Skip to main content
Log in

A Convex Programming Approach to Mid-course Trajectory Optimization for Air-to-Ground Missiles

  • Original Paper
  • Published:
International Journal of Aeronautical and Space Sciences Aims and scope Submit manuscript

Abstract

This paper addresses trajectory optimization in the mid-course phase of an air-to-ground missile, when the main objectives are (a) to ensure that the target is locked on in the center of the missile’s field-of-view at a specified flight path angle and (b) to attain maximum possible speed to allow for sufficient maneuverability in the terminal phase. The method presents as a second-order cone program (SOCP) formulation for this trajectory optimization, taking advantage of partial linearization and lossless convexification techniques that effectively handle underlying non-convex characteristics of the problem. A well-established SOCP solver can then be readily used to obtain the optimal solution to this convex program. The proposed approach is validated by (a) proving the losslessness of the convexification, and (b) numerically comparing the results with an existing pseudo-spectral method.

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
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  1. Weiss M, Bucco D (2005) Handover analysis for tactical guided weapons using the adjoint method. In: AIAA guidance, navigation, and control conference and exhibit, San Francisco

  2. Song TL, Shin SJ, Cho H (1999) Impact angle control for planar engagement. IEEE Trans Aerosp Electron Syst 35(4):1439–1444

    Article  Google Scholar 

  3. Ryoo CK, Cho H, Tahk MJ (2005) Optimal guidance laws with terminal impact angle. J Guid Control Dyn 28(4):724–732

    Article  Google Scholar 

  4. Ryoo CK, Cho H, Tahk MJ (2006) Time-to-go weighted optimal guidance with impact angle constraints. IEEE Trans Control Syst Technol 14(3):483–492

    Article  Google Scholar 

  5. Kim BS, Lee JG, Han HS (1998) Biased PNG law for impact angle with angular constraint. IEEE Trans Aerosp Electron Syst 34(1):277–288

    Article  Google Scholar 

  6. Erer KS, Merttopcuoglu O (2012) Indirect impact-angle-control against stationary targets using biased pure proportional navigation. J Guid Control Dyn 35(2):700–703

    Article  Google Scholar 

  7. Park BG, Kim TH, Tahk MJ (2013) Optimal impact angle control guidance law considering the seeker’s field-of-view limits. J Proc Inst Mech Eng Part G J Aerosp Eng 227(8):1347–1364

    Article  Google Scholar 

  8. Kim TH, Park BG, Tahk MJ (2013) Bias-shaping method for biased proportional navigation with terminal-angle constraint. J Guid Control Dyn 36(6):1810–1816

    Article  Google Scholar 

  9. Tekin R, Erer KS (2015) Switched-gain guidance for impact angle control under physical constraints. J Guid Control Dyn 38(2):205–216

    Article  Google Scholar 

  10. Hargraves CR, Paris SW (1987) Direct trajectory optimization using nonlinear programming and collocation. J Guid Control Dyn 10(4):338–342

    Article  Google Scholar 

  11. Betts JT (1998) Survey of numerical methods for trajectory optimization. J Guid Control Dyn 21(2):193–207

    Article  Google Scholar 

  12. Fahroo F, Ross IM (2002) Direct trajectory optimization by a Chebyshev pseudospectral method. J Guid Control Dyn 25(1):160–166

    Article  Google Scholar 

  13. Hull DG (1997) Conversion of optimal control problems into parameter optimization problems. J Guid Control Dyn 20(1):57–60

    Article  Google Scholar 

  14. Lu P (2017) Introducing computational guidance and control. J Guid Control Dyn 40(2):193

    Article  Google Scholar 

  15. Liu X, Lu P, Pan B (2017) Survey of convex optimization for aerospace applications. Astrodynamics 1(1):23–40

    Article  Google Scholar 

  16. Acikmese B, Ploen SR (2007) Convex programming approach to powered descent guidance for mars landing. J Guid Control Dyn 30(5):1353–1366

    Article  Google Scholar 

  17. Blackmore L, Acikmese B, Scharf DP (2010) Minimum landing error powered descent guidance for mars landing using convex optimization. J Guid Control Dyn 33(4):1161–1171

    Article  Google Scholar 

  18. Lu P, Liu X (2013) Autonomous trajectory planning for rendezvous and proximity operations by conic programming. J Guid Control Dyn 36(2):375–389

    Article  Google Scholar 

  19. Liu X, Shen Z, Lu P (2016) Exact convex relaxation for optimal flight of aerodynamically controlled missiles. IEEE Trans Aerosp Electron Syst 52(4):1881–1892

    Article  Google Scholar 

  20. Dueri D, Acikmese B, Scharf DP, Harris MW (2017) Customized real-time interior-point methods for onboard powered-descent guidance. J Guid Control Dyn 40(2):197–212

    Article  Google Scholar 

  21. Liu X, Shen Z, Lu P (2015) Solving the maximum-crossrange problem via successive second-order cone programming with a line search. Aerosp Sci Technol 47:10–20

    Article  Google Scholar 

  22. Liu X, Shen Z, Lu P (2016) Entry trajectory optimization by second-order cone programming. J Guid Control Dyn 39(2):227–241

    Article  Google Scholar 

  23. Szmuk M, Acikmese B, Berning AW Jr (2016) Successive convexification for fuel-optimal powered landing with aerodynamic drag and non-convex constraints. In: AIAA guidance, navigation and control conference, AIAA SciTech Forum, San Diego, California

  24. Mao Y, Szmuk M, Acikmese B (2016) Successive convexification of non-convex optimal control problems and its convergence properties. In: IEEE 55th conference on decision and control, Las Vegas

  25. Szmuk M, Eren U, Berning AW Jr (2017) Successive convexification for 6-DoF powered descent landing guidance. In: AIAA guidance, navigation and control conference, AIAA SciTech Forum, Grapevine

  26. Mao Y, Dueri D, Szmuk M, Acikmese B (2017) Successive convexification of non-convex optimal control problems with state constraints. In: 20th IFAC world congress, Toulouse

  27. Acikmese B, Blackmore L (2011) Lossless convexification of a class of optimal control problems with non-convex control constraints. Automatica 47(2):341–347

    Article  MathSciNet  Google Scholar 

  28. Harris MW, Acikmese B (2013) Lossless convexification for a class of optimal control problems with linear state constraints. In: 52nd IEEE conference on decision and control, Florence

  29. Acikmese B, Carson JM III, Blackmore L (2013) Lossless convexification of nonconvex control bound and pointing constraints of the soft landing optimal control problem. IEEE Trans Control Syst Technol 21(6):2104–2113

    Article  Google Scholar 

  30. Harris MW, Acikmese B (2013) Lossless convexification for a class of optimal control problems with quadratic state constraints. In: American control conference, Washington, DC

  31. Boyd S, Vandenberghe L (2004) Convex optimization. Cambridge University Press, Cambridge

    Book  Google Scholar 

  32. Hartl RF, Sethi SP, Vickson RG (1995) A survey of the maximum principles for optimal control problems with state constraints. SIAM Rev 37(2):181–218

    Article  MathSciNet  Google Scholar 

  33. Sagliano M (2018) Pseudospectral convex optimization for powered descent and landing. J Guid Control Dyn 41(2):320–334

    Article  Google Scholar 

  34. Sagliano M (2019) Generalized hp pseudospectral convex programming for powered descent and landing. J Guid Control Dyn 42(7):1562–1570

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Han-Lim Choi.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix: Proof of Lossless Convexification

Appendix: Proof of Lossless Convexification

The proof of lossless convexification is carried out using the proof by contradiction. Let us assume that there exists a certain interval \( \left[ {x_{1} ,x_{2} } \right] \subset \left[ {x_{0} ,x_{\text{f}} } \right] \) that satisfies \( u_{1}^{2} < u_{2} \). There should exist a constant \( p_{0} \le 0 \) for an optimal solution \( \left( {q^{*} ,u^{*} } \right) \) satisfying the conditions in Eqs. (19)–(33) in Theorem 1.

  1. (a)

    In the interval \( \left[ {x_{1} ,x_{2} } \right] \subset \left[ {x_{0} ,x_{\text{f}} } \right] \), the following relationship is obtained:

    $$ u_{1}^{2} < u_{2} ,\;u_{2} > 0. $$
    (55)

    The values of \( \lambda_{1} ,\lambda_{2} \) can be determined from complementary slackness conditions (42) and (43):

    $$ \lambda_{0} = \lambda_{1} = 0. $$
    (56)
  2. (b)

    From the stationary condition (40), the following equation is derived:

    $$ \begin{aligned} \frac{\partial L}{{\partial u_{1} }} = p_{\gamma } b_{31} = p_{\gamma } \frac{{\left( {T + qS_{\text{ref}} C_{{{\text{L}}_{\alpha } }} } \right)}}{{mV^{2} \cos \gamma }} = 0, \hfill \\ \to p_{\gamma } = 0\;\quad \because T \ge 0,\;q > 0,\;S_{\text{ref}} > 0,\;C_{{{\text{L}}_{\alpha } }} > 0. \hfill \\ \end{aligned} $$
    (57)
  3. (c)

    From the stationary condition (41), the following equation is derived:

    $$ \begin{aligned} \frac{\partial L}{{\partial u_{2} }} = p_{V} b_{22} + \lambda_{2} = 0, \hfill \\ \to p_{V} b_{22} = - \lambda_{2} \ge 0. \hfill \\ \end{aligned} $$
    (58)

    Hamiltonian in Eq. (34) can be re-formulated as follows:

    $$ H = p_{y} \left( {a_{13} \gamma + c_{y} } \right) + p_{V} \left( {a_{22} V + a_{23} \gamma + c_{V} } \right) + p_{\gamma } \left( {a_{32} V + c_{\gamma } } \right) + p_{\gamma } b_{31} u_{1} + p_{V} b_{22} u_{2} . $$
    (59)

    Based on the pointwise maximum condition (20), \( u_{2} \) is determined according to the switching function \( p_{V} b_{22} \):

    $$ {\text{if}}\; p_{V} b_{22} = 0 \to u_{2} \in \left[ {0,u_{2\rm{max} } } \right], $$
    (60)
    $$ {\text{if }} p_{V} b_{22} > 0 \to u_{2} = 0. $$
    (61)

    Since \( u_{2} \) should be positive as in Eq. (55), \( p_{\text{V}} \) is determined as follows:

    $$ p_{V} = 0\quad \;\because b_{22} < 0. $$
    (62)
  4. (d)

    Since \( p_{\gamma } = p_{V} = 0 \) from (57) and (62), the differential equations (37)–(39) are described as follows:

    $$ p_{y}^{'} = - \frac{\partial L}{\partial y} = \nu_{y1} - \nu_{y2} , $$
    (63)
    $$ p_{V}^{'} = - \frac{\partial L}{\partial V} = - \left( {p_{V} a_{22} + p_{\gamma } a_{32} } \right) = 0, $$
    (64)
    $$ p_{\gamma }^{'} = - \frac{\partial L}{\partial \gamma } = - \left( {p_{y} a_{13} + p_{V} a_{23} } \right) = - p_{y} a_{13} = 0. $$
    (65)

    Since the downrange is monotonically increasing and \( a_{13} \) cannot be 0, then

    $$ p_{y} = 0. $$
    (66)

    Therefore, we obtain \( \nu_{y1} = \nu_{y2} \) from Eq. (63). If the altitude is larger than 0, then \( \nu_{y1} = 0 \) from Eq. (45) and if the altitude is less than \( y_{\hbox{max} } \), then \( \nu_{y2} = 0 \) from Eq. (46). Thus, the multipliers \( \nu_{y1} \) and \( \nu_{y2} \) should be 0 regardless of the altitude condition:

    $$ \nu_{y1} = \nu_{y2} = 0. $$
    (67)
  5. (e)

    From Assumption 2, Eqs. (47), (48), (49) give

    $$ \zeta_{y1} = \zeta_{y2} = p_{y} \left( {x_{\text{f}} } \right) = 0. $$
    (68)
  6. (f)

    The altitude constraint is described as follows:

    $$ S\left( q \right) = y - y_{\rm{max} } \le 0. $$

    The constraint \( S\left( q \right) \) is second-order constraint with respect to the control and then the derivatives of the constraint is calculated

    $$ \begin{aligned} \frac{S\left( q \right)}{\partial q} = \left( {\begin{array}{*{20}c} 1 \\ 0 \\ 0 \\ \end{array} } \right), \;\frac{{S^{\prime}\left( q \right)}}{\partial q} = \left( {\begin{array}{*{20}c} 0 \\ 0 \\ {\sec^{2} \gamma \left( {1 + 2\gamma \tan \gamma } \right)} \\ \end{array} } \right) , \hfill \\ {\text{where }}S^{\prime}\left( q \right) = y^{\prime} = \gamma \sec^{2} \gamma . \hfill \\ \end{aligned} $$
    (69)

    The jump condition (29) for the costate is

    $$ \left( {\begin{array}{*{20}c} {p_{y} \left( {\tau_{1}^{ - } } \right)} \\ {p_{V} \left( {\tau_{1}^{ - } } \right)} \\ {p_{\gamma } \left( {\tau_{1}^{ - } } \right)} \\ \end{array} } \right) = \left( {\begin{array}{*{20}c} {p_{y} \left( {\tau_{1}^{ + } } \right)} \\ {p_{V} \left( {\tau_{1}^{ + } } \right)} \\ {p_{\gamma } \left( {\tau_{1}^{ + } } \right)} \\ \end{array} } \right) + \eta_{1} \left( {\tau_{1} } \right)\left( {\begin{array}{*{20}c} 1 \\ 0 \\ 0 \\ \end{array} } \right) + \eta_{2} \left( {\tau_{1} } \right)\left( {\begin{array}{*{20}c} 0 \\ 0 \\ {\sec^{2} \gamma \left( {1 + 2\gamma \tan \gamma } \right)} \\ \end{array} } \right). $$
    (70)

    If the altitude constraint is active, the flight path angle should be 0. Therefore,

    $$ \left( {\begin{array}{*{20}c} {p_{y} \left( {\tau_{1}^{ - } } \right)} \\ {p_{V} \left( {\tau_{1}^{ - } } \right)} \\ {p_{\gamma } \left( {\tau_{1}^{ - } } \right)} \\ \end{array} } \right) = \left( {\begin{array}{*{20}c} {p_{y} \left( {\tau_{1}^{ + } } \right) + \eta_{1} \left( {\tau_{1} } \right)} \\ {p_{V} \left( {\tau_{1}^{ + } } \right)} \\ {p_{\gamma } \left( {\tau_{1}^{ + } } \right) + \eta_{2} \left( {\tau_{1} } \right)} \\ \end{array} } \right). $$
    (71)

    Since \( p_{\gamma } ,p_{V} ,p_{y} \) are 0 from (57), (62), (66), then \( \eta_{1} \left( {\tau_{1} } \right),\eta_{2} \left( {\tau_{1} } \right) \) should be 0:

    $$ \eta_{1} \left( {\tau_{1} } \right) = \eta_{2} \left( {\tau_{1} } \right) = 0. $$
    (72)
  7. (g)

    Because the Hamiltonian, endpoint, and state constraint functions are autonomous, then

    $$ H\left( t \right) = 0 \forall t. $$
    (73)

From Eqs. (51) and (68), \( p_{\gamma } \left( {x_{\text{f}} } \right) \) and \( p_{y} \left( {x_{\text{f}} } \right) \) are 0 and then:

$$ p_{V} \left( {x_{\text{f}} } \right) = - p_{0} = 0. $$
(74)

From the above results, we conclude

$$ \left( {p_{0} ,p\left( t \right),\lambda \left( t \right),\nu \left( t \right),\xi ,\zeta ,\eta_{1} ,\eta_{2} , \ldots } \right) = 0. $$
(75)

This contradicts the non-triviality condition (19). Therefore, the following equation holds:

$$ u_{1}^{2} \left( x \right) = u_{2} \left( x \right)\;{\text{a}} . {\text{e}} . x \in \left[ {x_{0} ,x_{\text{f}} } \right]. $$
(76)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kwon, HH., Choi, HL. A Convex Programming Approach to Mid-course Trajectory Optimization for Air-to-Ground Missiles. Int. J. Aeronaut. Space Sci. 21, 479–492 (2020). https://doi.org/10.1007/s42405-019-00219-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s42405-019-00219-9

Keywords

Navigation