Is there an algebraic proof for $sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k} = binom{n+1}{2k+1}, nge2kge0$












3












$begingroup$


$sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k} = binom{n+1}{2k+1}, nge2kge0$



An combinatorial proof of the identity above states as follow:



(1)Number of ways of picking (2k+1) numbers from 1 to (n+1) should be $binom{n+1}{2k+1}$



(2)We pick (2k+1) numbers from 1 to (n+1) with median value (m+1). Then, k numbers must be selected from 1~m, and the other k numbers must be chosen from (m+2)~(n+1). Thus there are $binom{m}{k}binom{n-m}{k}$ ways for picking (2k+1) numbers with median value (m+1). Since $n-kge mge k$, there are total $sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k}$ ways.



Since (1)=(2), the statement is true.
But is it possible to sketch an algebraic proof that doesn't require building combinatorial models?










share|cite|improve this question









$endgroup$












  • $begingroup$
    Consider coefficient of $x^{n}$ in $[x^k(1-x)^{k+1}][x^k(1-x)^{k+1}]$.
    $endgroup$
    – Lord Shark the Unknown
    Jan 1 '18 at 11:22
















3












$begingroup$


$sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k} = binom{n+1}{2k+1}, nge2kge0$



An combinatorial proof of the identity above states as follow:



(1)Number of ways of picking (2k+1) numbers from 1 to (n+1) should be $binom{n+1}{2k+1}$



(2)We pick (2k+1) numbers from 1 to (n+1) with median value (m+1). Then, k numbers must be selected from 1~m, and the other k numbers must be chosen from (m+2)~(n+1). Thus there are $binom{m}{k}binom{n-m}{k}$ ways for picking (2k+1) numbers with median value (m+1). Since $n-kge mge k$, there are total $sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k}$ ways.



Since (1)=(2), the statement is true.
But is it possible to sketch an algebraic proof that doesn't require building combinatorial models?










share|cite|improve this question









$endgroup$












  • $begingroup$
    Consider coefficient of $x^{n}$ in $[x^k(1-x)^{k+1}][x^k(1-x)^{k+1}]$.
    $endgroup$
    – Lord Shark the Unknown
    Jan 1 '18 at 11:22














3












3








3


2



$begingroup$


$sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k} = binom{n+1}{2k+1}, nge2kge0$



An combinatorial proof of the identity above states as follow:



(1)Number of ways of picking (2k+1) numbers from 1 to (n+1) should be $binom{n+1}{2k+1}$



(2)We pick (2k+1) numbers from 1 to (n+1) with median value (m+1). Then, k numbers must be selected from 1~m, and the other k numbers must be chosen from (m+2)~(n+1). Thus there are $binom{m}{k}binom{n-m}{k}$ ways for picking (2k+1) numbers with median value (m+1). Since $n-kge mge k$, there are total $sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k}$ ways.



Since (1)=(2), the statement is true.
But is it possible to sketch an algebraic proof that doesn't require building combinatorial models?










share|cite|improve this question









$endgroup$




$sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k} = binom{n+1}{2k+1}, nge2kge0$



An combinatorial proof of the identity above states as follow:



(1)Number of ways of picking (2k+1) numbers from 1 to (n+1) should be $binom{n+1}{2k+1}$



(2)We pick (2k+1) numbers from 1 to (n+1) with median value (m+1). Then, k numbers must be selected from 1~m, and the other k numbers must be chosen from (m+2)~(n+1). Thus there are $binom{m}{k}binom{n-m}{k}$ ways for picking (2k+1) numbers with median value (m+1). Since $n-kge mge k$, there are total $sum_{m=k}^{n-k} binom{m}{k}binom{n-m}{k}$ ways.



Since (1)=(2), the statement is true.
But is it possible to sketch an algebraic proof that doesn't require building combinatorial models?







combinatorics binomial-coefficients combinatorial-proofs






share|cite|improve this question













share|cite|improve this question











share|cite|improve this question




share|cite|improve this question










asked Jan 1 '18 at 11:19









user3737735user3737735

161




161












  • $begingroup$
    Consider coefficient of $x^{n}$ in $[x^k(1-x)^{k+1}][x^k(1-x)^{k+1}]$.
    $endgroup$
    – Lord Shark the Unknown
    Jan 1 '18 at 11:22


















  • $begingroup$
    Consider coefficient of $x^{n}$ in $[x^k(1-x)^{k+1}][x^k(1-x)^{k+1}]$.
    $endgroup$
    – Lord Shark the Unknown
    Jan 1 '18 at 11:22
















$begingroup$
Consider coefficient of $x^{n}$ in $[x^k(1-x)^{k+1}][x^k(1-x)^{k+1}]$.
$endgroup$
– Lord Shark the Unknown
Jan 1 '18 at 11:22




$begingroup$
Consider coefficient of $x^{n}$ in $[x^k(1-x)^{k+1}][x^k(1-x)^{k+1}]$.
$endgroup$
– Lord Shark the Unknown
Jan 1 '18 at 11:22










2 Answers
2






active

oldest

votes


















1












$begingroup$

Here is an algebraic proof based upon generating functions. It is convenient to use the coefficient of operator $[z^n]$ to denote the coefficient of $z^n$ in a series. This way we can write for instance
begin{align*}
[z^k](1+z)^n=binom{n}{k}
end{align*}




We obtain
begin{align*}
color{blue}{sum_{m=k}^{n-k}}&color{blue}{binom{m}{k}binom{n-m}{k}}\
&=sum_{m=0}^{n-2k}binom{m+k}{m}binom{n-m-k}{k}tag{1}\
&=sum_{m=0}^{n-2k}binom{-k-1}{m}(-1)^mbinom{n-m-k}{k}tag{2}\
&=sum_{m=0}^infty[z^m](1-z)^{-k-1}[u^k](1+u)^{n-m-k}tag{3}\
&=[u^k](1+u)^{n-k}sum_{m=0}^inftyleft(frac{1}{1+u}right)^{m}[z^m](1-z)^{-k-1}tag{4}\
&=[u^k](1+u)^{n-k}left(1-frac{1}{1+u}right)^{-k-1}tag{5}\
&=[u^k](1+u)^{n-k}u^{-k-1}(1+u)^{k+1}tag{6}\
&=[u^{2k+1}](1+u)^{n+1}tag{7}\
&color{blue}{=binom{n+1}{2k+1}}tag{8}
end{align*}
and the claim follows.




Comment:




  • In (1) we shift the index to start with $m=0$ and use the binomial identity $binom{p}{q}=binom{p}{p-q}$.


  • In (2) we use the binomial identity $binom{-p}{q}=binom{p+q-1}{q}(-1)^q$.


  • In (3) we apply the coefficient of operator twice and we set the upper bound of the series to $infty$ without changing anything since we are adding zeros only.


  • In (4) we use the linearity of the coefficient of operator and we apply the rule $[z^{p-q}]A(z)=[z^p]z^qA(z)$.


  • In (5) we apply the substitution rule of the coefficient of operator with $z=frac{1}{1+u}$

    begin{align*}
    A(u)=sum_{m=0}^infty a_m u^m=sum_{m=0}^infty u^m [z^m]A(z)
    end{align*}


  • In (6) we do some simplifications.


  • In (7) we do some more simplifications and apply the same rule as we did in (4).


  • In (8) we select the coefficient of $u^{2k+1}$.







share|cite|improve this answer











$endgroup$





















    0












    $begingroup$

    More generally:




    Theorem 1. For every nonnegative integers $x$, $y$ and $n$, we have
    begin{equation}
    dbinom{n+1}{x+y+1}=sum_{m=0}^{n}dbinom{m}{x}dbinom{n-m}{y}.
    end{equation}




    This equality rewrites as
    begin{equation}
    dbinom{n+1}{x+y+1} = sum_{m=x}^{n-y} dbinom{m}{x}dbinom{n-m}{y} ,
    end{equation}

    because all addends $dbinom{m}{x}dbinom{n-m}{y}$ are zero except for those where $x leq m leq n-y$.



    Your identity is the particular case of the latter equality for $x = k$ and $y = k$.



    Your nice bijective proof of the original identity can easily be generalized to a proof of Theorem 1; just count all ways of picking an $left(x+y+1right)$-element subset of $left{1,2,ldots,n+1right}$ according to the value of the $left(x+1right)$-th-lowest element of the subset. Indeed, for any given $m in left{0,1,ldots,nright}$, picking an $left(x+y+1right)$-element subset $S$ of $left{1,2,ldots,n+1right}$ whose $left(x+1right)$-th-lowest element is $m+1$ is tantamount to picking an $x$-element subset of $left{1,2,ldots,mright}$ (which will contain the $x$ elements of $S$ lower than $m+1$) and picking a $y$-element subset of $left{m+2,m+3,ldots,n+1right}$ (which will contain the $y$ elements of $S$ higher than $m+1$). This gives $dbinom{m}{x}dbinom{n-m}{y}$ for the number of choices.



    As for other proofs:




    • An elementary proof is given in Proposition 3.32 (f) in Darij Grinberg, Notes on the combinatorial fundamentals of algebra, Version of 10 January 2019. The proposition is obtained from Vandermonde convolution by several applications of upper negation.


    • Theorem 1 is also equivalent to the identity proven in Proof of $sum_{m=0}^{n}binom{m}{j}binom{n-m}{k-j}=binom{n+1}{k+1}$ (Another form of the Chu–Vandermonde identity) (just set $j = x$ and $k = x+y$); the proof uses generating functions.







    share|cite|improve this answer











    $endgroup$













      Your Answer





      StackExchange.ifUsing("editor", function () {
      return StackExchange.using("mathjaxEditing", function () {
      StackExchange.MarkdownEditor.creationCallbacks.add(function (editor, postfix) {
      StackExchange.mathjaxEditing.prepareWmdForMathJax(editor, postfix, [["$", "$"], ["\\(","\\)"]]);
      });
      });
      }, "mathjax-editing");

      StackExchange.ready(function() {
      var channelOptions = {
      tags: "".split(" "),
      id: "69"
      };
      initTagRenderer("".split(" "), "".split(" "), channelOptions);

      StackExchange.using("externalEditor", function() {
      // Have to fire editor after snippets, if snippets enabled
      if (StackExchange.settings.snippets.snippetsEnabled) {
      StackExchange.using("snippets", function() {
      createEditor();
      });
      }
      else {
      createEditor();
      }
      });

      function createEditor() {
      StackExchange.prepareEditor({
      heartbeatType: 'answer',
      autoActivateHeartbeat: false,
      convertImagesToLinks: true,
      noModals: true,
      showLowRepImageUploadWarning: true,
      reputationToPostImages: 10,
      bindNavPrevention: true,
      postfix: "",
      imageUploader: {
      brandingHtml: "Powered by u003ca class="icon-imgur-white" href="https://imgur.com/"u003eu003c/au003e",
      contentPolicyHtml: "User contributions licensed under u003ca href="https://creativecommons.org/licenses/by-sa/3.0/"u003ecc by-sa 3.0 with attribution requiredu003c/au003e u003ca href="https://stackoverflow.com/legal/content-policy"u003e(content policy)u003c/au003e",
      allowUrls: true
      },
      noCode: true, onDemand: true,
      discardSelector: ".discard-answer"
      ,immediatelyShowMarkdownHelp:true
      });


      }
      });














      draft saved

      draft discarded


















      StackExchange.ready(
      function () {
      StackExchange.openid.initPostLogin('.new-post-login', 'https%3a%2f%2fmath.stackexchange.com%2fquestions%2f2587436%2fis-there-an-algebraic-proof-for-sum-m-kn-k-binommk-binomn-mk%23new-answer', 'question_page');
      }
      );

      Post as a guest















      Required, but never shown

























      2 Answers
      2






      active

      oldest

      votes








      2 Answers
      2






      active

      oldest

      votes









      active

      oldest

      votes






      active

      oldest

      votes









      1












      $begingroup$

      Here is an algebraic proof based upon generating functions. It is convenient to use the coefficient of operator $[z^n]$ to denote the coefficient of $z^n$ in a series. This way we can write for instance
      begin{align*}
      [z^k](1+z)^n=binom{n}{k}
      end{align*}




      We obtain
      begin{align*}
      color{blue}{sum_{m=k}^{n-k}}&color{blue}{binom{m}{k}binom{n-m}{k}}\
      &=sum_{m=0}^{n-2k}binom{m+k}{m}binom{n-m-k}{k}tag{1}\
      &=sum_{m=0}^{n-2k}binom{-k-1}{m}(-1)^mbinom{n-m-k}{k}tag{2}\
      &=sum_{m=0}^infty[z^m](1-z)^{-k-1}[u^k](1+u)^{n-m-k}tag{3}\
      &=[u^k](1+u)^{n-k}sum_{m=0}^inftyleft(frac{1}{1+u}right)^{m}[z^m](1-z)^{-k-1}tag{4}\
      &=[u^k](1+u)^{n-k}left(1-frac{1}{1+u}right)^{-k-1}tag{5}\
      &=[u^k](1+u)^{n-k}u^{-k-1}(1+u)^{k+1}tag{6}\
      &=[u^{2k+1}](1+u)^{n+1}tag{7}\
      &color{blue}{=binom{n+1}{2k+1}}tag{8}
      end{align*}
      and the claim follows.




      Comment:




      • In (1) we shift the index to start with $m=0$ and use the binomial identity $binom{p}{q}=binom{p}{p-q}$.


      • In (2) we use the binomial identity $binom{-p}{q}=binom{p+q-1}{q}(-1)^q$.


      • In (3) we apply the coefficient of operator twice and we set the upper bound of the series to $infty$ without changing anything since we are adding zeros only.


      • In (4) we use the linearity of the coefficient of operator and we apply the rule $[z^{p-q}]A(z)=[z^p]z^qA(z)$.


      • In (5) we apply the substitution rule of the coefficient of operator with $z=frac{1}{1+u}$

        begin{align*}
        A(u)=sum_{m=0}^infty a_m u^m=sum_{m=0}^infty u^m [z^m]A(z)
        end{align*}


      • In (6) we do some simplifications.


      • In (7) we do some more simplifications and apply the same rule as we did in (4).


      • In (8) we select the coefficient of $u^{2k+1}$.







      share|cite|improve this answer











      $endgroup$


















        1












        $begingroup$

        Here is an algebraic proof based upon generating functions. It is convenient to use the coefficient of operator $[z^n]$ to denote the coefficient of $z^n$ in a series. This way we can write for instance
        begin{align*}
        [z^k](1+z)^n=binom{n}{k}
        end{align*}




        We obtain
        begin{align*}
        color{blue}{sum_{m=k}^{n-k}}&color{blue}{binom{m}{k}binom{n-m}{k}}\
        &=sum_{m=0}^{n-2k}binom{m+k}{m}binom{n-m-k}{k}tag{1}\
        &=sum_{m=0}^{n-2k}binom{-k-1}{m}(-1)^mbinom{n-m-k}{k}tag{2}\
        &=sum_{m=0}^infty[z^m](1-z)^{-k-1}[u^k](1+u)^{n-m-k}tag{3}\
        &=[u^k](1+u)^{n-k}sum_{m=0}^inftyleft(frac{1}{1+u}right)^{m}[z^m](1-z)^{-k-1}tag{4}\
        &=[u^k](1+u)^{n-k}left(1-frac{1}{1+u}right)^{-k-1}tag{5}\
        &=[u^k](1+u)^{n-k}u^{-k-1}(1+u)^{k+1}tag{6}\
        &=[u^{2k+1}](1+u)^{n+1}tag{7}\
        &color{blue}{=binom{n+1}{2k+1}}tag{8}
        end{align*}
        and the claim follows.




        Comment:




        • In (1) we shift the index to start with $m=0$ and use the binomial identity $binom{p}{q}=binom{p}{p-q}$.


        • In (2) we use the binomial identity $binom{-p}{q}=binom{p+q-1}{q}(-1)^q$.


        • In (3) we apply the coefficient of operator twice and we set the upper bound of the series to $infty$ without changing anything since we are adding zeros only.


        • In (4) we use the linearity of the coefficient of operator and we apply the rule $[z^{p-q}]A(z)=[z^p]z^qA(z)$.


        • In (5) we apply the substitution rule of the coefficient of operator with $z=frac{1}{1+u}$

          begin{align*}
          A(u)=sum_{m=0}^infty a_m u^m=sum_{m=0}^infty u^m [z^m]A(z)
          end{align*}


        • In (6) we do some simplifications.


        • In (7) we do some more simplifications and apply the same rule as we did in (4).


        • In (8) we select the coefficient of $u^{2k+1}$.







        share|cite|improve this answer











        $endgroup$
















          1












          1








          1





          $begingroup$

          Here is an algebraic proof based upon generating functions. It is convenient to use the coefficient of operator $[z^n]$ to denote the coefficient of $z^n$ in a series. This way we can write for instance
          begin{align*}
          [z^k](1+z)^n=binom{n}{k}
          end{align*}




          We obtain
          begin{align*}
          color{blue}{sum_{m=k}^{n-k}}&color{blue}{binom{m}{k}binom{n-m}{k}}\
          &=sum_{m=0}^{n-2k}binom{m+k}{m}binom{n-m-k}{k}tag{1}\
          &=sum_{m=0}^{n-2k}binom{-k-1}{m}(-1)^mbinom{n-m-k}{k}tag{2}\
          &=sum_{m=0}^infty[z^m](1-z)^{-k-1}[u^k](1+u)^{n-m-k}tag{3}\
          &=[u^k](1+u)^{n-k}sum_{m=0}^inftyleft(frac{1}{1+u}right)^{m}[z^m](1-z)^{-k-1}tag{4}\
          &=[u^k](1+u)^{n-k}left(1-frac{1}{1+u}right)^{-k-1}tag{5}\
          &=[u^k](1+u)^{n-k}u^{-k-1}(1+u)^{k+1}tag{6}\
          &=[u^{2k+1}](1+u)^{n+1}tag{7}\
          &color{blue}{=binom{n+1}{2k+1}}tag{8}
          end{align*}
          and the claim follows.




          Comment:




          • In (1) we shift the index to start with $m=0$ and use the binomial identity $binom{p}{q}=binom{p}{p-q}$.


          • In (2) we use the binomial identity $binom{-p}{q}=binom{p+q-1}{q}(-1)^q$.


          • In (3) we apply the coefficient of operator twice and we set the upper bound of the series to $infty$ without changing anything since we are adding zeros only.


          • In (4) we use the linearity of the coefficient of operator and we apply the rule $[z^{p-q}]A(z)=[z^p]z^qA(z)$.


          • In (5) we apply the substitution rule of the coefficient of operator with $z=frac{1}{1+u}$

            begin{align*}
            A(u)=sum_{m=0}^infty a_m u^m=sum_{m=0}^infty u^m [z^m]A(z)
            end{align*}


          • In (6) we do some simplifications.


          • In (7) we do some more simplifications and apply the same rule as we did in (4).


          • In (8) we select the coefficient of $u^{2k+1}$.







          share|cite|improve this answer











          $endgroup$



          Here is an algebraic proof based upon generating functions. It is convenient to use the coefficient of operator $[z^n]$ to denote the coefficient of $z^n$ in a series. This way we can write for instance
          begin{align*}
          [z^k](1+z)^n=binom{n}{k}
          end{align*}




          We obtain
          begin{align*}
          color{blue}{sum_{m=k}^{n-k}}&color{blue}{binom{m}{k}binom{n-m}{k}}\
          &=sum_{m=0}^{n-2k}binom{m+k}{m}binom{n-m-k}{k}tag{1}\
          &=sum_{m=0}^{n-2k}binom{-k-1}{m}(-1)^mbinom{n-m-k}{k}tag{2}\
          &=sum_{m=0}^infty[z^m](1-z)^{-k-1}[u^k](1+u)^{n-m-k}tag{3}\
          &=[u^k](1+u)^{n-k}sum_{m=0}^inftyleft(frac{1}{1+u}right)^{m}[z^m](1-z)^{-k-1}tag{4}\
          &=[u^k](1+u)^{n-k}left(1-frac{1}{1+u}right)^{-k-1}tag{5}\
          &=[u^k](1+u)^{n-k}u^{-k-1}(1+u)^{k+1}tag{6}\
          &=[u^{2k+1}](1+u)^{n+1}tag{7}\
          &color{blue}{=binom{n+1}{2k+1}}tag{8}
          end{align*}
          and the claim follows.




          Comment:




          • In (1) we shift the index to start with $m=0$ and use the binomial identity $binom{p}{q}=binom{p}{p-q}$.


          • In (2) we use the binomial identity $binom{-p}{q}=binom{p+q-1}{q}(-1)^q$.


          • In (3) we apply the coefficient of operator twice and we set the upper bound of the series to $infty$ without changing anything since we are adding zeros only.


          • In (4) we use the linearity of the coefficient of operator and we apply the rule $[z^{p-q}]A(z)=[z^p]z^qA(z)$.


          • In (5) we apply the substitution rule of the coefficient of operator with $z=frac{1}{1+u}$

            begin{align*}
            A(u)=sum_{m=0}^infty a_m u^m=sum_{m=0}^infty u^m [z^m]A(z)
            end{align*}


          • In (6) we do some simplifications.


          • In (7) we do some more simplifications and apply the same rule as we did in (4).


          • In (8) we select the coefficient of $u^{2k+1}$.








          share|cite|improve this answer














          share|cite|improve this answer



          share|cite|improve this answer








          edited Jan 1 '18 at 20:02

























          answered Jan 1 '18 at 19:25









          Markus ScheuerMarkus Scheuer

          61.3k456146




          61.3k456146























              0












              $begingroup$

              More generally:




              Theorem 1. For every nonnegative integers $x$, $y$ and $n$, we have
              begin{equation}
              dbinom{n+1}{x+y+1}=sum_{m=0}^{n}dbinom{m}{x}dbinom{n-m}{y}.
              end{equation}




              This equality rewrites as
              begin{equation}
              dbinom{n+1}{x+y+1} = sum_{m=x}^{n-y} dbinom{m}{x}dbinom{n-m}{y} ,
              end{equation}

              because all addends $dbinom{m}{x}dbinom{n-m}{y}$ are zero except for those where $x leq m leq n-y$.



              Your identity is the particular case of the latter equality for $x = k$ and $y = k$.



              Your nice bijective proof of the original identity can easily be generalized to a proof of Theorem 1; just count all ways of picking an $left(x+y+1right)$-element subset of $left{1,2,ldots,n+1right}$ according to the value of the $left(x+1right)$-th-lowest element of the subset. Indeed, for any given $m in left{0,1,ldots,nright}$, picking an $left(x+y+1right)$-element subset $S$ of $left{1,2,ldots,n+1right}$ whose $left(x+1right)$-th-lowest element is $m+1$ is tantamount to picking an $x$-element subset of $left{1,2,ldots,mright}$ (which will contain the $x$ elements of $S$ lower than $m+1$) and picking a $y$-element subset of $left{m+2,m+3,ldots,n+1right}$ (which will contain the $y$ elements of $S$ higher than $m+1$). This gives $dbinom{m}{x}dbinom{n-m}{y}$ for the number of choices.



              As for other proofs:




              • An elementary proof is given in Proposition 3.32 (f) in Darij Grinberg, Notes on the combinatorial fundamentals of algebra, Version of 10 January 2019. The proposition is obtained from Vandermonde convolution by several applications of upper negation.


              • Theorem 1 is also equivalent to the identity proven in Proof of $sum_{m=0}^{n}binom{m}{j}binom{n-m}{k-j}=binom{n+1}{k+1}$ (Another form of the Chu–Vandermonde identity) (just set $j = x$ and $k = x+y$); the proof uses generating functions.







              share|cite|improve this answer











              $endgroup$


















                0












                $begingroup$

                More generally:




                Theorem 1. For every nonnegative integers $x$, $y$ and $n$, we have
                begin{equation}
                dbinom{n+1}{x+y+1}=sum_{m=0}^{n}dbinom{m}{x}dbinom{n-m}{y}.
                end{equation}




                This equality rewrites as
                begin{equation}
                dbinom{n+1}{x+y+1} = sum_{m=x}^{n-y} dbinom{m}{x}dbinom{n-m}{y} ,
                end{equation}

                because all addends $dbinom{m}{x}dbinom{n-m}{y}$ are zero except for those where $x leq m leq n-y$.



                Your identity is the particular case of the latter equality for $x = k$ and $y = k$.



                Your nice bijective proof of the original identity can easily be generalized to a proof of Theorem 1; just count all ways of picking an $left(x+y+1right)$-element subset of $left{1,2,ldots,n+1right}$ according to the value of the $left(x+1right)$-th-lowest element of the subset. Indeed, for any given $m in left{0,1,ldots,nright}$, picking an $left(x+y+1right)$-element subset $S$ of $left{1,2,ldots,n+1right}$ whose $left(x+1right)$-th-lowest element is $m+1$ is tantamount to picking an $x$-element subset of $left{1,2,ldots,mright}$ (which will contain the $x$ elements of $S$ lower than $m+1$) and picking a $y$-element subset of $left{m+2,m+3,ldots,n+1right}$ (which will contain the $y$ elements of $S$ higher than $m+1$). This gives $dbinom{m}{x}dbinom{n-m}{y}$ for the number of choices.



                As for other proofs:




                • An elementary proof is given in Proposition 3.32 (f) in Darij Grinberg, Notes on the combinatorial fundamentals of algebra, Version of 10 January 2019. The proposition is obtained from Vandermonde convolution by several applications of upper negation.


                • Theorem 1 is also equivalent to the identity proven in Proof of $sum_{m=0}^{n}binom{m}{j}binom{n-m}{k-j}=binom{n+1}{k+1}$ (Another form of the Chu–Vandermonde identity) (just set $j = x$ and $k = x+y$); the proof uses generating functions.







                share|cite|improve this answer











                $endgroup$
















                  0












                  0








                  0





                  $begingroup$

                  More generally:




                  Theorem 1. For every nonnegative integers $x$, $y$ and $n$, we have
                  begin{equation}
                  dbinom{n+1}{x+y+1}=sum_{m=0}^{n}dbinom{m}{x}dbinom{n-m}{y}.
                  end{equation}




                  This equality rewrites as
                  begin{equation}
                  dbinom{n+1}{x+y+1} = sum_{m=x}^{n-y} dbinom{m}{x}dbinom{n-m}{y} ,
                  end{equation}

                  because all addends $dbinom{m}{x}dbinom{n-m}{y}$ are zero except for those where $x leq m leq n-y$.



                  Your identity is the particular case of the latter equality for $x = k$ and $y = k$.



                  Your nice bijective proof of the original identity can easily be generalized to a proof of Theorem 1; just count all ways of picking an $left(x+y+1right)$-element subset of $left{1,2,ldots,n+1right}$ according to the value of the $left(x+1right)$-th-lowest element of the subset. Indeed, for any given $m in left{0,1,ldots,nright}$, picking an $left(x+y+1right)$-element subset $S$ of $left{1,2,ldots,n+1right}$ whose $left(x+1right)$-th-lowest element is $m+1$ is tantamount to picking an $x$-element subset of $left{1,2,ldots,mright}$ (which will contain the $x$ elements of $S$ lower than $m+1$) and picking a $y$-element subset of $left{m+2,m+3,ldots,n+1right}$ (which will contain the $y$ elements of $S$ higher than $m+1$). This gives $dbinom{m}{x}dbinom{n-m}{y}$ for the number of choices.



                  As for other proofs:




                  • An elementary proof is given in Proposition 3.32 (f) in Darij Grinberg, Notes on the combinatorial fundamentals of algebra, Version of 10 January 2019. The proposition is obtained from Vandermonde convolution by several applications of upper negation.


                  • Theorem 1 is also equivalent to the identity proven in Proof of $sum_{m=0}^{n}binom{m}{j}binom{n-m}{k-j}=binom{n+1}{k+1}$ (Another form of the Chu–Vandermonde identity) (just set $j = x$ and $k = x+y$); the proof uses generating functions.







                  share|cite|improve this answer











                  $endgroup$



                  More generally:




                  Theorem 1. For every nonnegative integers $x$, $y$ and $n$, we have
                  begin{equation}
                  dbinom{n+1}{x+y+1}=sum_{m=0}^{n}dbinom{m}{x}dbinom{n-m}{y}.
                  end{equation}




                  This equality rewrites as
                  begin{equation}
                  dbinom{n+1}{x+y+1} = sum_{m=x}^{n-y} dbinom{m}{x}dbinom{n-m}{y} ,
                  end{equation}

                  because all addends $dbinom{m}{x}dbinom{n-m}{y}$ are zero except for those where $x leq m leq n-y$.



                  Your identity is the particular case of the latter equality for $x = k$ and $y = k$.



                  Your nice bijective proof of the original identity can easily be generalized to a proof of Theorem 1; just count all ways of picking an $left(x+y+1right)$-element subset of $left{1,2,ldots,n+1right}$ according to the value of the $left(x+1right)$-th-lowest element of the subset. Indeed, for any given $m in left{0,1,ldots,nright}$, picking an $left(x+y+1right)$-element subset $S$ of $left{1,2,ldots,n+1right}$ whose $left(x+1right)$-th-lowest element is $m+1$ is tantamount to picking an $x$-element subset of $left{1,2,ldots,mright}$ (which will contain the $x$ elements of $S$ lower than $m+1$) and picking a $y$-element subset of $left{m+2,m+3,ldots,n+1right}$ (which will contain the $y$ elements of $S$ higher than $m+1$). This gives $dbinom{m}{x}dbinom{n-m}{y}$ for the number of choices.



                  As for other proofs:




                  • An elementary proof is given in Proposition 3.32 (f) in Darij Grinberg, Notes on the combinatorial fundamentals of algebra, Version of 10 January 2019. The proposition is obtained from Vandermonde convolution by several applications of upper negation.


                  • Theorem 1 is also equivalent to the identity proven in Proof of $sum_{m=0}^{n}binom{m}{j}binom{n-m}{k-j}=binom{n+1}{k+1}$ (Another form of the Chu–Vandermonde identity) (just set $j = x$ and $k = x+y$); the proof uses generating functions.








                  share|cite|improve this answer














                  share|cite|improve this answer



                  share|cite|improve this answer








                  edited Jan 10 at 1:46

























                  answered Jan 1 '18 at 12:21









                  darij grinbergdarij grinberg

                  10.8k33062




                  10.8k33062






























                      draft saved

                      draft discarded




















































                      Thanks for contributing an answer to Mathematics Stack Exchange!


                      • Please be sure to answer the question. Provide details and share your research!

                      But avoid



                      • Asking for help, clarification, or responding to other answers.

                      • Making statements based on opinion; back them up with references or personal experience.


                      Use MathJax to format equations. MathJax reference.


                      To learn more, see our tips on writing great answers.




                      draft saved


                      draft discarded














                      StackExchange.ready(
                      function () {
                      StackExchange.openid.initPostLogin('.new-post-login', 'https%3a%2f%2fmath.stackexchange.com%2fquestions%2f2587436%2fis-there-an-algebraic-proof-for-sum-m-kn-k-binommk-binomn-mk%23new-answer', 'question_page');
                      }
                      );

                      Post as a guest















                      Required, but never shown





















































                      Required, but never shown














                      Required, but never shown












                      Required, but never shown







                      Required, but never shown

































                      Required, but never shown














                      Required, but never shown












                      Required, but never shown







                      Required, but never shown







                      Popular posts from this blog

                      Can a sorcerer learn a 5th-level spell early by creating spell slots using the Font of Magic feature?

                      Does disintegrating a polymorphed enemy still kill it after the 2018 errata?

                      A Topological Invariant for $pi_3(U(n))$