Skip to main content
Log in

Cutting planes from extended LP formulations

  • Full Length Paper
  • Series A
  • Published:
Mathematical Programming Submit manuscript

Abstract

Given a mixed-integer set defined by linear inequalities and integrality requirements on some of the variables, we consider extended formulations of its continuous (LP) relaxation and study the effect of adding cutting planes in the extended space. In terms of optimization, extended LP formulations do not lead to better bounds as their projection onto the original space is precisely the original LP relaxation. However, adding cutting planes in the extended space can lead to stronger bounds. In this paper we show that for every 0–1 mixed-integer set with n integer and k continuous variables, there is an extended LP formulation with \(2n+k-1\) variables whose elementary split closure is integral. The proof is constructive but it requires an inner description of the LP relaxation. We then extend this idea to general mixed-integer sets and construct the best extended LP formulation for such sets with respect to lattice-free cuts. We also present computational results on the two-row continuous group relaxation showing the strength of cutting planes derived from extended LP formulations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

References

  1. Andersen, K., Louveaux, Q., Weismantel, R.: Mixed-integer sets from two rows of two adjacent simplex bases. Math. Program. 124, 455–480 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  2. Andersen, K., Louveaux, Q., Weismantel, R., Wolsey, L.A.: Cutting planes from two rows of a simplex tableau. In: Proceedings of the IPCO XII, vol. 4513 of Lecture Notes in Computer Science, pp. 1–15

  3. Balas, E.: Disjunctive programming and a hierarchy of relaxations for discrete optimization problems. SIAM. J. Algebr. Discrete Methods 6(3), 466–486 (1985)

    Article  MATH  MathSciNet  Google Scholar 

  4. Balas, E.: Projection, lifting and extended formulation in integer and combinatorial optimization. Ann. Oper. Res. 140, 125–161 (2005)

    Article  MATH  MathSciNet  Google Scholar 

  5. Balas, E., Pulleyblank, W.R.: The perfectly matchable subgraph polytope of a bipartite graph. Networks 13, 495–518 (1983)

    Article  MATH  MathSciNet  Google Scholar 

  6. Balinski, M.L.: On maximum matching, minimum covering and their connections. In: Kuhn, H.W. (eds.) Proceedings of the Princeton Symposium on Mathematical Programming, pp. 303–312. Princeton University Press, Princeton (1970)

  7. Barahona, F., Mahjoub, A.R.: Compositions of graphs and polyhedra I: balanced induced subgraphs and acyclic subgraphs. SIAM J. Discrete Math. 7, 344–358 (1994)

    Article  MATH  MathSciNet  Google Scholar 

  8. Bodur, M., Dash, S., Günlük, O., Luedtke, J.: Strengthened benders cuts for stochastic integer programs with continuous recourse. IBM Research Report RC25452. IBM Research, Yorktown Heights, NY, USA (2014)

  9. Chvátal, V., Cook, W., Hartmann, M.: On cutting-plane proofs in combinatorial optimization. Linear Algebra Appl. 114(115), 455–499 (1989)

    Article  MATH  MathSciNet  Google Scholar 

  10. Conforti, M., Cornuéjols, G., Zambelli, G.: Extended formulations in combinatorial optimization. Ann. Oper. Res. 204, 97–143 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  11. Cook, W., Kannan, R., Schrijver, A.: Chvátal closures for mixed integer programming problems. Math. Program. 47, 155–174 (1990)

    Article  MATH  Google Scholar 

  12. Cornuéjols, G., Li, Y.: On the rank of mixed 0, 1 polyhedra. Math. Program. 91(2), 391–397 (2002)

    Article  MATH  MathSciNet  Google Scholar 

  13. Dash, S., Günlük, O.: On mixing inequalities: rank, closure, and cutting-plane proofs. SIAM J. Optim. 20(2), 1090–1109 (2009)

    Article  MATH  MathSciNet  Google Scholar 

  14. Del Pia, A.: On the rank of disjunctive cuts. Math. Oper. Res. 37(2), 372–378 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  15. Dey, S.S.: A note on the split rank of intersection cuts. Math. Program. 130(1), 107–124 (2011)

    Article  MATH  MathSciNet  Google Scholar 

  16. Dey, S.S., Lodi, A., Wolsey, L.A., Tramontani, A.: On the practical strength of two-row tableau cuts. INFORMS J. Comput. 26, 222–237 (2014)

  17. Fiorini, S., Massar, S., Pokutta, S., Tiwary, H.R., de Wolf, R.: Linear vs. semidefinite extended formulations: exponential separation and strong lower bounds. In: Proceedings of the STOC, pp. 95–106 (2012)

  18. Günlük, O., Pochet, Y.: Mixing mixed-integer inequalities. Math. Program. 90(3), 429–457 (2001)

    Article  MATH  MathSciNet  Google Scholar 

  19. Hartmann, M.: Cutting Planes and the Complexity of the Integer Hull, Technical Report No. 819 Cornell University Operations Research and Industrial Engineering, Cornell University, Ithaca, NY, USA (1998)

  20. Matoušek, J.: Lectures on discrete geometry. Springer, New York (2002)

    Book  MATH  Google Scholar 

  21. Modaresi, S., Kilinç, M.R., Vielma, J.P.: Split cuts and extended formulations for mixed integer conic quadratic programming. Oper. Res. Lett. 43, 10–15 (2015)

    Article  MathSciNet  Google Scholar 

  22. Nemhauser, G.L., Trotter Jr., L.E.: Properties of vertex packing and independence system polyhedra. Math. Program. 6, 48–61 (1974)

    Article  MATH  MathSciNet  Google Scholar 

  23. Rothvoss, T.: The matching polytope has exponential extension complexity. In: Proceedings of the 46th ACM STOC, pp. 263–272 (2014)

  24. Sherali, H.D., Adams, W.P.: A hierarchy of relaxations between the continuous and convex hull representations for zeroone programming problems. SIAM J. Discrete Math. 3, 411–430 (1990)

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

We would like to thank the anonymous referees for their careful reading of the paper and constructive comments. We also thank Jim Luedtke for useful discussions on this topic.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Oktay Günlük.

Appendix: Proof of Lemma 3.7

Appendix: Proof of Lemma 3.7

Let \(n=n_1+n_2\). As \(P\) is anti-blocking, it is defined by nonredundant system of inequalities \(P=\{x\in {\mathbb {R}}^n:Ax\le b;~x\ge 0\}\) where \(A,b\ge 0\). Assume that the cut \(c^{T} x\le d\) can be derived using the split set \(S(\pi ,{ \gamma })\) and let \(c_k<0\) for some index k.

As \(c^{ T}x\le d\) is valid for \(P^1=\{x\in {\mathbb {R}}^n:Ax\le b, \pi ^{ T} x\le {\gamma }, x\ge 0\}\), there exists multipliers \({ \mu _1,\mu _2\in {\mathbb {R}}_{+}}\) and \(\lambda \in {\mathbb {R}}^n_{+}\), such that

$$\begin{aligned} c=\mu _1 a+\mu _2\pi -\lambda ,\quad \text {and}\quad d \ge \mu _1f+\mu _2{\gamma }, \end{aligned}$$

where \(a^{T}x\le f\) is a non-negative linear combination of the rows of \(Ax\le b\) (and therefore \(a\ge 0\)). Now consider the split set \(S(\pi ^+,{ \gamma })\) where \(\pi ^+\) is identical to \(\pi \) except \(\pi ^+_k=0\). Further let \(\lambda ^+\) be identical to \(\lambda \) except \(\lambda ^+_k=\mu _1 a_k\ge 0\). Now, observe that the inequality \({ {\bar{c}}^T}x\le d\) is valid for \(\bar{P}^1=\{x\in {\mathbb {R}}^n:Ax\le b,~{ (\pi ^+)^T}x\le { \gamma }, x\ge 0\}\) where \({ \bar{c}}=\mu _1 a+\mu _2\pi ^+-\lambda ^+\) and note that \({ \bar{c}}\) is identical to c except \({ \bar{c}}_k=0\).

Using the same argument for \(P^2=\{x\in {\mathbb {R}}^n:Ax\le b, \pi ^{ T} x\ge {\gamma }+1, x\ge 0\}\) establishes that \({ {\bar{c}}^T}x\le d\) is valid for \(P^2\) and therefore is a valid split cut for \(P^{\textit{IP}}\).

Repeating this process for all negative components of c completes the proof. \(\square \)

1.1 Proof of Theorem 3.9

Given a point \((x,y) \in X\), let \(T_{ny} = \{j \in N : y_j > 0\}\), i.e., \(T_{ny}\) stands for the indices of nonzero components of \(y\). Let \(y_{max}\) be the maximum value among all components of \(y\), and let \(T_{max} = \{j \in N : y_j = y_{max}\}\). Let

$$\begin{aligned} T_2= & {} \{ i \in N : (18) \text { for index { i} is tight for } (x,y)\} \text{ and } \\ T_3= & {} \{ i \in N : (19) \text { for index { i} is tight for } (x,y)\}. \end{aligned}$$

If \((x,y)\) is an extreme point of X, it must be the unique solution of a list L of some 2n linearly independent constraints from (17) to (20). From among these 2n constraints, let \(L_1, L_2, L_3, L_{y}\) stand for, respectively, the sets of indices of constraints of types (17)–(20) that are contained in L. By definition, we have

$$\begin{aligned}&L_2 \subseteq T_2,\quad L_3 \subseteq T_3, \quad T_{ny} \ne \emptyset \ { \Longrightarrow } \ T_{max} \subseteq T_{ny}, \end{aligned}$$
(42)
$$\begin{aligned}&L_{y} \subseteq N {\setminus } T_{ny}\quad \text{ and }\quad |L_1| + |L_2| + |L_3| + |L_{y}| = 2n. \end{aligned}$$
(43)

Using these definitions, we first show some properties which we need in the proof.

Lemma 6.1

Let \((x,y)\) be an extreme point of X with \(y\ne 0\). Then

  1. (i)

    \(T_2,T_3 \subseteq T_{max}\) and \(|L_2\cup L_3|+|L_{y}|\le n.\)

  2. (ii)

    \(|T_{ny}| \ge 3\).

  3. (iii)

    \(|L_2 \cap L_3| \le 1\) and \(|L_1| \ge n-1\).

  4. (iv)

    \(T_{max} = T_{ny}\).

Proof

  1. (i)

    Let \(k \in T_{max}\). If the inequality (18) is tight for some index \(i \in N {\setminus } T_{max}\) then \(y_i < y_k\), and inequality (18) for index k is violated, contradicting the fact that \((x,y) \in X\). Therefore \(T_2 \subseteq T_{max}\). The proof of the fact that \(T_3 \subseteq T_{max}\) is very similar. To see the last inequality, notice that by inequality (42) we have \(L_2,L_3\subset T_{max}\subset T_{ny}\), and therefore \(|L_2\cup L_3|+|L_{y}|\le n\).

  2. (ii)

    Let \(y_{i_1} >0\) for some \(i_1 \in N\) and let \(i_2 \in N {\setminus }\{i_1\}\). Adding the constraints in (18) corresponding to the indices \(i_1\) and \(i_2\), we obtain

    $$\begin{aligned} 2\sum _{j \in N} y_j \ge 3(y_{i_1} + y_{i_2}) \Rightarrow 2\sum _{j \in N {\setminus } \{i_1, i_2\}} y_j \ge y_{i_1} + y_{i_2}, \end{aligned}$$

    which implies that at least one of the components of \(y\) different from \(i_1, i_2\) is nonzero. Repeating the same argument with this new component and \(y_{i_1}\) shows that there are at least three such entries.

  3. (iii)

    Suppose that \(|L_2 \cap L_3| \ge 2\). Then there are two indices \(i_1 \ne i_2\) such that the corresponding constraints from (18) and (19) are contained in L. But subtracting \(\sum _{j \in N}y_j = 3y_{i_2}\) from \(\sum _{j \in N} y_j = 3y_{i_1}\), we get \(y_{i_1} - y_{i_2} = 0\). Similarly, subtracting \(2 = \sum _{j \in N} (2x_j - y_j) + 2y_{i_2}\) from \(2 = \sum _{j \in N} (2x_j - y_j) + 2y_{i_1}\), we also get \(y_{i_1} - y_{i_2} = 0\), which means that the above four constraints from L are not linearly independent, a contradiction. Therefore \(|L_2 \cap L_3| \le 1\). But this fact, combined with (i) yields \(|L_2| + |L_3| \le n-|L_{y}| + 1\). Then Eq. (43) implies that \(|L_1| \ge n-1\).

  4. (iv)

    Note that \(T_{max} \subseteq T_{ny}\) by (42). For contradiction, assume \(|T_{max}| \le |T_{ny}| - 1\) and therefore \(0 < y_\ell < y_{max}\) for some index \(\ell \in N\). If \(|L_2| = 0\), then (i) and Eq. (43) imply that

    $$\begin{aligned}&|L_3| \le |T_3| \le |T_{max}| \le |T_{ny}|-1 \Rightarrow |L_2| + |L_3| \le |T_{ny}|-1\\&\quad \qquad \Rightarrow |L_2| + |L_3| + |L_y| \le n-1. \end{aligned}$$

    As \(|L_1| \le n\), this means that L cannot have 2n constraints, a contradiction. Therefore, we can assume that \(|L_2| \ne 0\). Let \(k \in L_2\), i.e., \(\sum _{j \in N} y_j = 3y_k\). Then, (18) implies that \(y_k \ge y_i\) for all \(i \in N\). Thus, \(y_k \in T_{max}\) and \(y_k = y_{max}\). As \(0 < y_\ell < y_k\), there has to be at least another index of \(y\) different from \(\ell \) such that the corresponding component is positive and strictly less than \(y_k\) in order for the constraint \(\sum \nolimits _{j \in N} y_j = 3y_k\) to hold. Then we have \(|T_{max}|\le |T_{ny}|-2\). This fact, combined with (i) and (iii) and Eq. (42) implies that \(|L_2|+|L_3| \le |T_{ny}|-1\). As \(|L_1| \le n\), using Eq. (43), we get \(|L| \le 2n-1\), a contradiction. Therefore \(T_{max} = T_{ny}\).

\(\square \)

Proof of Theorem 3.9

We will first prove that \(\{q^1, \ldots , q^m\} \subseteq X\); as X is convex this will imply that \(Q^n \subseteq X\). Let \(q^j = (x^j, y^j)\) for some \(1 \le j \le m\). If \(x^j = \mathbf{0}\) or if \(x^j\) is a unit vector in \({\mathbb {R}}^n\), then \(y^j = {\varvec{0}}\), and the constraints (17)–(20) are trivially satisfied. Let . The \(i^\mathrm{th}\) component of \(y^j\) is 1 if and only if the \(i^\mathrm{th}\) component of \(x^j\) is 1/2. Therefore (17) holds with equality for all \(i=1, \ldots , n\); in this case the constraints (19) simply become \(y_i \le 1\) for \(i \in N\), which are all satisfied by \(y^j\). Finally, at least three components of \(x^j\) are 1/2, therefore at least three components of \(y^j\) are 1. This, combined with the fact that all components of \(y^j\) are at most 1, implies that all constraints in (18) are satisfied by \(q^j\). Therefore \(Q^n \subseteq X\).

For the reverse inclusion, we will show that each extreme point of X belongs to \(V^{\prime }\). Note that X is a pointed polyhedron as Eqs. (20) and (17) imply that it is contained in \({\mathbb {R}}^n_+\times {\mathbb {R}}^n_+\). Furthermore, inequalities (19) together with inequalities (17) imply that \(2y_i \le 2\) and therefore \(y_i \le 1\) for all \(i \in N\). In addition, when all \(y_i\) are bounded, (19) also implies that all \(x_i\) are bounded as well. Therefore, X is in fact a polytope and is equal to the convex hull of its extreme points.

Let \((x,y) \in {\mathbb {R}}^n \times {\mathbb {R}}^n\) be an extreme point of X. If \(x_i = 1\) for some index \(i \in N\), rewriting (19) as

$$\begin{aligned} 2 \ge \sum _{j \in N {\setminus } \{i\} } (2x_j-y_j) + (2x_i + y_i), \ \forall i \in N, \end{aligned}$$
(44)

it is clear that \(x_i = 1 \Rightarrow y_i = 0\) and \(2x_j = y_j\), for all \(j \in N {\setminus } \{i\}\). Moreover, for any index \(k \in N {\setminus } \{i\}\), Eq. (44) with i replaced by k becomes

$$\begin{aligned} 2 \ge \sum _{j \in N {\setminus } \{k\} } (2x_j-y_j) + (2x_k + y_k) = (2x_i - y_i) + (2x_k + y_k). \end{aligned}$$

Together with \(x_i = 1\) and \(y_i = 0\), the above inequality implies that \(2x_k + y_k = 0 \Rightarrow x_k = y_k = 0\) for \(k \in N {\setminus } \{i\}\). Therefore \((x,y) = (e_i, \mathbf{0})\) which shows that \((x,y) \in V^{\prime } \subseteq Q^n\).

From now on, we assume that \(x_i < 1\), for all \(i \in N\). We first assume \(y= \mathbf{0}\). By Lemma 6.1(iii), we have \(|L_1| \ge n-1\), and therefore at most one of the inequalities of type (17) is not tight for \((x,y)\). Therefore at most one component of x is nonzero. This implies that \((x,y)\) is a convex combination of \((e_k, \mathbf{0})\) and \((\mathbf{0}, \mathbf{0})\) for some \(k \in N\), and therefore \((x,y)\) is equal to one of these two points and is contained in \(V^{\prime }\).

Henceforth assume that \(y\ne \mathbf{0}\). If \(L_3 = \emptyset \), then L does not contain any constraints of type (19), and therefore \(x = \mathbf{0}, y= \mathbf{0}\) is a solution of the constraints in L. As these constraints have a unique solution, we have \((x,y) = (\mathbf{0}, \mathbf{0})\) which contradicts the assumption that \(y\ne \mathbf{0}\). Therefore, \(L_3 \ne \emptyset \).

We will now consider two cases.

Case 1 All constraints of type (17) are tight for \((x,y)\). Let \(k \in L_3\). As \(2x_i - y_i = 0\) for all \(i \in N\), the tight inequality (19) for index k implies that \(2 = 2y_k \Rightarrow y_k = 1\). But this means \(y_{max} = 1\), and, by Lemma 6.1(iv), all nonzero components of \(y\) have value 1. Furthermore, Lemma 6.1(ii) implies that at least three components of \(y\) have value 1. This combined with the fact that \(2x_i = y_i\) for all \(i \in N\) means that and therefore \((x,y) \in V^{\prime }\).

Case 2 At least one constraint of type (17) is not tight for \((x,y)\). Therefore, \(|L_1| \le n-1\) and by Lemma 6.1(iii) \(|L_1| = n-1\). Let \(\{\ell \}= N\setminus L_1\), therefore \(2x_\ell > y_\ell \). It also follows that \(y_{max} < 1\) as \(y_k = 1\) for any \(k \in N\) would imply that \(2x_i - y_i = 0\) for all \(i \in N\), a contradiction.

Further, if \(|L_2 \cap L_3| = 0\), then \(|L_2 \cup L_3|=|L_2| + | L_3|\) and Lemma 6.1(i) would imply that \(|L_2| + | L_3|+|L_y|\le n\). This, in turn, implies that \(|L_1| = n\), a contradiction. Therefore \(|L_2 \cap L_3| \ge 1\), and \(L_2\not =\emptyset \). Consider an index \(k\in L_2\), then \(y_k = y_{max}\) because \(L_2\subseteq T_{max}\). As all nonzero components of \(y\) have value \(y_{max}\) by Lemma 6.1(iv), and inequality (18) is tight for index k, we conclude that \(|T_{ny}|=3\).

Case 2a Let \(\ell \not \in T_{ny} \). Let \(\delta =\min \{y_{max}/3, x_\ell , 1-x_\ell \}>0\). Define \((x^1,y^1)\) by

$$\begin{aligned} x^1_i = \left\{ \begin{array}{ll} x_i + \delta /2, &{}\quad \text{ if } \;i \in T_{ny} \\ x_i - \delta , &{}\quad \text{ if }\; i = \ell \\ x_i, &{}\quad \text{ o.w. }\end{array},\right. \quad y^1_i = \left\{ \begin{array}{ll} y_i + \delta , &{}\quad \text{ if } \;i \in T_{ny} \\ y_i,&{}\quad \text{ o.w. }\end{array}\right. \end{aligned}$$

and \((x^2,y^2) = 2(x,y) - (x^1, y^1)\). Then both \((x^1,y^1)\) and \((x^2,y^2)\) belong to X and \((x,y)\) is a convex combination of these points, contradicting the fact that it is an extreme point of X.

Case 2b Let \(\ell \in T_{ny}\). Let \(\delta =\min \{1-x_\ell +y_{\ell }{/}2, y_{max}{/}3, (2x_\ell - y_\ell ){/}2\}>0\). Define \((x^1,y^1)\) by

$$\begin{aligned} x^1_i = \left\{ \begin{array}{ll} x_i + \delta {/}2, &{}\quad \text{ if } \;i \in T_{ny}{\setminus }\{\ell \} \\ x_i - \delta {/}2, &{}\quad \text{ if }\; i = \ell \\ x_i &{}\quad \text{ o.w. }\end{array},\right. \quad y^1_i = \left\{ \begin{array}{ll} y_i + \delta , &{}\quad \text{ if }\;i \in T_{ny} \\ y_i, &{}\quad \text{ o.w. }\end{array}\right. \end{aligned}$$

and \((x^2,y^2) = 2(x,y) - (x^1, y^1)\). Then both \((x^1,y^1)\) and \((x^2,y^2)\) belong to X and \((x,y)\) is a convex combination of these points, contradicting the fact that it is an extreme point of X.

As the claim holds for both cases, the proof is complete. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bodur, M., Dash, S. & Günlük, O. Cutting planes from extended LP formulations. Math. Program. 161, 159–192 (2017). https://doi.org/10.1007/s10107-016-1005-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10107-016-1005-7

Keywords

Navigation