Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article

On q-steepest descent method for unconstrained multiobjective optimization problems

  • Received: 28 April 2020 Accepted: 22 June 2020 Published: 28 June 2020
  • MSC : 90C29, 05A30, 34M60, 58E17, 90C53, 11Y16

  • The q-gradient is the generalization of the gradient based on the q-derivative. The q-version of the steepest descent method for unconstrained multiobjective optimization problems is constructed and recovered to the classical one as q equals 1. In this method, the search process moves step by step from global at the beginning to particularly neighborhood at last. This method does not depend upon a starting point. The proposed algorithm for finding critical points is verified in the numerical examples.

    Citation: Kin Keung Lai, Shashi Kant Mishra, Geetanjali Panda, Md Abu Talhamainuddin Ansary, Bhagwat Ram. On q-steepest descent method for unconstrained multiobjective optimization problems[J]. AIMS Mathematics, 2020, 5(6): 5521-5540. doi: 10.3934/math.2020354

    Related Papers:

    [1] Shexiang Hai, Liang He . The steepest descent method for fuzzy optimization problems under granular differentiability. AIMS Mathematics, 2025, 10(4): 10163-10186. doi: 10.3934/math.2025463
    [2] Chunhong Li, Dandan Yang, Chuanzhi Bai . Some Opial type inequalities in (p, q)-calculus. AIMS Mathematics, 2020, 5(6): 5893-5902. doi: 10.3934/math.2020377
    [3] Zhongbin Zheng, Jinwu Fang, Wentao Cheng, Zhidong Guo, Xiaoling Zhou . Approximation properties of modified (p, q)-Szász-Mirakyan-Kantorovich operators. AIMS Mathematics, 2020, 5(5): 4959-4973. doi: 10.3934/math.2020317
    [4] Jia-Tong Li . A redistributed cutting plane bundle-type algorithm for multiobjective nonsmooth optimization. AIMS Mathematics, 2022, 7(7): 12827-12841. doi: 10.3934/math.2022710
    [5] Xiaoli Zhang, Shahid Khan, Saqib Hussain, Huo Tang, Zahid Shareef . New subclass of q-starlike functions associated with generalized conic domain. AIMS Mathematics, 2020, 5(5): 4830-4848. doi: 10.3934/math.2020308
    [6] Waseem Ahmad Khan, Kottakkaran Sooppy Nisar, Dumitru Baleanu . A note on (p, q)-analogue type of Fubini numbers and polynomials. AIMS Mathematics, 2020, 5(3): 2743-2757. doi: 10.3934/math.2020177
    [7] Chunyan Luo, Tingsong Du, Muhammad Uzair Awan, Yao Zhang . Estimation-type results with respect to the parameterized (p, q)-integral inequalities. AIMS Mathematics, 2020, 5(1): 568-586. doi: 10.3934/math.2020038
    [8] Sina Etemad, Sotiris K. Ntouyas . Application of the fixed point theorems on the existence of solutions for q-fractional boundary value problems. AIMS Mathematics, 2019, 4(3): 997-1018. doi: 10.3934/math.2019.3.997
    [9] Sani Aji, Poom Kumam, Aliyu Muhammed Awwal, Mahmoud Muhammad Yahaya, Kanokwan Sitthithakerngkiet . An efficient DY-type spectral conjugate gradient method for system of nonlinear monotone equations with application in signal recovery. AIMS Mathematics, 2021, 6(8): 8078-8106. doi: 10.3934/math.2021469
    [10] İbrahim Aktaş . On some geometric properties and Hardy class of q-Bessel functions. AIMS Mathematics, 2020, 5(4): 3156-3168. doi: 10.3934/math.2020203
  • The q-gradient is the generalization of the gradient based on the q-derivative. The q-version of the steepest descent method for unconstrained multiobjective optimization problems is constructed and recovered to the classical one as q equals 1. In this method, the search process moves step by step from global at the beginning to particularly neighborhood at last. This method does not depend upon a starting point. The proposed algorithm for finding critical points is verified in the numerical examples.


    In multiobjective problems, several objective functions have to be minimized or maximized simultaneously. There are conflicts between the objective functions. It is not possible to get a single point that minimizes all objective functions simultaneously. Therefore, the concept of optimality is replaced by the concept of Pareto optimality, a measure of efficiency in multiobjective problems [1]. A point is called Pareto-optimal or efficient, if there does not exist a different point with same or smaller objective function values such that there is a decrease in at least one objective function value. Applications of such optimization problems can be found in space exploration [2,3], engineering [4], truss optimization [5,6], design [7,8], environmental analysis [9,10], statistics [11], management science [12,13,14], economic sciences [15], etc.

    There are many solution strategies for finding a Pareto optimal point of unconstrained multiobjective optimization problems (UP). One popular approach is to reformulate (UP) as a scalar optimization problem whose objective function is a weighted linear combination of all objective functions. However, this approach may yield unbounded scalar problems [20], if the weighted values are not properly chosen. Another popular approach is the evolutionary algorithm which searches for a Pareto optimum in a set of candidate solutions with some genetic operator. No convergence proofs are known in this case, and empirical results show that convergence is quite slow [16]. Another general solution strategy for multiobjective optimization problems is ϵ-constrained method [17]. A drawback of this method is that ϵ may be selected so that the feasible region becomes empty. Other parameter-free multiobjective optimization techniques use the ordering of the different criteria, that is, an ordering of the objective functions based on priority. In this case, the ordering has to be prespecified. Moreover, the corresponding optimization process is usually augmented by an interactive procedure [18]. Another kind of approach is descent methods, which extend the traditional descent methods for scalar optimization to solve (UP), such as the gradient descent method [19], Newton's method [10,20], etc. A remarkable property of this method is that it is a parameter free approach. This is quite different from the weighted approach for (UP). With this method, there is no need to analyze the prior information including the relationships and conflicting between the objectives. Note that such information is very important for choosing the weights for the weighted method. Specifically, an example is shown in [20] that the weighted method fails for a large range of weights which leads to unbounded weighted objective, but a descent method works well with any starting point.

    The steepest descent method is one of the oldest and simplest descent methods to minimize the real-valued functions proposed by Baron Augustin-Louis Cauchy (1789–1857) [21] in 1847. Cauchy developed the gradient method to solve nonlinear equations. The idea of this method is that the continuous function value always decreases if the negative descent direction is considered. Almost, all optimization books [26] discuss this method to learn advanced optimization algorithms. However, this method is used to solve system of ridge regression or regularized least squares [22], also see [23,24,25]. The steepest descent method for single-objective problems was extended for unconstrained multiobjective optimization problems (UP) [19,27]. Further, the method is presented to find critical points of (UP) using maps from a Euclidean space to a Banach space [28]. Recently, the descent method for (UP) decreases the gradient at the rate of 1k regardless of the starting point. This rate of convergence is derived for obtaining some point satisfying weak Pareto-optimality [29].

    On the other hand, quantum calculus (briefly, q-calculus) is dedicated to the study of calculus regardless of the limits. In fact, by introducing the parameter q, we consider q-analogue of calculus concepts such that they can be recaptured for the case when q1. Euler was the founder of this branch of mathematics, by using the parameter q in Newton's work of infinite series. The q-calculus is one of the research interests in the field of mathematics and physics for the last few decades. The q-analogue of ordinary derivative was first proposed by F.H. Jackson[30,31]. Its wide applications can be seen in several areas such as operator theory [32], mean value theorems of q-calculus [33], q-Taylor formula and its remainder [34,35], fractional integrals and derivatives [36], integral inequalities [37], variational calculus [38], transform calculus [39], sampling theory [40], etc. Recently, a qFunctions Mathematica package is developed for q-series, and partition theory applications [41].

    A q-version of the classical steepest descent method has been developed to solve single objective unconstrained optimization problems [42,43,44,45]. In this method, the quantum value q acts as a dilation parameter that controls the balance between global and local search direction in the context of q-gradient. The q-LMS (Least Mean Square) algorithm is proposed which utilizes the q-gradient to compute the secant of the cost function instead of the tangent [46]. The algorithm takes larger steps towards the optimum solution and achieves a higher convergence rate. A new class of stochastic gradient algorithm based on q-calculus is developed to enhance the q-LMS algorithm. The proposed approach utilizes a parameterless concept of error correlation energy, and normalization of signal for high convergence, stability, and low steady-state error. The q-Taylor formula for the functions of several variables and mean value theorems in q-calculus are utilized for solving the systems of equations [48]. In multiobjective optimization, Newton's method using q-calculus [49] is extended for unconstrained multiobjective optimization problems [50]. Following this trend, the Quasi-Newton's method [51] is further extended for solving (UP) [52] where components of the Hessian matrix are computed using q-calculus.

    The goal of this paper is to present the q-steepest descent method, a generalization of the steepest descent method for (UP) [10,27]. A subproblem is formulated using the quadratic approximation of all functions at every iteration, and a feasible descent direction due to q-gradient is obtained as a solution of this subproblem [53]. The ordering of the objective functions is avoided. Instead of the classical gradient, the q-gradient of the objective function is used in the proposed algorithm to find critical points and (weak) efficient points and its theoretical results are validated by giving convergence proofs. The advantage of using q-gradient is that it allows the steepest descent direction to perform in a more diverse set of directions which makes it possible to escape local critical points. Several examples are presented where q-analog is implemented to shift the search process from global in the beginning to almost local search in the end. To evaluate the performance of the proposed method, we compare it with [28] and the weighted sum method. In general, the involvement of q-calculus effectively reduces the number of iterations to reach the critical point or (weak) efficient points. Thus, the proposed method is more promising to solve convex problems, non-convex problems, and multimodal multiobjective optimization problems.

    The paper is arranged as follows: Section 2 recalls concepts related to q-calculus and multiobjective optimization. In Section 3, the subproblem for the descent direction using qderivative is presented. The proposed algorithm and convergence analysis with numerical examples are provided in Section 4. Comparison with the existing method is done in Section 5. The conclusion is given in the last section.

    Denote R as the set of real numbers, Rn+:={xRn|xi0,i=1,,m}, and as Rn++:={xRn|xi>0,i=1,,m}. A continuous function on any interval not containing 0, is called the continuous q-differentiable function. Let the q-integer [n] be defined as [n]=1qn1q, for nN, and the derivative of xn with respect to x be given as [n]xn1, then the q-derivative [30] is given as:

    Dqf(x)={f(qx)f(x)qxx,x0,q1,df(x)dx,otherwise. (2.1)

    The q-derivative reduces to an ordinary derivative as q equals 1. The first-order partial q-derivative [42] of function f with respect to the variable xi, where i=1,,n is:

    Dqi,xif(x)={f(x1,,qixi,,xn)f(x1,,xi,,xn)qixixi,xi0,qi1,f(x)xi,otherwise. (2.2)

    The q-gradient is the vector of n first-order partial q-derivatives of f is:

    qf(x)T=[Dq1,x1f(x)Dqi,xif(x)Dqn,xnf(x)], (2.3)

    where the parameter q is a vector q=(q1,,qi,,qn)Rn. Consider the following unconstrained multiobjective optimization problem (UP):

    minimize(f1(x),fm(x)),xRn, (2.4)

    where fi:RnR,  i=1,,m are real-valued continuously q-differentiable function. For two vectors x,yRn, we denote x=y if and only if xi=yi i=1,,m, xy if and only if xiyi i=1,,m, xy if and only if xiyi and xy, i=1,,m, x>y if and only if xi>yi  i=1,,m. A feasible solution xXRn is called a locally (weakly) efficient solution [20] if there is a neighborhood WX of x is an (weakly) efficient solution, then fi(x) is called a locally (weakly) non-dominated point. Let XRn be a convex set, and fi:XRm is called Rm-convex if fi,  i=1,,m is componentwise-convex [17]. If the domain of the multiobjective optimization problem is a convex set and the objective functions are convex component-wise then critical point will be the weak efficient point, and if the objective functions are strictly convex component-wise then critical point will be efficient point. The relationship between critical point and efficient point is discussed in [20]. The classical gradient is obtained in the limit of qi1, for all i=1,,m. The following example is from [49].

    Example 1. Consider a function f:R2R, where f(x,y)=xy2+x4. Then, the q-gradient [52] is computed as:

    qf(x)T=[y2+(1+q)(1+q2)x3(1+q)xy].

    The transpose of Jacobian matrix of q-partial derivative denoted as JqF(x) for the vector valued objective functions F is:

    JqF(x)=[Dq1,x1f1(x)Dq2,x2f1(x)Dqn,xnf1(x)Dq1,x1f2(x)Dq2,x2f2(x)Dqn,xnf2(x)Dq1,x1fm(x)Dq2,x2fm(x)Dqn,xnfm(x)]. (3.1)

    Note that a variant q(0,1) has been considered in the above matrix. The range of the linear mapping given by the Jacobian matrix JqF(x) is range(JqF(x)). A necessary condition for a (local) weak Pareto minimizer xRn is given as:

    range(JqF(x)T)(R++)m=ϕ. (3.2)

    Using Gordan's theorem of alternative one can show that x is a critical point if there exists λRm+, λ0 such that

    mi=0λiqfi(x)=0 (3.3)

    holds (see Lemma 3.1 of [55]). In the literature (3.2) or (3.3) is often considered as a necessary condition of weak efficiency. The points satisfying (3.2) are called Pareto critical points. If a point xRn is not Pareto critical then there exists a descent direction dqRn for F satisfying

    qF(x)dq(R++)m. (3.4)

    If dqRn is not a descent direction, then

    qF(x)dq(R++)m. (3.5)

    The descent gradient method for (UP) takes the schema xk+1=xk+αkdkq, where αk(0,1] is the step-length computed by Armijo-Wolfe line search with backtracking. For single-objective case, the negative q-gradient qf(x) is the steepest descent direction, and it is the solution of quadratic programming whose objective function is an approximation to f(x+dq)f(x), that is,

    min

    For (UP) case, the subproblem introduced in [27], is an approximation to the least reduction of all objective functions, which is given as:

    \begin{align} \min\limits_{d_{q}\in\mathbb{R}^{n}} \max\limits_{i = 1,\dots,m} \nabla_{q} f_{i}(x^{k})^{T}d_{q}+\frac{1}{1+q}\lVert d_{q} \rVert^{2}, \end{align} (3.6)

    where q\in(0, 1) , and the search direction d_{q}^{k} is the solution of (3.6). The min-max subproblem (3.6) is depicted in the form of q -differentiable quadratic optimization problem ( q -DQOP) with linear inequality constraints as:

    \begin{align*} \arg \min\limits_{d_{q}\in\mathbb{R}^{n},t\in\mathbb{R}}& \quad t +\frac{1}{(1+q)}\lVert d_{q} \rVert^{2},\quad q\in(0,1), \nonumber\\ \text{subject to}& \quad \nabla_{q} f_{i}(x^{k})^{T}d_{q}\le t ,\quad i = 1,\dots,m. \end{align*}

    Note that the term \frac{1}{(1+q)}\lVert d_{q}\rVert^{2} for q\in (0, 1) in the objective function ensures that the problem is bounded, and (d_{q}^{k}, t^{k}) are the optimial solution of the above problem ( q -DQOP). The above sub-problem is a convex programming problem and satisfies Slater's constraint qualification since for t = 1 , and d_{q} = 0 , the above inequalities become strict inequalities. The Karush-Kuhn-Tucker (KKT) conditions [54] for ( q -DQOP) at x^{k} are:

    \begin{align} d_{q}^{k}& = -\sum\limits_{i = 1}^{m}\lambda_{i}^{k}\nabla_{q}f_{i}(x^{k}), \end{align} (3.7)
    \begin{align} \nabla_{q}f_{i}(x^{k})^{T}d_{q}^{k}&\le t^{k},\;\;i = 1,\dots,m, \end{align} (3.8)
    \begin{align} \sum\limits_{i = 1}^{m} \lambda_{i}^{k}& = 1,\;\;\lambda_{i}^{k}\ge 0, \end{align} (3.9)
    \begin{align} \lambda_{i}^{k}(\nabla_{q}f_{i}(x^{k})^{T }d_{q}-t^{k})& = 0,\;\;i = 1,\dots,m, \end{align} (3.10)

    Note that \lambda_{i}^{k}\ge 0, where i = 1, \dots m, are the Lagrange multipliers associated with linear inequality constraints of (3.7). Taking summation of (3.10) over i from 1 to m , and from (3.7), we get

    \begin{align} t^{k} = \sum\limits_{i = 1}^{m}\lambda_{i}^{k}\nabla_{q} f_{i}(x^{k})^{T}d_{q}^{k} = -\lVert d_{q}^{k} \rVert^{2}. \end{align} (3.11)

    The solution of the ( q -DQOP) is related to Pareto critical as stated in [27].

    Lemma 1. [29] Let (d_{q}^{k}, t^{k}) be the solution of ( q -DQOP)

    1. If x^{k} is Pareto critical, then d_{q}^{k} = 0\in\mathbb{R}^{n} , and t^{k} = 0.

    2. If x^{k} is not Pareto critical, then

    \begin{align} t^{k}\le -\frac{1}{(1+q)}\lVert d_{q}^{k} \rVert^{2} & \lt 0;\;\;q\in(0,1), \end{align} (3.12)
    \begin{align} \nabla_{q} f_{i}(x^{k})^{T}d_{q}^{k} &\le t^{k}, \;\; i = 1,\dots,m. \end{align} (3.13)

    Proof. If x^{k} is Pareto critical, then there is no d_{q}^{k} such that \nabla_{q}f_{i}(x^{k})^{T}d_{q}^{k} < 0, \forall\; i = 1, \dots, m, otherwise Part 2 would not be satisfied, and for any fix \bar{i} , we have t^{k}\ge \nabla_{q}f_{\bar{i}}(x^{k})^{T}d_{q}^{k}. Part 1 of this lemma holds when (t^{k}, d^{k}_{q}) = (0, 0) , is a feasible point of ( q -DQOP). The constraint of ( q -DQOP) is equivalent to (3.13). Since d_{q} = 0 , and t = 0 , then we have (d_{q}^{k}, t^{t})\le (0, 0) , thus (3.12) holds. If x^{k} is not a Pareto critical, then there exists a descent search direction d_{q}^{k} such that \nabla f_{i}(x^{k})^{T}d^{k} < 0, \; \forall\; i = 1, \dots, m to get t^{k} < 0.

    Suppose (t^k, d_{q}^k) is the solution of the sub-problem and d_{q}^k\neq 0 , following the arguments in [50,55] we select a suitable step length. Suppose \theta_i^k is the angle between \nabla f_i^k and d_q^k . If \cos^2(\theta_i^k)\geq \delta, where \delta > 0 holds for all i then we select \alpha_k satisfying

    \begin{align} f_{i}(x^{k+1})- f_{i}(x^{k})\le \beta_{1}\alpha _{k}\sum\limits_{i = 1}^{m}\lambda_{i}^{k}(\nabla_{q}f_{i}(x^{k}))^{T}d_{q}(x^{k}), \end{align} (3.14)

    and

    \begin{align} \beta_{2}\sum\limits_{i = 1}^{m}\lambda_{i}^{k}(\nabla_{q} f_{i}(x^{k}))^{T}d_{q}(x^{k})\le \sum\limits_{i = 1}^{m}\lambda_{i}^{k}\nabla_{q}f_{i}(x^{k+1})d_{q}(x^{k}), \end{align} (3.15)

    where \beta\in(0, 1) , \beta_{1} < \beta_{2} < 1 , and \alpha \in (0, 1] . If cos^2(\theta_j^k) < \delta for any j then \alpha_k satisfying (3.14) is selected. The last inequality is a criterion for accepting a step-length in the multiobjective descent direction. We start with \alpha = 1 , and if the above condition is not satisfied, then set \alpha: = \frac{\alpha}{2} , and process will be continued.

    Proposition 1. Let \beta_{1}\in (0, 1), x^{k}\in X\subseteq\mathbb{R}^{n} , and d_{q}^{k}\in X . If f_{i}:\mathbb{R}^{n}\rightarrow\mathbb{R} for i = 1, \dots, m are continuously q -differentiable functions, and \nabla_{q}f_{i}(x^{k})d_{q}^{k} < 0, for i = 1, \dots, m then there exists \varepsilon > 0 such that f_{i}(x^{k}+\alpha d^{k})\le f_{i}(x^{k})+\beta_{1} \alpha\nabla_{q}f_{i}(x^{k}) d_{q}^{k} for all i = 1, \dots, m and \alpha \in(0, \varepsilon] .

    Proof. Since f_{i} for all i\in 1, \dots, m , are q -differentiable functions, then

    \begin{align} f_{i}(x^{k}+\alpha d_{q}^{k}) = f_{i}(x^{k})+\alpha(\nabla_{q}f_{i}(x^{k})d^{k}_{q}+R(t)), \end{align} (3.16)

    where \lim\limits_{t\rightarrow 0}R(t) = 0 . For \beta_{1}\in (0, 1) , we have

    \begin{align} (\nabla_{q}f_{i}(x^{k}))^{T}d_{q} \lt \beta_{1}(\nabla_{q}f_{i}(x^{k}))^{T}d_{q}, \end{align} (3.17)

    where i = 1, \dots, m . Combining (3.16) and (3.17), we get

    \begin{align} f_{i}(x^{k}+\alpha d_{q}^{k})\le f_{i}(x^{k})+\beta_{1} \alpha(\nabla_{q}f_{i}(x^{k})d^{k}_{q}, \end{align} (3.18)

    for all \alpha\in(0, \varepsilon] , i = 1, \dots, m.

    Let us now summarize the steepest descent method for multiobjective proposed in [19] based on q -derivative which is given in Algorithm 1 as:

    Algorithm 1 ends up with a critical point or produces (weak) efficient points. In this case, if \cos^{2}(\theta^k_i) > \delta , where i = 1, \dots, m is assumed, then assumptions of Theorem 1 are satisfied and every accumulation point of the sequence \{x^{k} \} will be a (weak) efficient point otherwise assumptions of Theorem 2 are considered and the accumulation point becomes a critical point of (UP).

    Theorem 1. Let f_{i} where i = 1, \dots, m be continuously q -differentiable on a set X\subseteq \mathbb{R}^{n} for \beta_{1}, q\in(0, 1) and \{x^{k}\} be the sequence updated by x^{k+1} = x^{k}+\alpha_{k}d_{q}(x^{k}) , where \alpha_{k} satisfies

    \begin{align} f_{i}(x^{k+1})- f_{i}(x^{k})\le \beta_{1}\alpha _{k}\sum\limits_{i = 1}^{m}\lambda_{i}^{k}(\nabla_{q}f_{i}(x^{k}))^{T}d_{q}(x^{k}) , \end{align} (4.1)

    for all i = 1, \dots, m. Suppose that L_{0} = \{x\in X:f(x) < f(x^{0}) \} is bounded and convex, where x^{0}\in X is a starting point. The function f_{i}(x) is bounded below for at least one i\in\{1, \dots, m\} . Then, the accumulation point of \{x^{k}\} is a critical point x^{*} of (UP).

    Proof. We know that

    \begin{align*} f_{i}(x^{k+1})- f_{i}(x^{k})\le \beta_{1}\alpha_{k}\sum\limits_{i = 1}^{m}\lambda_{j}^{k}(\nabla_{q}f_{i}(x^{k}))^{T}d_{q}(x^{k}). \end{align*}

    Since \sum_{i = 1}^{m}\lambda_{i}^{k} = 1, where \lambda_{i}^{k}\ge 0, then

    \begin{align*} f_{i}(x^{k+1})- f_{i}(x^{k})\le \beta_{1} \alpha_{k}\max\limits_{i = 1,\dots m} \big( (\nabla_{q}f_{i}(x^{k}))^{T} d_{q}(x^{k})\big), \end{align*}

    that is,

    \begin{align*} f_{i}(x^{k+1})-f_{i}(x^{k}) \lt \beta_{1}\alpha_{k} \max\limits_{i = 1,\dots m} \bigg[ (\nabla_{q}f_{i}(x^{k}))^{T} d_{q}(x^{k})+\frac{1}{(1+q)}\lVert d_{q} \rVert^{2}\bigg] = \beta_{1}\alpha_{k} d_{q}^{k}. \end{align*}

    We obtain

    \begin{align*} f_{i}(x^{k+1}) \lt f_{i}(x^{0})+c\sum\limits_{j = 0}^{k}\alpha_{i} d_{q}(x^{i})\;\;\text{for all}\;i = 1,\dots,m. \end{align*}

    For at least one i_{1} from i = 1, \dots, m for which f_{i}(x) is bounded below such that f_{i1}(x) > -\infty for all x\in X . The sequence \{f_{i1}(x^{k})\} is also monotonically decreasing sequence and bounded below so that f_{i1}(x^{k}) converges to f_{i1}(x^{*}) as k\rightarrow\infty where f_{i1}(x^{*}) > -\infty . We obtain the inequality as

    \begin{align*} f_{i1}(x^{0})-f_{i1}(x^{k+1}) \gt -\beta_{1}\sum\limits_{j = 0}^{k}\alpha_{j}d_{q}(x^{j}). \end{align*}

    As k\rightarrow\infty , we have

    \begin{align} \beta_{1} \sum\limits_{j = 0}^{\infty}\alpha_{j}(-d_{q}(x^{j})) \lt f_{i1}(x^{0})-f_{i1}(x^{*}) \lt \infty \end{align} (4.2)

    Since d_{q}(x^{j})\le 0 for all j due to [10,52], and \beta_{1} \sum_{j = 0}^{\infty}\alpha_{j}(-d_{q}(x^{j})) is finite. Thus, we obtain \beta_{1}\alpha_{k}(-d_{q}(x^{k}))\rightarrow 0 as k\rightarrow \infty . Since the step length is bounded above so \alpha_{k}\rightarrow\infty for some k implies L_{0} unbounded which is contradiction to the assumption. If \alpha_{k}\ge\beta_{1} for all k and for some \beta_{1}\in(0, 1) , then we get -d_{q}(x^{k})\rightarrow 0 as k\rightarrow\infty . Note that L_{0} is bounded sequence, and has at least one accumulation point. Let \{R_{1}^{*}, R_{2}^{*}, \dots, R_{r}^{*}\} be the set of accumulation points \{x^{k}\} . Since R_{s}^{*} is an accumulation point for every s\in\{1, 2, \dots, r\} , and d_{q} is a continuous function, then d_{q}(R_{s}^{*}) is a critical point of f_{s} for every s\in\{1, 2, \dots, r\} .

    Theorem 2. Let f_{i} for all i = 1, \dots, m be a continuously q -differentiable on a set X\subset \mathbb{R}^{n} , and \{x^{k}\} be the sequence by x^{k+1} = x^{k}+\alpha_{k}d^{k}_{q}(x^{k}) , and given that

    1.

    \begin{align} \beta_{2}\sum\limits_{i = 1}^{m}\lambda_{i}^{k}(\nabla_{q} f_{i}(x^{k}))^{T}d_{q}(x^{k})\le \sum\limits_{i = 1}^{m}\lambda_{i}^{k}\nabla_{q}f_{i}(x^{k+1})d_{q}(x^{k}),\;\mathit{\text{where}}\; \beta_{1} \lt \beta_{2} \lt 1, \end{align} (4.3)

    2. \lVert \nabla_{q} f_{i}(x^{k})-\nabla_{q} f_{i}(x^{k+1})\lVert\le L_{i}\lVert x^k-x^{k+1} \rVert, \; \forall\; x^k, x^{k+1}\in\mathbb{R}^{n}, i = 1, \dots, m, and

    3. \lVert\cos^{2}\theta_{i}^{k}\rVert\ge\delta for some \delta > 0, for all i = 1, \dots, m , where \theta_i^k is the angle between d_{q}(x^{k}) and \nabla_{q}f_{i}(x^k) .

    Then, every accumulation point of \{x^{k}\} generated is a weak efficient solution of (UP).

    Proof. We have proved that every accumulation point of \{x^{k}\} is a critical point of f_{i} , for all i = 1, \dots, m . Let x^{*} be an accumulation point of \{x^{k}\} . Fix one i_{0} from i = 1, \dots, m for which \nabla_{q}f_{i_{0}}(x^{*}) = 0 , then x^{*} will be a (weak) efficient solution. Form Cauchy-Schwartz inequality for all i = 1\dots, m , we obtain

    \begin{align*} ( \nabla_{q}f_{i}(x^{k+1})-\nabla_{q}f_{i}(x^{k}))^{T}d_{q}(x^{k})&\le \lVert \nabla_{q}f_{i}(x^{k+1})-\nabla_{q}f_{i}(x^{k}) \rVert\lVert d_{q}(x^{k}) \rVert \\ &\le L_{i}\lVert x^{k+1} -x^{k}\rVert \lVert d_{q}(x^{k}) \rVert \\ &\le L_{i}\alpha_{k}\lVert d_{q}(x^{k}) \rVert^{2}, \end{align*}

    Since L = \max L_{i} , where i = 1, 2\dots, m, then

    \begin{align*} ( \nabla_{q}f_{i}(x^{k+1})-\nabla_{q}f_{i}(x^{k}))^{T}d_{q}(x^{k})\le L\alpha_{k}\lVert d_{q}(x^{k}) \rVert^{2}. \end{align*}

    We obtain

    \begin{align*} L\alpha_{k}\lVert d_{q}(x^{k}) \rVert^{2}&\ge \max\limits_{i = 1,\dots m}(\nabla_{q}f_{i}(x^{k+1})- \nabla_{q}f_{i}(x^{k}))^{T} d_{q}(x^{k}) \\ &\ge\sum\limits_{i = 1}^{m}\lambda_{i}^{k}(\nabla_{q}f_{i}(x^{k+1})-\nabla_{q}f_{i}(x^{k}))^{T}d_{q}(x^{k}). \end{align*}

    From part 1 of this theorem,

    \begin{align*} L\alpha_{k}\lVert d_{q}(x^{k}) \rVert^{2} &\ge (\beta_{2}-1)\sum\limits_{i = 1}^{m}\lambda_{i}^{k} \nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k})\\ &\ge (\beta_{2}-1)\max\limits_{i = 1,\dots,m} \nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k}), \end{align*}

    this is,

    \begin{align*} \alpha_{k}\ge \frac{\beta_{2}-1}{L\lVert d_{q}(x^{k}) \rVert^{2}}\max\limits_{i = 1,\dots,m} \nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k}). \end{align*}

    Since \max_{i = 1, \dots, m} \nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k}) < 0, then

    \begin{align*} \alpha_{k}\bigg(\max\limits_{i = 1,\dots,m} \nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k})\bigg)\le \frac{\beta_{2}-1}{L\lVert d_{q}(x^{k}) \rVert^{2}}\bigg(\max\limits_{i = 1,\dots,m}\nabla_{q}f_{i}(x^{k})^{T} d_{q}(x^{k})\bigg)^2, \end{align*}

    that is,

    \begin{align*} -\beta_{1} \alpha_{k}\bigg(\max\limits_{i = 1,\dots,m}\nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k})\bigg)\ge \frac{\beta_{1}(\beta_{2}-1)}{L\lVert d_{q}(x^{k}) \rVert^{2}}\min\limits_{i = 1,\dots,m}\bigg( \nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k})\bigg)^2. \end{align*}

    Since (\nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k}))^2 = (\nabla_{q}f_{i}(x^{k})^{T})^2(d_{q}(x^{k}))^{2} (\cos^2\theta_{i}^{k}) , for all i = 1, \dots, m then

    \begin{align*} -\beta_{1} \alpha_{k}\bigg(\max\limits_{i = 1,\dots,m}\nabla_{q}f_{i}(x^{k})^{T}d_{q}(x^{k})\bigg) \ge\frac{\beta_{1}(1-\beta_{2})}{L}\min\limits_{i = 1,\dots,m}[\lVert \nabla_{q}f_{i}(x^{k})\rVert^{2}\cos^{2}(\theta_{i})^{T}], \end{align*}

    where \theta_{i}^{k} is the angle between \nabla_{q} f_{i}(x^{k}) and d_{q}(x^{k}). We have

    \begin{align*} \infty \gt f_{i1}(x^{0})-f_{i1}(x^{k+1})\ge -\beta_{1}\sum\limits_{j = 0}^{k}\alpha_{j}\max\limits_{y\in Y, i = 1,\dots,m}(\nabla_{q}f_{i}(x^{i}))^{T}d_{q}(x^{i})\\ = \sum\limits_{j = 0}^{k}\alpha_{j}(-\beta_{1}\max\limits_{y\in Y, i = 1,\dots,m}(\nabla_{q}f_{i}(x^{i}))^{T}d_{q}(x^{i})). \end{align*}

    Taking k\rightarrow\infty ,

    \begin{align*} \infty \gt f_{i1}(x^{0})-f_{i1}(x^{*})\ge \sum\limits_{j = 0}^{\infty}\alpha_{j}(-\beta_{1}\max\limits_{i = 1,\dots,m}(\nabla_{q}f_{i}(x^{i}))^{T}d_{q}(x^{i})). \end{align*}

    Since -\beta_{1}\max_{i = 1, \dots, m}\nabla_{q}f_{i}(x^{i})^{T}d_{q}(x^{i}) > 0, then

    \alpha_{k}(-\beta_{1}\max\limits_{i = 1,\dots,m}(\nabla_{q}f_{i}(x^{k}))^{T}d_{q}(x^{k}))\rightarrow0\; \text{as}\; k\rightarrow\infty.

    We also have

    \begin{align*} \frac{\beta_{1}(1-\beta_{2})}{L}\min\limits_{i = 1,\dots,m}[\lVert \nabla_{q}f_{i}(x^{k}) \rVert^{2}\cos^{2}\theta_{i}^{k}]\rightarrow0\;\;\;\text{as} \;\; k\rightarrow \infty. \end{align*}

    Since \cos^{2}\theta_{i}^{k} > \delta for i = 1, \dots, m , then \min_{i = 1, \dots, m}\lVert \nabla_{q}f_{i}(x^{k})\rVert^{2}\rightarrow0 as k\rightarrow\infty. Fix any i_{0} from i = 1, \dots, m such that \lVert \nabla_{q}f_{i0}(x^{k}) \rVert^{2}\rightarrow0 as k\rightarrow\infty. Since \lVert \nabla_{q}f_{i0}(x^{k}) \rvert is a continuous function, and \lVert \nabla_{q}f_{i0}(x^{k})\rVert\rightarrow0 as k\rightarrow\infty , then \nabla_{q}f_{i0}(x^{*}) = 0 for every accumulation points x^{*} of \{x^{k}\} . Thus, x^{*} is a local weak efficient solution.

    Iteration Complexity of Algorithm 1: Based on the ideas above, we consider iteration complexity. For a given iteration k , a point solution for a particular x^{k} , reaches to an optimal point under a desired tolerance \epsilon . The steepest descent direction moves from d_{q}^{k} to some d_{q}^{k+1} where d_{q}^{k+1}-d_{q}^{k} is small enough to be able to determine the associated solution on the Pareto front. For non-convex, Algorithm 1 has a convergence rate of the order of \frac{1}{\sqrt{k}} , and for the convex case, Algorithm 1 establishes the desired \frac{1}{k} rate for a certain sequence of weights \{\bar{\lambda}^{k}\} (see Theorem 1 of [29]) since for large value of k , q -gradient behaves as a classical gradient [43]. The global rates translate into worst-case complexity bounds of the order of \frac{1}{\epsilon^{2}} , \frac{1}{\epsilon} , and \log\frac{1}{\epsilon} iterations, respectively, to reach an approximate optimality criterion of the form \lVert d_{q}^{k}\rVert\le\epsilon for some \epsilon\in(0, 1) where d_{q}^{k} is the steepest descent direction (3.6). Due to the existence of a uniform lower bound on the step-length \alpha_{k} in Algorithm 1 will always stop in a finite number of steps.

    It is important to note that (weak) efficient solution of (UP) is not unique. Therefore, one can execute Algorithm 1 with any starting point to reach at one of these (weak) efficient points. User may consider a sufficiently large compact subset of \mathbb{R}^{n} as a domain of (UP) to solve. We have implemented algorithm 1 with the Armijo-Wolfe line search with backtracking. We have considered both convex and nonconvex problems. To avoid unbounded solutions of the subproblem at any iterating point x^k , we consider the following subproblem:

    \begin{align} &\underset{t,d_{q}}{\min }\; \; \; \; \; \; \; \; \; \; \; \; t+\frac{1}{2} d_q^Td_q \\ &\mbox{subject to}\; \; \nabla_q f_i(x^k)^Td_q\leq t\; \; i = 1,2,...,m, \\ &\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; lb\leq x^k+d_q\leq ub, \end{align} (4.4)

    where lb and ub are lower and upper bound of x^{k} , respectively. We now illustrate the methodology in numerical examples.

    Example 2 (One Dimension [55]). Consider the problem (UP): Minimize _{x\in\mathbb{R}} (f_{1}(x), f_{2}(x)) , where f_{1}(x) = x^{2} -4 and \; f_{2}(x) = (x- 1)^2 .

    This problem has two objective functions and one decision variable. Both functions are convex. The set of efficient solutions to this problem is [0, 1] which can be shown in Figure 1. Using Algorithm 1 with any starting point x^{0} > 1 , one can find the efficient point approximately equal to 1. At starting point x^{0} = 10 , the approximate solution is obtained at 29th step as x^{29} = 1.00240356 \approx 1 in [55]. Under the same starting point and stopping criterion, i.e, \epsilon_{1} = 10^{-7}, \epsilon_{2} = 10^{-11}, and \delta = 10^{-4}, our algorithm provides the efficient solution at 2nd step. The Pareto front using Algorithm 1 and [28] for -2\le x\le 2 is found as a convex curve shown in Figure 2.

    Figure 1.  Graph of Example 2.
    Figure 2.  (a) Approximate Pareto Front without q -derivative for Example 2. (b) Approximate Pareto Front with q -derivative for Example 2.

    Example 3 (Two Dimension). Consider an unconstrained multiobjective optimization problem [55] \text{minimize}_{x\in\mathbb{R}^{2}} (f_{1}(x), f_{2}(x)) , where f_{1}(x_{1}, x_{2}) = \frac{1}{4}\big[(x_{1}-1)^{4}+2(x_{2}-2)^{4}\big], f_{2}(x_{1}, x_{2}) = (x_{2}-x_{1}^{2})^{2}+(1-x_{1})^{2},

    After running the Algorithm 1, and methodology used in [28], we obtain the point x^{*} = (1, 2)^{T} and \lambda^{*}_{1} = 0.9774 , \lambda^{*}_{2} = 0.226 , in two iterations as final solution with a starting point x^{0} = (1, 3)^{T} . One can verify that

    \begin{align*} \lambda_{1}^{*}\nabla_{q}f_{1}(x^{*})+\lambda_{2}^{*}\nabla_{q}f_{2}(x^{*})\approx\begin{pmatrix} 0 \\ 0 \end{pmatrix}, \end{align*}

    Note that function f_{2} is not convex. A multiobjective optimization does not produce a unique solution but a set of efficient solutions. The algorithm does not depend upon the choice of the starting point. The sequence converges to one of the sets of critical or weak efficient solutions, starting at any starting point. For example, with starting point x^{0} = (0, 0)^{T} , Algorithm 1 converges to x^{*} = (0.9975, 0.9936)^{T} in 60 iterations, while methodology used in [28] converges to x^{*} = (0.9981, 0.9950) in 64 iterations.

    Example 4 (Two Dimension Non-Convex Problem). Consider the test problem (DTLZP) from [56]: \underset{0\leq x_1, x_2\leq 1}{\min}\; \; \; (f_{1}(x_{1}, x_{2}), f_{2}(x_{1}, x_{2})) , where f_{1}(x_{1}, x_{2}) = (1 + gx)\cos(0.5 \pi x_1) , f_{2}(x_{1}, x_{2}) = (1+ gx)\sin(0.5 \pi x_1) , gx = (x_{2}- 0.5)^2 .

    Consider a starting point x^0 = (0.5060, 0.6991)^T . Then, we have f(x^0) = (0.7282, 0.7420)^T . Note that both f_1 and f_2 are non-convex. At x^0 , the q -gradients are |\nabla_q f_1(x^0)| = -2.7161 < 0 and |\nabla_q f_2(x^0)| = -2.8055 < 0, respectively. Solution of subproblem at x^0 is t^0 = -0.0793 , and steepest descent direction is d_q^0 = (0.0006, -0.2816)^T . Because \cos^2(\theta_1) = 0.0552 > 10^{-4} and \cos^2(\theta_2) = 0.0571 > 10^{-4} , thus step length \alpha_0 = 1 is selected to satisfying both (3.14) and (3.15). Next iterating point is x^1 = x^0+\alpha_0d_q^0 = (0.5066, 0.4175) . Clearly, we get f(x^1) = (0.7044, 0.7193)^T < f(x^0) . Final solution is obtained at 7^{th} step as x^7 = (0.5068, 0.5010)^T . One can verify that x^7 is an approximate critical point since

    \|0.4957 \nabla_q f_1(x^7)+0.5053\nabla_q f_2(x^7)\| = 1.4\times 10^{-3}\approx 0.

    Comparison of approximate Pareto front with weighted sum method: To obtain an approximate Pareto front we have considered a multi-start technique. Here, 100 uniformly distributed random points are selected to execute Algorithm 1 and weighted sum (WS) algorithm individually. Approximate Pareto front obtained by Algorithm 1 is compared with the approximate Pareto front obtained by the weighted sum method. In (WS) method we have used weights (1, 0) , (0, 1) , and 98 random weights. The single objective optimization problem is solved by a single objective steepest descent method with some random initial approximation. Approximate Pareto fronts generated by Algorithm 1 for Example 2 and Example 4 are given in Figure 3(a), (b), respectively. One can observe that Algorithm 1 generates an approximate Pareto front for both problems but (WS) fails for Example 4 (non-convex problem). Table 1 lists the average of number of iterations it , that of function evaluations evalf , that of gradient evaluations evalg and that of subproblem solved evalsp , respectively to show the performance of q -steepest descent method for (UP). The reported data in Table 1 represent a typical execution of algorithm 1. We conclude that Algorithm 1 performs better in terms of the number of iterations than the method of [28]. It is noteworthy that the number of steepest descent direction calculations is equal to the number of iterations.

    Figure 3.  (a) Approximate Pareto Front for Example 2. (b) Approximate Pareto front for Example 4.
    Table 1.  Performance of q-Steepest Descent Method.
    Problem (q-SDUP) [28]
    it evalf evalg evalsp it evalf evalg evalsp
    Example 2 171 238 238 1623 176 250 176 1929
    Example 4 684 777 775 4561 705 715 714 4903

     | Show Table
    DownLoad: CSV

    Example 5 (Three Dimension). Consider multiobjective problem: Minimize (f_{1}(x), f_{2}(x), f_{3}(x)) , where f_{1}(x_{1}, x_{2}, x_{3}) = x_{1}^{2}+x_{2}^{2}+x_{3}^{2} , f_{2}(x_{1}, x_{2}, x_{3}) = (x_{1}-2)^2+x_{2}^2+x_{3}^{2} , f_{3}(x_{1}, x_{2}, x_{3}) = x_{1}^{2}+x_{2}^{2}+(x_{3}-2)^{2} .

    We solve the above problem with the method described in [28] and compare it with the proposed method of this paper. Consider a starting point x^{0} = (2, 1, 3)^{T}, and stopping criteria \min\{\nabla_{q} f_{1}, \nabla_{q} f_{2}, \nabla_{q} f_{3}\} < 10^{-5} . A program in MATLAB (2017a) is written in the lines of Algorithm 1. Of course, the search process moves from global at the beginning to particularly neighborhood at last due to the advantage of q -derivative shown in Table 2. The solution to this problem is found at 14th iterations, and this point is a local weak efficient solution of the above problem.

    Table 2.  Computational Details of Example 5.
    k x^{k} \alpha_{k} d_{q}^{k} x^{k+1} [\lambda_{1}^{k+1}, \lambda_{2}^{k+1}, \lambda_{3}^{k+1}] \min{\cos(\theta_{k}^{i})}
    0 \begin{pmatrix} 2 \\ 1 \\ 3 \end{pmatrix} 0.5 \begin{pmatrix} -34.6980 \\ -4.3275 \\ -55.3018 \end{pmatrix} \begin{pmatrix} 1.6510 \\ 7.8363 \\ 0.3491 \end{pmatrix} [ 0.0000, 0.8250, 0.1750] 1.9734
    1 \begin{pmatrix} 1.6510 \\ 7.8363 \\ 0.3491 \end{pmatrix} 0.5 \begin{pmatrix} 0.6983 \\ -0.9999 \\ -0.6971 \end{pmatrix} \begin{pmatrix} 2.3493 \\ 6.8364 \\ -0.3480 \end{pmatrix} [0.0000, 1.0000, 0.0000] 1.9738
    2 \begin{pmatrix} 2.3493 \\ 6.8364 \\ -0.3480 \end{pmatrix} 0.5 \begin{pmatrix} -0.6987 \\ -1.0000 \\ 0.6969 \end{pmatrix} \begin{pmatrix} 1.6506\\ 5.8364 \\ 0.3489 \end{pmatrix} [0.0000 1.0000 0.0000] 1.9751
    3 \begin{pmatrix} 1.6506\\ 5.8364 \\ 0.3489 \end{pmatrix} 0.5 \begin{pmatrix} 0.6990 \\ -1.0000 \\ -0.6975 \end{pmatrix} \begin{pmatrix} 2.3496 \\ 4.8364 \\ -0.3486 \end{pmatrix} [0.0000 1.0000 0.0000] 1.9758
    4 \begin{pmatrix} 2.3496 \\ 4.8364 \\ -0.3486 \end{pmatrix} 0.5 \begin{pmatrix} -0.6989 \\ -1.0000\\ 0.6981 \end{pmatrix} \begin{pmatrix} 1.6507\\ 3.8364 \\ 0.3495 \end{pmatrix} [0.0000 1.0000 0.0000] 1.9769
    5 \begin{pmatrix} 1.6507\\ 3.8364 \\ 0.3495 \end{pmatrix} 0.5 \begin{pmatrix} 0.6990\\-0.9999\\-0.6990 \end{pmatrix} \begin{pmatrix} 2.3496 \\ 2.8365 \\ -0.3495 \end{pmatrix} [0.0000 1.0000 0.0000] 1.9774
    6 \begin{pmatrix} 2.3496 \\ 2.8365 \\ -0.3495 \end{pmatrix} 0.5 \begin{pmatrix} -0.6993 \\ -1.0000 \\ 0.6989 \end{pmatrix} \begin{pmatrix} 1.6503 \\ 1.8365 \\ 0.3493 \end{pmatrix} [0.0000 1.0000 0.0000] 1.7603
    7 \begin{pmatrix} 1.6503 \\ 1.8365 \\ 0.3493 \end{pmatrix} 0.5 \begin{pmatrix} 0.4009 \\ -1.1996\\ -0.4007 \end{pmatrix} \begin{pmatrix} 1.8507 \\ 1.2367 \\ 0.1490 \end{pmatrix} [0  0.9254   0.0746] 1.1536
    8 \begin{pmatrix} 1.8507 \\ 1.2367 \\ 0.1490 \end{pmatrix} 0.5 \begin{pmatrix} 0.1917 \\ -1.0393 \\ -0.1915 \end{pmatrix} \begin{pmatrix} 1.9466 \\ 0.7170 \\ 0.0533 \end{pmatrix} [0 0.9733 0.0267] 1.0172
    9 \begin{pmatrix} 1.9466 \\ 0.7170 \\ 0.0533 \end{pmatrix} 0.5 \begin{pmatrix} 0.0546 \\ -1.0056 \\ -0.0539 \end{pmatrix} \begin{pmatrix} 2.0012 \\ -0.2886 \\ -0.0007 \end{pmatrix} [0 0.9869 0.0131] 0.8673
    10 \begin{pmatrix} 2.0012 \\ -0.2886 \\ -0.0007 \end{pmatrix} 0.5 \begin{pmatrix} -0.3413 \\ -0.8665 \\ 0.0010 \end{pmatrix} \begin{pmatrix} 1.8305 \\ -0.7219 \\ -0.0002 \end{pmatrix} [0.0843 0.9153 0.0004] 0.579
    11 \begin{pmatrix} 1.8305 \\ -0.7219 \\ -0.0002 \end{pmatrix} 0.5 \begin{pmatrix} -0.3678 \\ -0.5775 \\ 0 \end{pmatrix} \begin{pmatrix} 1.6542 \\ -1.0106 \\ -0.0002 \end{pmatrix} [0.0865  0.8271 0.0865] 0.1452
    12 \begin{pmatrix} 1.6542 \\ -1.0106 \\ -0.0002 \end{pmatrix} 0.5 \begin{pmatrix} -0.2297 \\ -0.3041\\ 0 \end{pmatrix} \begin{pmatrix} 1.4245 \\ -1.3147 \\ -0.0002 \end{pmatrix} [0 0.7697 0.2303] 0.0011
    13 \begin{pmatrix} 1.4245 \\ -1.3147 \\ -0.0002 \end{pmatrix} 0.5 \begin{pmatrix} 0.0221\\ 0.0243 \\ 0 \end{pmatrix} \begin{pmatrix} 1.4466\\ -1.2904 \\ -0.0002 \end{pmatrix} [0.1411 0.7178 0.1411] 4.5278e-05
    14 \begin{pmatrix} 1.4466\\ -1.2904 \\ -0.0002 \end{pmatrix} 0.5 \begin{pmatrix} -0.0045 \\ -0.0050 \\ 0 \end{pmatrix} \begin{pmatrix} 1.4421 \\ -1.2954 \\ -0.0002 \end{pmatrix} [0.2779 0.7221 0] 1.5914e-06
    15 \begin{pmatrix} 1.4421 \\ -1.2954 \\ -0.0002 \end{pmatrix} 0.5 \begin{pmatrix} 0.8427\times 10^{-3}\\ 0.9387\times 10^{-3}\\0 \end{pmatrix} \begin{pmatrix} 1.4421\\ -1.2954 \\ -0.0002 \end{pmatrix} [0.2788 0.7212 0.0000] 0.0013

     | Show Table
    DownLoad: CSV

    Example 6 (Two Dimension Problem). Consider multimodal multiobjective optimization problem [57]: \underset{-5\leq x_1, x_2\leq 5}{minimize}\; \; \; (f_{1}(x_{1}, x_{2}), f_{2}(x_{1}, x_{2})) , where f_{1}(x_{1}, x_{2}) = 100(x_{1}^{2}-x_{2})^{2}+(x_{1}-1)^{2} , f_{2}(x_{1}, x_{2}) = 100(x_{1}^{2}-x_{2})^{2}+(x_{1}-2)^{2} .

    Functions f_{1} and f_{2} have global minimizers at x^{*} = (1, 1)^{T} and \hat{x} = (2, 4)^{T} , respectively. It is noteworthy the f_{1}(x^{*}) = 0 = f_{2}(\hat{x}) = 0 and f_{1}(\hat{x}) = f_{2}(x^{*}) = 1. We execute the Algorithm 1 100 times using starting points from a uniform random distribution belonging to ]-5, 5[ \times ]-5, 5[ . In all instances, Algorithm 1 stopped at a point satisfying the convergence criterion. We see that Algorithm 1 could estimate the Pareto front as shown in Figure 4.

    Figure 4.  Approximate Pareto Front for Example 6.

    We select the work of [28] to compare the proposed method. The basic difference between the method of [28] and our method is that q -gradient based steepest descent algorithm gives faster convergence for q\in (0, 1) . The q -drivative evaluates tangent, computes the secant of the objective functions, and therefore takes larger steps towards the optimum solution. Table 3 summarizes the results after executing the method of [28] and the proposed algorithm of this paper, with several starting points for the problem of Example 5. Both methods are solved using the MATLAB (2017a) program. It is seen that number of iterations using Algorithm 1 is less than a number of iterations in the methodology by [28].

    Table 3.  Comparison Results with [28] for Example 5.
    Sl. No. Starting Point Number of iterations in Algorithm 1 Final result in Algorithm 1 Number of iterations in [28] Final result in [28]
    1 \begin{pmatrix} 1 \\ 5 \\ 6 \end{pmatrix} 10 \begin{pmatrix} 1.4434 \\ -1.2907 \\ 0.0035 \end{pmatrix} 15 \begin{pmatrix} 1.3652 \\ -1.0703 \\ -0.0000 \end{pmatrix}
    2 \begin{pmatrix} 3 \\ 4 \\ 1 \end{pmatrix} 10 \begin{pmatrix} 1.4117 \\ -1.2015 \\ 0.0003 \end{pmatrix} 14 \begin{pmatrix} 1.4472 \\ -1.3035 \\ -0.0000 \end{pmatrix}
    3 \begin{pmatrix} 5 \\ 7 \\ 3 \end{pmatrix} 13 \begin{pmatrix} 1.4048 \\ -1.1773 \\ 0.0001 \end{pmatrix} 18 \begin{pmatrix} 1.3975 \\ -1.1569 \\ -0.0000 \end{pmatrix}
    4 \begin{pmatrix} 10\\ 8\\ 9 \end{pmatrix} 13 \begin{pmatrix} 1.4559 \\ -1.3422 \\ -0.0011 \end{pmatrix} 16 \begin{pmatrix} 1.2051 \\ -0.7542 \\ -0.0000 \end{pmatrix}
    5 \begin{pmatrix} 7 \\ 3 \\ 8 \end{pmatrix} 9 \begin{pmatrix} 1.3421 \\ -1.0236 \\ 0.3291 \end{pmatrix} 13 \begin{pmatrix} 1.2603 \\ -0.8490 \\ 0.3862 \end{pmatrix}
    6 \begin{pmatrix} 2 \\ 10 \\ 7 \end{pmatrix} 17 \begin{pmatrix} 1.3766 \\ -1.1076 \\ 0.1774 \end{pmatrix} 20 \begin{pmatrix} 1.3556 \\ -1.0486 \\ 0.1748 \end{pmatrix}
    7 \begin{pmatrix} 5 \\ 7 \\ 6 \end{pmatrix} 15 \begin{pmatrix} 1.3805 \\ -1.1174 \\ 0.1072 \end{pmatrix} 17 \begin{pmatrix} 1.3793 \\ -1.1073 \\ 0.1094 \end{pmatrix}
    8 \begin{pmatrix} -9 \\ -5 \\ -1 \end{pmatrix} 8 \begin{pmatrix} 1.2285 \\ -0.7931 \\ 0.0020 \end{pmatrix} 8 \begin{pmatrix} 1.2274 \\ -0.7917 \\ -0.0000 \end{pmatrix}
    9 \begin{pmatrix} 3 \\ 9 \\ 5 \end{pmatrix} 17 \begin{pmatrix} 1.4242 \\ -1.2351 \\ 0.0938 \end{pmatrix} 19 \begin{pmatrix} 1.3861 \\ -1.1252 \\ 0.0935 \end{pmatrix}

     | Show Table
    DownLoad: CSV

    We have applied the gradient descent approach to solve (UP), and gave necessary optimality condition based on q -derivative. We proved the convergence based on the proposed algorithm. The steepest descent method for multiobjective optimization with q-derivative converges independently of the starting point to a critical Pareto point. The solution strategy does not require any ordering information. The proposed algorithm has been compared with the existing method to measure the efficiency of algorithm.

    The authors would like to sincerely thanks the anonymous referees for their valuable comments and suggestions, which helped us to improve the article. The research is supported by the Science and Engineering Research Board (Grant No. DST-SERB-MTR-2018/000121) and the University Grants Commission (IN) (Grant No. UGC-2015-UTT–59235).

    The authors declare that they have no competing interests.



    [1] Marusciac, On Fritz John type optimality criterion in multi-objective optimization, Anal. Numer. Theor. Approx., 11 (1982), 109-114.
    [2] G. Palermo, C. Silvano, S. Valsecchi, et al. A system-level methodology for fast multi-objective design space exploration, ACM, New York, 2003.
    [3] M. Tavana, A subjective assessment of alternative mission architectures for the human exploration of mars at NASA using multicriteria decision making, Comput. Oper. Res., 31 (2004), 1147-1164.
    [4] H. Eschenauer, J. Koski, A. Osyczka, Multicriteria design optimization: Procedures and applications, Springer-Verlag, Berlin, 1990.
    [5] I. Das, J. E. Dennis, Normal-Boundary intersection: A new method for generating the Pareto surface in nonlinear multicriteria optimization problems, SIAM J. Optim.,8 (1998), 631-657.
    [6] I. Y. Kim, O. L. de Weck, Adaptive weighted-sum method for bi-objective optimization: Pareto front generation, Struct. Multidisc. Optim., 29 (2005), 149-158.
    [7] Y. Fu, U. M. Diwekar, An efficient sampling approach to multiobjective optimization, Ann. Oper. Res., 132 (2004), 109-134.
    [8] A. Jüschke, J. Jahn, A. Kirsch, A bicriterial optimization problem of antenna design, Comput. Optim. Appl., 7 (1997), 261-276.
    [9] T. M. Leschine, H. Wallenius, W. A. Verdini, Interactive multiobjective analysis and assimilative capacity-based ocean disposal decisions, Eur. J. Oper. Res., 56 (1992), 278-289.
    [10] J. Fliege, OLAF-A general modeling system to evaluate and optimize the location of an air polluting facility, OR Spektrum, 23 (2001), 117-136.
    [11] E. Carrizosa, J. B. G. Frenk, Dominating sets for convex functions with some applications, J. Optim. Theory Appl., 96 (1998), 281-295.
    [12] G. W. Evan, An overview of techniques for solving multiobjective mathematical programs, Manage. Sci., 30 (1984), 1268-1282.
    [13] D. Prabuddha, J. B. Ghosh, C. E. Wells, On the minimization of completion time variance with a bicriteria extension, Oper. Res., 40 (1992), 1148-1155.
    [14] D. J. White, Epsilon-dominating solutions in mean-variance portfolio analysis, Eur. J. Oper. Res., 105 (1998), 457-466.
    [15] M. G. C. Tapia, C. A. C. Coello, Applications of multi-objective evolutionary algorithms in economics and finance: A survey, IEEE Congr. Evol. Comput., (2007), 532-539.
    [16] E. Zitzler, K. Deb, L. Thiele, Comparison of multiobjective evolutionary algorithms: Empirical results, Evol. Comput., 8 (2000), 173-195.
    [17] M. Ehrgott, Multicriteria optimization, Springer, Berlin, 2005.
    [18] K. Miettinen, M. M. Makelä, Interactive bundle-based method for nondifferentiable multiobjeective optimization: NIMBUS, Optimization, 34 (1995), 231-246.
    [19] L. M. G. Drummond, B. F. Svaiter, A steepest descent method for vector optimization, J. Comput. Appl. Math., 175 (2005), 395-414.
    [20] J. Fliege, L. M. G. Drummond, B. F. Svaiter, Newton's method for multiobjective optimization, SIAM J. Optim., 20 (2009), 602-626.
    [21] A. Cauchy, Méthode générale pour la résolution des systèms d'équations simultanées, Comp. Rend. Sci. Paris, 25 (1847), 536-538.
    [22] H. Zhao, J. Zheng, W. Deng, et al. Semi-supervised broad learning system based on manifold regularization and broad network, IEEE Trans. Circuits Syst., 67 (2020), 983-994.
    [23] W. Deng, J. Xu, Y. Song, et al. An effective improved co-evolution ant colony optimization algorithm with multi-strategies and its application, Int. J. Bio-Inspir. Com., (2020) 1-10.
    [24] W. Deng, J. Xu, H. Zhao, An improved ant colony optimization algorithm based on hybrid strategies for scheduling problem, IEEE access, 7 (2019), 20281-20292.
    [25] W. Deng, H. Liu, J. Xu, et al. An improved quantum-inspired differential evolution algorithm for deep belief network, IEEE Trans. Instrum. Meas., 2020.
    [26] S. K. Mishra, B. Ram, Introduction to Unconstrained Optimization with R, Springer Nature, Singapore, 2019.
    [27] J. Fliege, B. F. Svaiter, Steepest descent methods for multicriteria optimization, Math. Method Oper. Res., 51 (2000), 479-494.
    [28] T. D. Chuong, J. C. Yao, Steepest descent methods for critical points in vector optimization problems, Appl. Anal., 91 (2012), 1811-1829.
    [29] J. Fliege, A. I. F. Vaz, L. N. Vicente, Complexity of gradient descent for multiobjective optimization, Optim. Method Softw., 34 (2019), 949-959.
    [30] F. H. Jackson, On q-functions and a certain difference operator, Trans. Roy Soc. Edin., 46 (1908) 253-281.
    [31] F. H. Jackson, On q-Definite Integrals, Pure Appl. Math. Q., 41 (1910), 193-203.
    [32] A. Aral, V. Gupta, R. P. Agarwal, Applications of q-calculus in operator theory, Springer, New York, 2013.
    [33] P. M. Rajković, M. S. Stanković, S. D. Marinković, Mean value theorems in q-calculus, Mat. Vesn., 54 (2002), 171-178.
    [34] M. E. Ismail, D. Stanton, Applications of q-Taylor theorems, J. Comput. Appl. Math., 153 (2003), 259-272.
    [35] S. C. Jing, H. Y. Fan, q-Taylor's formula with its q-remainder, Commun. Theor. Phys., 23 (1995), 117.
    [36] P. M. Rajković, S. D. Marinković, M. S. Stanković, Fractional integrals and derivatives in qcalculus, Appl. Anal. Discrete Math., 1 (2007), 311-323.
    [37] H. Gauchman, Integral inequalities in q-calculus, Comput. Math. Appl., 47 (2004), 281-300.
    [38] G. Bangerezako, Variational q-calculus, J. Math. Anal. Appl., 289 (2004), 650-665.
    [39] L. Abreu, A q-sampling theorem related to the q-Hankel transform, Proc. Am. Math. Soc., 133 (2005), 1197-1203
    [40] T. H. Koornwinder, R. F. Swarttouw, On q-analogues of the Fourier and Hankel transforms, Trans. Am. Math. Soc., 333 (1992), 445-461.
    [41] J. Ablinger, A. K. Uncu, qFunctions-A Mathematica package for q-series and partition theory applications, arXiv preprint arXiv:1910.12410, 2019.
    [42] A. C. Soterroni, R. L. Galski, F. M. Ramos, The q-gradient vector for unconstrained continuous optimization problems, In: Operations research proceedings 2010. Springer-Verlag Berlin Heidelberg, (2011), 365-370.
    [43] A. C. Soterroni, R. L. Galski, F. M. Ramos, The q-gradient method for global optimization, arXiv:1209.2084, 2012.
    [44] A. C. Soterroni1, R. L. Galski, M. C. Scarabello1, et al. The q-G method: A q-version of the steepest descent method for global optimization, SpringerPlus, 4 (2015), 647.
    [45] A. C. Soterroni, R. L. Galski, F. M. Ramos, Satellite constellation design using the q-G global optimization method, Third World Conference on Complex Systems (WCCS), IEEE, (2015), 1-6.
    [46] U. M. Al-Saggaf, M. Moinuddin, M. Arif, et al. The q-least mean squares algorithm, Signal Process., 111 (2015), 50-60.
    [47] A. Sadiq, S. Khan, I. Naseem, et al. Enhanced q-least mean square, Circ. Syst. Signal Process., 38 (2019), 4817-4839.
    [48] P. M. Rajković, S. D. Marinković, M. S. Stanković, On q-Newton-Kantorovich method for solving systems of equations, Appl. Math. Comput., 168 (2005), 1432-1448.
    [49] S. K. Chakraborty, G. Panda, Newton like line search method using q-calculus, In: D. Giri, R. N. Mohapatra, H. Begehr, M. Obaidat, Communications in Computer and Information Science, Singapore: Springer, 2017.
    [50] S. K. Mishra, G. Panda, M. A. T. Ansary, et al. On q-Newton's method for unconstrained multiobjective optimization problems, J. Appl. Math. Comput., (2020), 1-20.
    [51] Ž. Povalej, Quasi-Newton's method for multiobjective optimization, J. Comput. Appl. Math., 255 (2014), 765-777.
    [52] K. K. Lai, S. K. Mishra, B. Ram, On q-Quasi-Newton's Method for Unconstrained Multiobjective Optimization Problems, Mathematics, 8 (2020), 616.
    [53] M. A. T. Ansary, G. Panda, A sequential quadratically constrained quadratic programming technique for a multi-objective optimization problem, Eng. Optim., 51 (2019), 22-41.
    [54] W. E. Karush, Minima of functions of several variables with inequalities as side conditions. PhD thesis, Univ. Chicago, 1939. 55. M. A. Ansary, G. Panda, A modified quasi-Newton method for vector optimization problem, Optimization, 64 (2015), 2289-2306.
    [55] M. A. T. Ansary, G. Panda, A modified quasi-Newton method for vector optimization problem, Optimization, 64 (2015), 2289-2306.
    [56] K. Deb, L. Thiele, M. Laumanns, et al. Scalable multi-objective optimization test problems, IEEE, 1 (2002), 825-830.
    [57] O. P. Ferreira, M. S. Louzeiro, L. F. Prudente, Iteration-complexity and asymptotic analysis of steepest descent method for multiobjective optimization on Riemannian manifolds, J. Optim. Theory Appl., 184 (2020), 507-533.
  • This article has been cited by:

    1. Shashi Kant Mishra, Suvra Kanti Chakraborty, Mohammad Esmael Samei, Bhagwat Ram, A q-Polak–Ribière–Polyak conjugate gradient algorithm for unconstrained optimization problems, 2021, 2021, 1029-242X, 10.1186/s13660-021-02554-6
    2. Mohammad Esmael Samei, Ahmad Ahmadi, Sayyedeh Narges Hajiseyedazizi, Shashi Kant Mishra, Bhagwat Ram, The existence of nonnegative solutions for a nonlinear fractional q-differential problem via a different numerical approach, 2021, 2021, 1029-242X, 10.1186/s13660-021-02612-z
    3. Kin Keung Lai, Mohd Hassan, Jitendra Kumar Maurya, Sanjeev Kumar Singh, Shashi Kant Mishra, Multiobjective Convex Optimization in Real Banach Space, 2021, 13, 2073-8994, 2148, 10.3390/sym13112148
    4. Danijela Protic, Miomir Stankovic, Vladimir Antic, 2023, Chapter 44, 978-981-19-8492-1, 595, 10.1007/978-981-19-8493-8_44
    5. Dinesh Kumar, Geetanjali Panda, A line search technique for a class of multi-objective optimization problems using subgradient, 2024, 28, 1385-1292, 10.1007/s11117-024-01051-6
    6. Kin Keung Lai, Shashi Kant Mishra, Bhagwat Ram, Ravina Sharma, A Conjugate Gradient Method: Quantum Spectral Polak–Ribiére–Polyak Approach for Unconstrained Optimization Problems, 2023, 11, 2227-7390, 4857, 10.3390/math11234857
    7. Bhagwat Ram, Shashi Kant Mishra, Kin Keung Lai, Predrag Rajković, 2024, Chapter 1, 978-981-97-2434-5, 1, 10.1007/978-981-97-2435-2_1
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(6747) PDF downloads(393) Cited by(7)

Figures and Tables

Figures(4)  /  Tables(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog