroup$
                        The premise of the question is not to use power series.
                        $endgroup$
                        – FUZxxl
                        Dec 4 '18 at 12:34











                        1












                        $begingroup$

                        We can also use Euler's formula to prove the limit:



                        $$e^{ix} = cos x + isin x$$



                        $$lim_{xto 0}dfrac{sin x}{x} = implies lim_{xto 0} dfrac{e^{ix}- e^{-ix}}{2i x}$$



                        $$= lim_{xto 0} dfrac{e^{2ix}-1}{2ix}timesdfrac 1{ e^{ix} }= 1 times 1 = 1$$



                        since:



                        $lim_{f(x)to 0}dfrac{e^{f(x)}-1}{f(x)} = 1$






                        share|cite|improve this answer









                        $endgroup$


















                          1












                          $begingroup$

                          We can also use Euler's formula to prove the limit:



                          $$e^{ix} = cos x + isin x$$



                          $$lim_{xto 0}dfrac{sin x}{x} = implies lim_{xto 0} dfrac{e^{ix}- e^{-ix}}{2i x}$$



                          $$= lim_{xto 0} dfrac{e^{2ix}-1}{2ix}timesdfrac 1{ e^{ix} }= 1 times 1 = 1$$



                          since:



                          $lim_{f(x)to 0}dfrac{e^{f(x)}-1}{f(x)} = 1$






                          share|cite|improve this answer









                          $endgroup$
















                            1












                            1








                            1





                            $begingroup$

                            We can also use Euler's formula to prove the limit:



                            $$e^{ix} = cos x + isin x$$



                            $$lim_{xto 0}dfrac{sin x}{x} = implies lim_{xto 0} dfrac{e^{ix}- e^{-ix}}{2i x}$$



                            $$= lim_{xto 0} dfrac{e^{2ix}-1}{2ix}timesdfrac 1{ e^{ix} }= 1 times 1 = 1$$



                            since:



                            $lim_{f(x)to 0}dfrac{e^{f(x)}-1}{f(x)} = 1$






                            share|cite|improve this answer









                            $endgroup$



                            We can also use Euler's formula to prove the limit:



                            $$e^{ix} = cos x + isin x$$



                            $$lim_{xto 0}dfrac{sin x}{x} = implies lim_{xto 0} dfrac{e^{ix}- e^{-ix}}{2i x}$$



                            $$= lim_{xto 0} dfrac{e^{2ix}-1}{2ix}timesdfrac 1{ e^{ix} }= 1 times 1 = 1$$



                            since:



                            $lim_{f(x)to 0}dfrac{e^{f(x)}-1}{f(x)} = 1$







                            share|cite|improve this answer












                            share|cite|improve this answer



                            share|cite|improve this answer










                            answered Jan 13 at 12:24









                            AbcdAbcd

                            3,05331236




                            3,05331236

















                                protected by Community Mar 13 '15 at 19:29



                                Thank you for your interest in this question.
                                Because it has attracted low-quality or spam answers that had to be removed, posting an answer now requires 10 reputation on this site (the association bonus does not count).



                                Would you like to answer one of these unanswered questions instead?



                                Popular posts from this blog

                                Biblatex bibliography style without URLs when DOI exists (in Overleaf with Zotero bibliography)

                                ComboBox Display Member on multiple fields

                                Is it possible to collect Nectar points via Trainline?