删除或更新信息,请邮件至freekaoyan#163.com(#换成@)

New project gradient methods for quadratic programming with linear equality constraints

本站小编 Free考研考试/2021-12-25

李明强, 郭田德, 韩丛英
中国科学院大学数学科学学院, 北京 100049
摘要: 近些年来,众多****提出基于新步长选择策略的加速梯度投影算法求解大规模优化问题。本文针对线性约束二次规划问题提出两种基于新步长的梯度投影算法。一种是基于采用自适应线搜索和Barzilai-Borwein步长的非单调投影算法。另一种是基于Yuan步长的单调投影算法。在较弱的假设条件下,给出这两种算法的全局收敛性。数值实验表明新算法比传统的梯度投影算法求解效率更高。
关键词: Barzilai-Borwein步长线性等式投影梯度
In this paper, we consider the quadratic programming problem
$\mathop {\min }\limits_{x \in \Omega } f\left( \mathit{\boldsymbol{x}} \right) = \frac{1}{2}{\mathit{\boldsymbol{x}}^{\rm{T}}}\mathit{\boldsymbol{Qx}} + {\mathit{\boldsymbol{c}}^{\rm{T}}}\mathit{\boldsymbol{x}},$ (1)
where QRn×n is symmetric and positive definite, x, cRn, and the feasible region Ω is defined by
$\Omega = \left\{ {\mathit{\boldsymbol{x}} \in {{\bf{R}}^n},\mathit{\boldsymbol{Ax}} = \mathit{\boldsymbol{b}}} \right\},$ (2)
where ARm×n, rank (A)=m, and m < n.
First of all, we give a brief review for the development of the efficient step length selections in gradient methods for minimizing a large scale strictly convex quadratic function f(x). Assume that g(x)=?f(x) can be obtained at every x and gk=g(xk) is the gradient at xk. The classical examples of step size selections are the line searches used in the steepest descent (SD) and the minimal gradient (MG) methods, which minimize f(xk-αgk) and ‖g(xk-αgk)‖2, respectively,
$\begin{array}{l}\alpha _k^{SD} = \mathop {\arg \min }\limits_\alpha f\left( {{\mathit{\boldsymbol{x}}_k} - \alpha {\mathit{\boldsymbol{g}}_k}} \right)\\\;\;\;\;\;\; = \frac{{\mathit{\boldsymbol{g}}_k^{\rm{T}}{\mathit{\boldsymbol{g}}_k}}}{{\mathit{\boldsymbol{g}}_k^{\rm{T}}\mathit{\boldsymbol{Q}}{\mathit{\boldsymbol{g}}_k}}},\end{array}$ (3)
$\begin{array}{l}\alpha _k^{MG} = \mathop {\arg \min }\limits_\alpha {\left\| {\mathit{\boldsymbol{g}}\left( {{\mathit{\boldsymbol{x}}_k} - \alpha {\mathit{\boldsymbol{g}}_k}} \right)} \right\|_2}\\\;\;\;\;\;\; = \frac{{\mathit{\boldsymbol{g}}_k^{\rm{T}}\mathit{\boldsymbol{Q}}{\mathit{\boldsymbol{g}}_k}}}{{\mathit{\boldsymbol{g}}_k^{\rm{T}}{\mathit{\boldsymbol{Q}}^2}{\mathit{\boldsymbol{g}}_k}}}.\end{array}$ (4)
Though the SD gradient method converges linearly, the convergence rate may be very slow, especially when Q is ill-conditioned in the sense that the condition number cond (Q) is very large. The cond (Q) is defined as follows:
${\rm{cond}}\left( \mathit{\boldsymbol{Q}} \right) = \frac{{{\lambda _n}\left( \mathit{\boldsymbol{Q}} \right)}}{{{\lambda _1}\left( \mathit{\boldsymbol{Q}} \right)}},$
where λn(Q) and λ1(Q) represent the largest and smallest eigenvalues of Q, respectively.
In 1988, a significant step size selection is proposed by Barzilai and Borwein (see Ref.[1]),
$\alpha _k^{BB1} = \frac{{\mathit{\boldsymbol{s}}_{k - 1}^{\rm{T}}{\mathit{\boldsymbol{s}}_{k - 1}}}}{{\mathit{\boldsymbol{s}}_{k - 1}^{\rm{T}}{\mathit{\boldsymbol{y}}_{k - 1}}}},$ (5)
$\alpha _k^{BB2} = \frac{{\mathit{\boldsymbol{s}}_{k - 1}^{\rm{T}}{\mathit{\boldsymbol{y}}_{k - 1}}}}{{\mathit{\boldsymbol{y}}_{k - 1}^{\rm{T}}{\mathit{\boldsymbol{y}}_{k - 1}}}},$ (6)
where sk-1=xk-xk-1 and yk-1=gk-gk-1. We refer to the step size (5) or (6) as BB (Barzilai and Borwein) step size. From (3) and (4), we can easily obtained the fact that in the unconstrained case, αkBB1=αk-1SD and αkBB2=αk-1MG (see Ref.[2]). In other words, the step lengths (5) and (6) use the SD and MG step lengths with one delay, respectively. A feature of BB step size is that the sequences {f(xk)} and {‖gk2} are non-monotone, in contrast to the SD and MG methods. Global convergence of the BB method was established by Raydan[3] in the strictly convex quadratic case with either (5) or (6). Dai and Liao[4] have shown that the BB method is R-linearly convergent in the n-dimensional case. Moreover, numerical results indicate that the BB method outperforms the method with (3) or (4) for convex quadratic functions (see also Ref.[2]). Starting from (5) and (6), BB-like methods with longer delays have been studied[5-7]. In Ref.[5], Dai and Fletcher proposed a formula
$\alpha _k^M = \frac{{\sum\nolimits_{i = 1}^M {\mathit{\boldsymbol{s}}_{k - i}^{\rm{T}}{\mathit{\boldsymbol{s}}_{k - i}}} }}{{\sum\nolimits_{i = 1}^M {\mathit{\boldsymbol{s}}_{k - i}^{\rm{T}}{\mathit{\boldsymbol{y}}_{k - i}}} }},$ (7)
where M is a positive integer, and they suggest that M=2 is a better choice for box-constrained quadratic programming.
Recently, some monotone methods which are competitive to the BB-like method are proposed. One of these methods with step length
$\alpha _k^Y = \frac{2}{{\varphi _k^Y + 1\alpha _{k - 1}^{SD} + 1\alpha _k^{SD}}},$ (8)
where
$\varphi _k^Y = \sqrt {{{\left( {1\alpha _{k - 1}^{SD} - 1\alpha _k^{SD}} \right)}^2} + 4\left\| {{\mathit{\boldsymbol{g}}_k}} \right\|_2^2\left\| {{\mathit{\boldsymbol{s}}_{k - 1}}} \right\|_2^2} ,$
was proposed by Yuan in Ref.[8]. A property of (8) is that if αkY has been taken after αkSD, only one more step αkSD is needed to get the solution to the minimizer of the two-dimensional strictly convex quadratics. If xk is obtained by (3), then sk-1=-αk-1SDgk-1. Thus a variant of (8) was proposed in Ref.[9] as follows:
$\alpha _k^{YY} = \frac{2}{{\varphi _k^{VY} + 1\alpha _{k - 1}^{SD} + 1\alpha _k^{SD}}},$ (9)
where
$\varphi _k^{VY} = \sqrt {{{\left( {1\alpha _{k - 1}^{SD} - 1\alpha _k^{SD}} \right)}^2} + \frac{{4\left\| {{\mathit{\boldsymbol{g}}_k}} \right\|_2^2}}{{{{\left( {\alpha _{k - 1}^{SD}{{\left\| {{\mathit{\boldsymbol{g}}_{k - 1}}} \right\|}_2}} \right)}^2}}}} .$
Based on (9), Dai and Yuan[9] proposed a gradient method whose step length is given by
${\alpha _k} = \left\{ \begin{array}{l}\alpha _k^{SD}\;\;\;\;{\rm{if}}\;{\rm{mod}}\left( {k,4} \right) = 1\;{\rm{or}}\;{\rm{2,}}\\\alpha _k^{VY}\;\;\;\;{\rm{otherwise}}.\end{array} \right.$ (10)
It is not difficult to verify the monotonicity of the method because αkVY≤2αkSD.
The efficiency of the project gradient methods with the above step length selections has been verified by a large amount of numerical experiments on box-constrained quadratic programming[10-13] and Support Vector Machines(SVMs)[5, 14-16]. Inspired by the success of these new step selection rules, we improve project gradient methods for (1).
1 PBB method analysisIn this section, we focus our attention on designing efficient project gradient type methods for solving (1). Here we let P denote the projection operator on Ω:
$\mathit{\boldsymbol{P}}\left( \mathit{\boldsymbol{x}} \right) = \mathop {\arg \min }\limits_{\mathit{\boldsymbol{y}} \in \Omega } {\left\| {\mathit{\boldsymbol{x}} - \mathit{\boldsymbol{y}}} \right\|_2}$ (11)
$ = \mathit{\boldsymbol{x}} - \mathit{\boldsymbol{T}}\left( {\mathit{\boldsymbol{Ax}} - \mathit{\boldsymbol{b}}} \right),$ (12)
where T=AT(AAT)-1. Let d(x) be defined in terms of the gradient g(x) as follows:
$\begin{array}{l}\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right) = \mathit{\boldsymbol{P}}\left( {\mathit{\boldsymbol{x}} - \mathit{\boldsymbol{g}}\left( \mathit{\boldsymbol{x}} \right)} \right) - \mathit{\boldsymbol{x}}\\\;\;\;\;\;\;\;\;\mathit{\boldsymbol{ = }} - \mathit{\boldsymbol{Hg}}\left( \mathit{\boldsymbol{x}} \right),\end{array}$ (13)
where H=I-AT(AAT)-1A.
Given an iterate point xk, the project gradient method for (1) computes the next point in the following form:
${\mathit{\boldsymbol{x}}_{k + 1}} = {\mathit{\boldsymbol{x}}_k} + {\alpha _k}{\mathit{\boldsymbol{d}}_k},$ (14)
where dk=d(xk) and αk>0 is a step length. It is easy to know that sk-i=αk-idk-i and yk-i=αk-iQdk-i, so we obtain another form of (7)
$\alpha _k^M = \frac{{\sum\nolimits_{i = 1}^M {\alpha _{k - i}^2\mathit{\boldsymbol{d}}_{k - i}^{\rm{T}}{\mathit{\boldsymbol{d}}_{k - i}}} }}{{\sum\nolimits_{i = 1}^M {\alpha _{k - i}^2\mathit{\boldsymbol{d}}_{k - i}^{\rm{T}}\mathit{\boldsymbol{Q}}{\mathit{\boldsymbol{d}}_{k - i}}} }}.$ (15)
We refer to the project gradient method with BB-like step length (15) as the projected BB method or PBB method. Obviously, the formula (5) is as a special case of (15) with M=1.
Compared with original project gradient method(PSD), the PBB method is much more effective. In PSD, there is αk=αkSD, where
$\begin{array}{l}\alpha _k^{SD'} = \mathop {\min }\limits_\alpha f\left( {{\mathit{\boldsymbol{x}}_k} + \alpha {\mathit{\boldsymbol{d}}_k}} \right)\\\;\;\;\;\;\;\; = - \frac{{\mathit{\boldsymbol{g}}_k^{\rm{T}}\mathit{\boldsymbol{Q}}{\mathit{\boldsymbol{d}}_k}}}{{\mathit{\boldsymbol{d}}_k^{\rm{T}}\mathit{\boldsymbol{Q}}{\mathit{\boldsymbol{d}}_k}}}.\end{array}$ (16)
In order to illustrate the performance of the PBB method for (1), we have generated 10 random test problems with n=1 000, m=200 and M=2. A, c and the initial point x0 is generated randomly, b=Ax0. Also Q is randomly generated with a range of condition numbers from 102 to 104. The terminal condition is ‖dk < 10-4. The numerical results of the PBB method are presented in Table 1.
Table 1
Table 1 Numerical results of PSD and PBB methods
Problem PBB PSD
cond(Q) iter sec iter sec
10e+3 126 0.86 1145 5.57
10e+2 47 0.39 213 1.30
10e+3 94 0.69 977 4.64
10e+4 334 1.97 5 333 23.37
10e+3 115 0.81 1 019 4.61
10e+3 154 0.94 1 151 5.23
10e+4 351 1.66 4 673 20.40
10e+2 49 0.52 207 1.02
10e+2 56 0.47 219 1.06
10e+4 302 1.55 4 411 19.05
average 162.8 0.98 1 934.8 8.67

Table 1 Numerical results of PSD and PBB methods

In Table 1, cond(Q) is the condition number of Q, iter and sec denote the iteration numbers and the time required respectively.
From Table 1, we can see that the PBB method is much more effective in solving the 10 test problems compared with the PSD method. PBB method with an average of 162.8 iterations and 0.98 s have an obvious advantage than PSD with an average of 1 934.8 and 8.67 s.
2 Two efficient project gradient methodsThough the PBB method performs very well in solving (1), there is no theory to guarantee its global convergence in constrained case[10]. To ensure global convergence of this method, we modify this method by incorporating some sort of line search. It is important that the line search does not degrade the performance of the unmodified PBB method.
Firstly, the sequence {f(xk)} generated by the PBB method is non-monotone, so a non-monotone line search is necessary. Secondly, we take αkM as the step length at each iteration as much as possible. Based on these two considerations, the GLL (Grippo-Lampariello-Lucidi)[17] non-monotone line search is a good choice. We refer to the modified PBB method with the GLL non-monotone line search as MPBB method (see Algorithm 1).
Table
Algorithm 1 MPBB method
Step 1: Let x0 be a feasible point of (1), f0r=∞, f0s=f(x0), k=0, l=0, L, MN+, H=I-AT (AAT)-1A, α0M=1, k0s=k0r=0, i=h=0.
Step 2: dk=-Hgk and terminate if ‖dk=0.
Step 3: If f(xk+αkMdk)≥fhr
????????xk+1=xk+αkSDdk
????else
????????xk+1=xk+αkMdk
????end.
Step 4: If f(xk+1) < f(xkis)
????????ki+1s=k+1, kh+1r=k+1, i=i+1, l=0
????else
????????l=l+1
if l=L
????h=h+1, $f_h^r = \mathop {\max }\limits_{0 \le j \le L} \left\{ {f\left( {{\mathit{\boldsymbol{x}}_{k_h^r + j}}} \right)} \right\}$ , kh+1r=khr+L, l=0
????????end
end.
Step 5: k=k+1, goto Step2.


f(xkis) is the smallest objective function value over the past kis iterations, that is,
$f\left( {{\mathit{\boldsymbol{x}}_{k_i^s}}} \right) = \mathop {\min }\limits_{0 \le j \le k_i^s} f\left( {{\mathit{\boldsymbol{x}}_j}} \right).$
Obviously {f(xkis)} is a strictly monotone decreasing sequence. If no smaller function value is found in L+1 iterations, the reference function value fhr will be updated by the maximum function value in most recent L+1 iterations.
Next, we will introduce a kind of monotone project gradient method with a new step length similar to (10). Now we give a general form of the monotone project gradient method for (1) (see Algorithm 2).
Table
Algorithm 2 Monotone algorithm
Step 1: Let x0 be a feasible point of (1), k=0, H=I-AT (AAT)-1A.
Step 2: dk=-Hgk and terminate if ‖dk=0.
Step 3: xk+1= xk+αkdk, where αk=ρkαkSD, ρk∈[0, 2].
Step 4: k=k+1; goto Step2.
Step 5: k=k+1, goto Step2.


The following step length similar to αkVY
$\alpha _k^{VY'} = \frac{2}{{\varphi _k^{VY'} + 1\alpha _{k - 1}^{SD'} + 1\alpha _k^{SD'}}}$ (17)
where
$\varphi _k^{VY'} = \sqrt {{{\left( {1\alpha _{k - 1}^{SD'} - 1\alpha _k^{SD'}} \right)}^2} + \frac{{4\left\| {{\mathit{\boldsymbol{g}}_k}} \right\|_2^2}}{{{{\left( {\alpha _{k - 1}^{SD'}{{\left\| {{\mathit{\boldsymbol{g}}_{k - 1}}} \right\|}_2}} \right)}^2}}}} $
is given by replacing αkSD with αkSD in (9). Following the formula (10), the step length
$\alpha _k^{SY} = \left\{ \begin{array}{l}\alpha _k^{SD'}\;\;\;\;{\rm{if}}\;{\rm{mod}}\left( {k,4} \right) = 1\;{\rm{or}}\;{\rm{2}}\\\alpha _k^{VY'}\;\;\;\;{\rm{otherwise}}\end{array} \right.$ (18)
is given. We refer to the project gradient method with (18) as the PSY method. Obviously, the methods PSD and PSY are special cases of the Algorithm 2.
Here we need to point out that αkVY may not have the similar property of (8) and (9) in two-dimensions for (1), but the numerical results will show that the PSY method is comparable to the PBB method.
3 Global convergence analysisIn this section, we analyze the global convergence of Algorithm 1 and Algorithm 2. In what follows, we denote the set of solutions to (1) by
${{\bf{X}}^ * } = \left\{ {{\mathit{\boldsymbol{x}}^ * } \in \Omega \left| {f\left( x \right) \ge f\left( {{x^ * }} \right),} \right.{\rm{for}}\;{\rm{all}}\;x \in \Omega } \right\},$
and the optimal value of (1) by f*.
Theorem 3.1????If Algorithm 1 does not terminate in a finite number of iterations, there must exist a strictly monotone decreasing infinite subsequence in {f(xk)}.
Proof????Let p be the length of the {f(xkis)}. If p=+∞, {f(xkis)} is the subsequence satisfying the requirement.
If p < +∞ is finite, there exists a const h0 which satisfies krh0=kps. Without loss of generality, let h0=0. Since no smaller function value is found after iteration kps, the reference function fhr will be updated every L+1 iterations and the length of {fhr} is infinite. It is easy to see that
$f_{{h_0} + q}^r = f_q^r = \mathop {\max }\limits_{0 \le j \le L} \left\{ {f\left( {{\mathit{\boldsymbol{x}}_{k_q^r + j}}} \right)} \right\},q = 1,2, \cdots $ (19)
On the other hand,
$f_q^r > f\left( {{\mathit{\boldsymbol{x}}_{k_{q + 1}^r + j}}} \right),j = 1,2, \cdots ,L.$ (20)
It follows from kq+1r=kqr+L, (19) and (20) that
$f_q^r \ge \mathop {\max }\limits_{0 \le j \le L} \left\{ {f\left( {{\mathit{\boldsymbol{x}}_{k_{q + 1}^r + j}}} \right)} \right\}$ (21)
$ = f_{q + 1}^r.$ (22)
Let q=q+1 in (20), then fq+1r>f(xkrq+2+j), j=1, 2, …, L. Combining (22), we conclude that
$f_q^r > f\left( {{\mathit{\boldsymbol{x}}_{k_{q + 2}^r + j}}} \right),j = 1,2, \cdots ,L.$ (23)
Let j=L in (20), then
$\begin{array}{l}f_q^r > f\left( {{\mathit{\boldsymbol{x}}_{k_{q + 1}^r + L}}} \right)\\\;\;\;\;\; = f\left( {{\mathit{\boldsymbol{x}}_{k_{q + 2}^r}}} \right)\end{array}$ (24)
It follows from (23) and (24) that
$f_q^r > \mathop {\max }\limits_{0 \le j \le L} \left\{ {f\left( {{\mathit{\boldsymbol{x}}_{k_{q + 2}^r + j}}} \right)} \right\} = f_{q + 2}^r$
which implies that either even or odd subsequence of {fqr} is strictly monotone decreasing. In conclusion, the theorem is true.
We give some properties of P(x) and d(x) which will be used in the following theorems, where P(x) and d(x) are defined as (11) and (13) respectively.
Proposition 3.1????(Properties of P(x) and d(x))
P1.?yRn and x∈Ω, (P(y)-y)T(x-P(y))≥0.
P2.?x∈Ω, g(x)Td(x)≤-d(x)Td(x).
P3.?x**∈Ω, d(x**)=0 if and only if g(x**)T(x-x**)≥0 for all x∈Ω.
P4.?xRn and x**X**={y∈Ω|d(y)=0}, we have
${\left\| {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right\|_2} \le \frac{{1 + {\lambda _n}}}{{{\lambda _1}}}{\left\| {\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)} \right\|_2}.$
Proof????Firstly, P1 is the optimal condition of (11)(see Ref.[11]).
From P1, we can easily obtain P2 by replacing y with x-g(x) in P1.
P3 is proved as follows.
From the definition of d(x), we know that if d(x**)=0, x**=P(x**-g(x**)). If y is replaced by x**-g(x**) in P1, P1 yields
$\mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\left( {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right) \ge 0\forall \mathit{\boldsymbol{x}} \in \Omega .$ (25)
Conversely, due to P(x**-g(x**))∈Ω, we have
$\mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\mathit{\boldsymbol{d}}\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right) \ge 0.$
From above inequality and P2, we obtain
$0 \le - \mathit{\boldsymbol{d}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\mathit{\boldsymbol{d}}\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right) \le 0$
which means d(x**)=0.
Finally, we consider P4. In P1, y is replaced with x-g(x) and x is replaced with x**, then we have inequality as follows,
$\begin{array}{l}{\left( {\mathit{\boldsymbol{P}}\left( {\mathit{\boldsymbol{x}} - \mathit{\boldsymbol{g}}\left( \mathit{\boldsymbol{x}} \right)} \right) - \mathit{\boldsymbol{x}} + \mathit{\boldsymbol{g}}\left( \mathit{\boldsymbol{x}} \right)} \right)^{\rm{T}}}\\\left( {{\mathit{\boldsymbol{x}}^{ * * }}\mathit{\boldsymbol{P}}\left( {\mathit{\boldsymbol{x}} - \mathit{\boldsymbol{g}}\left( \mathit{\boldsymbol{x}} \right)} \right)} \right) \ge 0.\end{array}$ (26)
In fact, by the definition of d(x) as (13), we know that (26) is equivalent to
$\begin{array}{*{20}{c}}{{{\left( {\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right) + \mathit{\boldsymbol{g}}\left( \mathit{\boldsymbol{x}} \right)} \right)}^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}} - \mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)} \right) + }\\{\mathit{\boldsymbol{g}}{{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)}^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right) \ge \mathit{\boldsymbol{g}}{{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)}^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right).}\end{array}$ (27)
Rearranging (27), we have
$\begin{array}{l}\mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right) - \mathit{\boldsymbol{g}}{\left( \mathit{\boldsymbol{x}} \right)^{\rm{T}}}\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)\\\;\; \ge \mathit{\boldsymbol{d}}{\left( \mathit{\boldsymbol{x}} \right)^{\rm{T}}}\left( {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right) + {\left\| {\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)} \right\|_2} + \\\;\;{\left( {\mathit{\boldsymbol{g}}\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right) - \mathit{\boldsymbol{g}}\left( \mathit{\boldsymbol{x}} \right)} \right)^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right)\\\;\; \ge \mathit{\boldsymbol{d}}{\left( \mathit{\boldsymbol{x}} \right)^{\rm{T}}}\left( {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right) + {\left\| {\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)} \right\|_2} + \\\;\;{\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right)^{\rm{T}}}\mathit{\boldsymbol{Q}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right)\\\;\; \ge \mathit{\boldsymbol{d}}{\left( \mathit{\boldsymbol{x}} \right)^{\rm{T}}}\left( {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right) + {\lambda _1}\left\| {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right\|_2^2.\end{array}$ (28)
On the other hand,
$\begin{array}{l}\mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right) - \mathit{\boldsymbol{g}}{\left( \mathit{\boldsymbol{x}} \right)^{\rm{T}}}\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)\\\;\;\; = \mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right) + {\left( {\mathit{\boldsymbol{g}}\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right) - \mathit{\boldsymbol{g}}\left( \mathit{\boldsymbol{x}} \right)} \right)^{\rm{T}}}\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right) - \\\;\;\;\mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)\\\;\;\; = \mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}} - \mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)} \right) + {\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right)^{\rm{T}}}\mathit{\boldsymbol{Qd}}\left( \mathit{\boldsymbol{x}} \right)\\\;\;\; = \mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{P}}\left( {\mathit{\boldsymbol{x}} - \mathit{\boldsymbol{g}}\left( \mathit{\boldsymbol{x}} \right)} \right)} \right) + \\\;\;\;\;\;\;{\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right)^{\rm{T}}}\mathit{\boldsymbol{Qd}}\left( \mathit{\boldsymbol{x}} \right).\end{array}$ (29)
Since P(x-g(x))∈Ω and from P3, then we have
$\begin{array}{l}\mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^{ * * }}} \right)^{\rm{T}}}\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right) - \mathit{\boldsymbol{g}}{\left( \mathit{\boldsymbol{x}} \right)^{\rm{T}}}\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)\\\;\;\; \le {\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right)^{\rm{T}}}\mathit{\boldsymbol{Qd}}\left( \mathit{\boldsymbol{x}} \right).\end{array}$ (30)
Combining (28) and (30), we obtain that
$\begin{array}{l}\mathit{\boldsymbol{d}}{\left( \mathit{\boldsymbol{x}} \right)^{\rm{T}}}\left( {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right) + {\lambda _1}\left\| {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right\|_2^2\\\;\; \le {\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right)^{\rm{T}}}\mathit{\boldsymbol{Qd}}\left( \mathit{\boldsymbol{x}} \right).\end{array}$ (31)
Rearranging (31), we have
$\begin{array}{l}{\left\| {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right\|_2} \le \frac{{{{\left( {{\mathit{\boldsymbol{x}}^{ * * }} - \mathit{\boldsymbol{x}}} \right)}^{\rm{T}}}\left( {\mathit{\boldsymbol{I}} + \mathit{\boldsymbol{Q}}} \right)\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)}}{{{\lambda _1}{{\left\| {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^{ * * }}} \right\|}_2}}}\\\;\;\;\;\;\;\;\;\;\;\;\;\;\; \le \frac{{1 + {\lambda _n}}}{{{\lambda _1}}}{\left\| {\mathit{\boldsymbol{d}}\left( \mathit{\boldsymbol{x}} \right)} \right\|_2}.\end{array}$ (32)
This completes the proof.
Now, we give the following theorem about the equivalent optimal conditions of (1).
Theorem 3.2????Suppose x* is a feasible point of (1), the following results are equivalent:
E1.x*X*.
E2.d(x*)Tg(x*)=0.
E3.d(x*)TQd(x*)=0.
E4.d(x*)=0.
Proof E1?E4:
From P3, d(x*)=0 is equivalent to
$\mathit{\boldsymbol{g}}{\left( {{\mathit{\boldsymbol{x}}^ * }} \right)^{\rm{T}}}\left( {\mathit{\boldsymbol{x}} - {\mathit{\boldsymbol{x}}^ * }} \right) \ge 0,\forall \mathit{\boldsymbol{x}} \in \Omega .$
The above inequality implies that x* is the solution of the following linear programming
$\begin{array}{*{20}{c}}{\min \mathit{\boldsymbol{g}}{{\left( {{\mathit{\boldsymbol{x}}^ * }} \right)}^{\rm{T}}}\mathit{\boldsymbol{x}}}\\{{\rm{s}}.\;{\rm{t}}.\;\mathit{\boldsymbol{x}} \in \Omega .}\end{array}$ (33)
Obviously, (1) and (37) have the same KKT system, so x*X*.
Since E1?E4, d(x*)=0. Then proposition E1?E2 and E1?E3 can be easily obtained. From E2 and P2, we have that d(x*)=0, which means E1 is established.
E3?E1:
Assume E1 is not established, that is, x*?X*. From P2 and E2, we have d(x*)Tg(x*) < 0. As f(x) is quadratic and d(x*)TQd(x*)=0, we have f(x*+αd(x*))=f(x*)+αg(x*)Td(x*). Let α→+∞, then f(x*+αd(x*))→-∞ which contradicts that (1) is bounded below. Thus, x* belongs to X*.
According to Theorem 3.2 and P4, we can see that X*=X**. Then, for all feasible points x?X*,
$\mathit{\boldsymbol{d}}{\left( \mathit{\boldsymbol{x}} \right)^{\rm{T}}}\mathit{\boldsymbol{Qd}}\left( \mathit{\boldsymbol{x}} \right) > 0.$
Thus, the formula (15) and (18) are well defined unless xk is a optimal point of (1).
Finally, we present the global convergence of Algorithm 2.
Theorem 3.3????If Algorithm 2 does not terminate in a finite number of iterations and the sequence {ρk} has an accumulation point $\tilde \rho \in \left({0, 2} \right)$ , then the sequence {f(xk)} generated by Algorithm 2 converges to f*.
Proof????Observe that for all k,
${\phi _k}\left( \rho \right) = f\left( {{\mathit{\boldsymbol{x}}_k} + \rho \alpha _k^{SD'}{\mathit{\boldsymbol{d}}_k}} \right)$
is a quadratic convex function of the variable ρ. From the definition of αkSD, we get that ?k(ρ) attains global minimum when ρ=1. Hence by symmetry of the quadratic convex function ?k(ρ), ?k(0)=?k(2) and for any ρ∈[0, 2], ?k(ρ)≤?k(0). Therefore
${\phi _k}\left( {{\rho _k}} \right) = f\left( {{\mathit{\boldsymbol{x}}_{k + 1}}} \right) \le f\left( {{\mathit{\boldsymbol{x}}_k}} \right),$
for all k. The above inequality implies that {f(xk)} is a monotonically decreasing sequence. Furthermore, f(xk) is bounded below, so {f(xk)} a convergent sequence,
$\mathop {\lim }\limits_{k \to + \infty } \left( {f\left( {{\mathit{\boldsymbol{x}}_k}} \right) - f\left( {{\mathit{\boldsymbol{x}}_{k + 1}}} \right)} \right) = 0.$ (34)
Because ${\tilde \rho }$ is an accumulation point of {ρk} and $\tilde \rho \in \left({0, 2} \right)$ , there exists a const β∈(0, 1) and subsequence ρkj∈[β, 2-β]. Using the properties of ?k we have
$\begin{array}{l}f\left( {{\mathit{\boldsymbol{x}}_{{k_j}}}} \right) - f\left( {{\mathit{\boldsymbol{x}}_{{k_{j + 1}}}}} \right) = {\phi _{{k_j}}}\left( 0 \right) - {\phi _{{k_j}}}\left( {{\rho _{{k_j}}}} \right)\\\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \ge {\phi _{{k_j}}}\left( 0 \right) - {\phi _{{k_j}}}\left( \beta \right)\\\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; = - \beta {{\phi '}_{{k_j}}}\left( 0 \right) - \frac{{{\beta ^2}}}{2}{{\phi ''}_{{k_j}}}\left( 0 \right)\\\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; = \beta \alpha _{{k_j}}^{SD'}\mathit{\boldsymbol{g}}_{{k_j}}^{\rm{T}}{\mathit{\boldsymbol{d}}_{{k_j}}} - \frac{{{\beta ^2}}}{2}\alpha _{{k_j}}^{SD'2}\mathit{\boldsymbol{d}}_{{k_j}}^{\rm{T}}\mathit{\boldsymbol{Q}}{\mathit{\boldsymbol{d}}_{{k_j}}}\\\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; = \frac{{\beta \left( {2 - \beta } \right)}}{2}\frac{{{{\left( {\mathit{\boldsymbol{g}}_{{k_j}}^{\rm{T}}{\mathit{\boldsymbol{d}}_{{k_j}}}} \right)}^2}}}{{\mathit{\boldsymbol{d}}_{{k_j}}^{\rm{T}}\mathit{\boldsymbol{Q}}{\mathit{\boldsymbol{d}}_{{k_j}}}}}\end{array}$
Since gTkjdkj≤-dTkj≤0, then |gTkjdkj|≥|dTkjdkj|. Thus, we have that
$\begin{array}{l}f\left( {{\mathit{\boldsymbol{x}}_{{k_j}}}} \right) - f\left( {{\mathit{\boldsymbol{x}}_{{k_j} + 1}}} \right) \ge \frac{{\beta \left( {2 - \beta } \right)}}{2}\frac{{{{\left( {\mathit{\boldsymbol{d}}_{{k_j}}^{\rm{T}}{\mathit{\boldsymbol{d}}_{{k_j}}}} \right)}^2}}}{{\mathit{\boldsymbol{d}}_{{k_j}}^{\rm{T}}\mathit{\boldsymbol{Q}}{\mathit{\boldsymbol{d}}_{{k_j}}}}}\\\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; \ge \frac{{\beta \left( {2 - \beta } \right)\mathit{\boldsymbol{d}}_{{k_j}}^{\rm{T}}{\mathit{\boldsymbol{d}}_{{k_j}}}}}{{2{\lambda _{\max }}\left( \mathit{\boldsymbol{Q}} \right)}}.\end{array}$ (35)
Let j→+∞ in the (35), then
$\begin{array}{*{20}{c}}{0 = \mathop {\lim }\limits_{j \to + \infty } \left( {f\left( {{\mathit{\boldsymbol{x}}_{{k_j}}}} \right) - f\left( {{\mathit{\boldsymbol{x}}_{{k_{j + 1}}}}} \right)} \right)}\\{ \ge \mathop {\lim }\limits_{j \to + \infty } \frac{{\beta \left( {2 - \beta } \right)\mathit{\boldsymbol{d}}_{{k_j}}^{\rm{T}}{\mathit{\boldsymbol{d}}_{{k_j}}}}}{{2{\lambda _{\max }}\left( \mathit{\boldsymbol{Q}} \right)}} \ge 0.}\end{array}$
From the above inequality, we conclude that dkj0, j→+∞.
It follows from P4 that xkj converges to some optimal point x*X*. Since f(x) is continuous, we have
$f\left( {{\mathit{\boldsymbol{x}}_{{k_j}}}} \right) \to {f^ * }.$
As {f(xk)} is decreasing and bounded from below, the whole sequence {f(xk)} converges to f*.
4 Numerical experimentsIn this section, we report the numerical results of the project gradient methods proposed in this paper. The test problems are designed based on nine parameters n, m, ncond, lx, ux, lA, uA, lc, uc. In particular, we denote Q=PΛPT, where
$\mathit{\boldsymbol{P}} = \mathop {\mathit{\Pi }}\limits_{i = 1}^3 \left( {\mathit{\boldsymbol{I}} - 2{\mathit{\boldsymbol{\omega }}_i}\mathit{\boldsymbol{\omega }}_i^{\rm{T}}} \right)$
and ω1, ω2, ω3 are random unit vectors, and Λ is diagonal matrix whose i-th component is defined by
$\log {\lambda _i} = \frac{{i - 1}}{{n - 1}}n{\rm{cond}},i = 1, \cdots ,n.$
The initial point x0, matrix A and the linear term c are generated by Matlab function rand with entries in [lx, ux], [lA, uA], [lc, uc] respectively. b is built as b=Ax0. In our tests, we have fixed (lx, ux, lA, uA, lc, uc)=(-5, 5, -10, 10, -10, 10). m, n, and ncond are positive integers chosen randomly in the interval [50, 800], [1 000, 2 000] and [2, 6] by the function randint respectively. As the stopping criterion we use
${\left\| {{\mathit{\boldsymbol{d}}_k}} \right\|_\infty } \le {10^{ - 4}}{\left\| {{\mathit{\boldsymbol{d}}_0}} \right\|_\infty }.$
All tests are conducted on a Windows XP professional computer (Pentium R, 2.79 GHZ) with Matlab 7.10.
Firstly, we generate 100 random test problems to test the PBB method and the MPBB method with M=1, 2, …, 15. The numerical results are reported in Table 2, where Aiter and Asec denote the average number of iterations and time required with regard to these 100 test problems respectively.
Table 2
Table 2 Testing the PBB and MPBB methods as M increases
M PBB MPBB
Aiter Asec Aiter Asec
1 309.96 3.85 176.37 2.58
2 282.61 3.61 138.73 2.14
3 277.23 3.57 145.46 2.25
4 268.54 3.49 154.15 2.30
5 277.43 3.57 149.62 2.27
6 268.04 3.47 154.87 2.29
7 277.60 3.56 155.44 2.23
8 282.33 3.61 156.89 2.33
9 278.53 3.57 156.09 2.33
10 282.42 3.58 157.24 2.32
11 285.95 3.64 166.88 2.44
12 291.98 3.72 170.28 2.49
13 297.73 3.76 172.13 2.45
14 303.85 3.80 176.87 2.56
15 295.85 3.73 177.33 2.52

Table 2 Testing the PBB and MPBB methods as M increases

From Table 2, we see that the PBB method performs worse as M(M>6) increases with rare exception, so does the MPBB method (M>2). On the whole, the iterations and time of these two methods have a tendency to ascend with the growth of M. At the same time, we suggest M=6 for the PBB method and M=2 for the MPBB method. Also we conclude that the MPBB method performs much better than PBB method.
Table 3 lists the numbers of iterations and time required by the PSD, PBB (M=6), MPBB (M=2) and PSY methods for 15 random problems. From this table, we can clearly see that the PSD method is significantly slower than the other three methods. Although, the PSY is slightly worse than the MPBB method, the PSY method performs better than PBB method. What's more, the MPBB method proposed by us is not only to guarantee the convergence of PBB method but also to perform much better than the latter one.
Table 3
Table 3 Comparison of numerical results for random test problems among the methods
Problem PSD PBB MPBB PSY
ncond m n Iter Sec Iter Sec Iter Sec Iter Sec
2 661 1 906 145 3.73 47 2.04 33 2.15 43 2.58
5 793 1 833 5 553 84.23 402 7.84 201 5.41 232 7.73
2 258 1 805 251 4.11 59 1.31 52 1.28 54 1.88
3 242 1 288 1 195 8.88 118 1.09 121 1.22 119 1.58
2 383 1 015 129 0.94 41 0.48 40 0.59 42 0.70
4 676 1 334 1 035 9.25 137 2.08 93 1.78 126 2.70
2 716 1 686 117 3.03 41 2.01 32 2.05 46 2.30
2 226 1 802 241 3.63 61 1.48 52 1.30 62 1.92
4 418 1 388 3 107 25.16 207 2.29 153 1.84 179 2.75
6 322 1 107 22 329 123.30 810 4.77 262 1.64 418 3.75
3 470 1 914 1 005 15.17 111 2.52 118 2.92 147 4.42
5 78 1 910 25 911 469 548 9.81 192 2.95 208 4.72
6 489 1 906 26 883 486.67 1 090 21.93 253 5.97 227 7.25
2 638 1 128 69 1.39 40 1.18 26 1.14 25 1.16
5 689 1 777 6 569 102.78 357 7.60 194 5.28 255 8.06
5 399 1 511 14 087 161.25 519 5.93 237 3.37 300 5.92
4 766 1 060 225 2.38 74 1.58 49 1.43 50 1.54
average 6 403 88.53 274.23 4.47 124 2.49 149 3.59

Table 3 Comparison of numerical results for random test problems among the methods

5 ConclusionsIn this paper, we proposed two methods, MPBB and PSY, for quadratic programming with linear equality constraints (QPLE) based on the efficient step length selections and have established their convergence under mild assumptions. Our numerical results demonstrate that the new project gradient methods with these step selection rules have superiority over the methods with classical step length.
References
[1] Barzilai J, Borwein J M. Two-point step size gradient methods[J].IMA Journal of Numerical Analysis, 1988, 8(1):141–148.DOI:10.1093/imanum/8.1.141
[2] Zhou B, Gao L, Dai Y H. Gradient methods with adaptive step-sizes[J].Computational Optimization and Applications, 2006, 35(1):69–86.DOI:10.1007/s10589-006-6446-0
[3] Raydan M. On the Barzilai and Borwein choice of steplength for the gradient method[J].IMA Journal of Numerical Analysis, 1993, 13(3):321–326.DOI:10.1093/imanum/13.3.321
[4] Dai Y H, Liao L Z. R-linear convergence of the Barzilai and Borwein gradient method[J].IMA Journal of Numerical Analysis, 2002, 22(1):1–10.
[5] Dai Y H, Fletcher R. New algorithms for singly linearly constrained quadratic programs subject to lower and upper bounds[J].Mathematical Programming, 2006, 106(3):403–421.DOI:10.1007/s10107-005-0595-2
[6] Raydan M. The Barzilai and Borwein gradient method for the large scale unconstrained minimization problem[J].SIAM Journal on Optimization, 1997, 7(1):26–33.DOI:10.1137/S1052623494266365
[7] Yuan Y. Gradient methods for large scale convex quadratic functions[M].Berlin: Springer Heidelberg, 2010: 141-155.
[8] Yuan Y. A new stepsize for the steepest descent method[J].Journal of Computational Mathematics, 2006:149–156.
[9] Dai Y, Yuan Y X. Analysis of monotone gradient methods[J].Journal of Industrial and Management Optimization, 2005, 1(2):181.DOI:10.3934/jimo
[10] Dai Y H, Fletcher R. Projected Barzilai-Borwein methods for large-scale box-constrained quadratic programming[J].Numerische Mathematik, 2005, 100(1):21–47.DOI:10.1007/s00211-004-0569-y
[11] Hager W W, Zhang H. A new active set algorithm for box constrained optimization[J].SIAM Journal on Optimization, 2006, 17(2):526–557.DOI:10.1137/050635225
[12] Zhou B, Gao L, Dai Y. Monotone projected gradient methods for largescale box-constrained quadratic programming[J].Science in China Series A, 2006, 49(5):688–702.DOI:10.1007/s11425-006-0688-2
[13] De Asmundis R, di Serafino D, Riccio F, et al. On spectral properties of steepest descent methods[J].IMA Journal of Numerical Analysis, 2013:drs056.
[14] Zanni L, Serafini T, Zanghirati G. Parallel software for training large scale support vector machines on multiprocessor systems[J].Journal of Machine Learning Research, 2006, 7(7):1467–1492.
[15] Serafini T, Zanghirati G, Zanni L. Parallel decomposition approaches for training support vector machines[J].Advances in Parallel Computing, 2004, 13:259–266.DOI:10.1016/S0927-5452(04)80035-2
[16] Zanghirati G, Zanni L. A parallel solver for large quadratic programs in training support vector machines[J].Parallel computing, 2003, 29(4):535–551.DOI:10.1016/S0167-8191(03)00021-8
[17] Grippo L, Lampariello F, Lucidi S. A nonmonotone line search technique for Newton's method[J].SIAM Journal on Numerical Analysis, 1986, 23(4):707–716.DOI:10.1137/0723046


相关话题/优化 规划 中国科学院大学 数学 科学学院

  • 领限时大额优惠券,享本站正版考研考试资料!
    大额优惠券
    优惠券领取后72小时内有效,10万种最新考研考试考证类电子打印资料任你选。涵盖全国500余所院校考研专业课、200多种职业资格考试、1100多种经典教材,产品类型包含电子书、题库、全套资料以及视频,无论您是考研复习、考证刷题,还是考前冲刺等,不同类型的产品可满足您学习上的不同需求。 ...
    本站小编 Free壹佰分学习网 2022-09-19
  • 不确定传输速率下无线资源调度问题的鲁棒优化模型
    田雷霞,杨文国,高随祥,姜志鹏中国科学院大学数学科学学院,北京100049;中国科学院大数据挖掘与知识管理重点实验室,北京1001902017年02月20日收稿;2017年03月27日收修改稿基金项目:国家自然科学基金(11571015,11331012),中国科学院战略性先导科技专项(XDA060 ...
    本站小编 Free考研考试 2021-12-25
  • 超密集网络中基于多连接的用户归属和功率控制联合优化
    张剑,邱玲,陈正中国科学技术大学中国科学院无线光电通信重点实验室,合肥2300272016年11月25日收稿;2017年03月20日收修改稿基金项目:国家自然科学基金(61672484)资助通信作者:邱玲,E-mail:lqiu@ustc.edu.cn摘要:超密集网络是第五代移动通信系统提升容量的关 ...
    本站小编 Free考研考试 2021-12-25
  • WFST解码器词图生成算法中的非活跃节点检测与内存优化
    丁佳伟1,刘加1,张卫强1,冯运波2,刘利军2,于乐21.清华大学电子工程系,北京100084;2.中国移动通信信息安全管理与运行中心,北京1000532017年12月22日收稿;2018年3月2日收修改稿基金项目:国家自然科学基金(U1836219)资助通信作者:刘加,E-mail:liuj@ts ...
    本站小编 Free考研考试 2021-12-25
  • 武汉东湖新技术开发区产业创新与产业结构优化升级的耦合研究
    陈妤凡1,2,王开泳11.中国科学院地理科学与资源研究所/中国科学院区域可持续发展分析与模拟重点实验室,北京100101;2.中国科学院大学,北京1000492017年5月31日收稿;2017年10月24日收修改稿基金项目:国家自然科学基金(41471126,41371178)资助通信作者:王开泳, ...
    本站小编 Free考研考试 2021-12-25
  • 基于功率谱估计的航磁补偿优化处理方法
    吴佩霖1,2,张群英1,陈路昭1,2,费春娇1,2,朱万华1,方广有11.中国科学院电子学研究所电磁辐射与探测技术重点实验室,北京100190;2.中国科学院大学,北京1000492017年4月10日收稿;2017年7月13日收修改稿基金项目:国家自然科学基金(41374186)和863计划项目(2 ...
    本站小编 Free考研考试 2021-12-25
  • 基于PSO的多星座GNSS垂直保护级优化方法*
    接收机自主完好性监测(RAIM)算法可以在单星座、单故障下为航空用户提供非精密进近阶段的完好性监测。随着多星座组合导航系统的应用,使得可见卫星数增加,单星座、单故障假设不再成立[1]。GEAS(GNSSEvolutionaryArchitectureStudy)报告提出的高级接收机自主完好性监测(A ...
    本站小编 Free考研考试 2021-12-25
  • 基于可靠性的CFRP螺栓连接优化设计方法*
    先进的碳纤维增强复合材料(CFRP)具有高比强度/比刚度、优异的疲劳性能、可裁剪设计等优点,已被广泛应用于航空航天结构。目前,先进复合材料的应用水平和用量是衡量飞机先进性和市场竞争力的标志。为与美国波音、欧洲空客形成三足鼎立的竞争格局,中国下一代远程宽体客机CR929的复合材料用量将由目前C919的 ...
    本站小编 Free考研考试 2021-12-25
  • 固体火箭冲压发动机设计点性能优化分析*
    冲压发动机通过气流减速增压的工作方式,省去了涡喷涡扇发动机的转动部件,结构复杂度降低,同时有效利用空气中的氧气作为氧化剂,相比火箭发动机比冲提高,在超声速飞行器中得到了广泛应用[1-3]。固体火箭冲压发动机通过燃气发生器产生贫氧燃气,具有推进剂供应系统简单、推进剂密度比冲高、结构布局紧凑等优势,在超 ...
    本站小编 Free考研考试 2021-12-25
  • 高碳醇/膨胀石墨复合相变热沉多目标优化*
    随着电子芯片和集成电路的高速发展,在性能不断提升的同时带来发热量增大的问题。为保证芯片的正常工作,必须设计合适的散热装置,控制芯片在合理的温度范围内工作。传统的风冷、液冷等主动或被动式芯片散热器均需要冷边提供冷量,通常采用冷空气作为冷边介质。对于一些特殊工况,如封闭空间或高温环境,传统散热器无法满足 ...
    本站小编 Free考研考试 2021-12-25
  • 优化Landweber迭代快速电磁层析成像图像重建算法*
    电学层析成像(ElectricalTomography,ET)技术作为一种高新检测技术,包括电阻层析成像(ElectricalResistanceTomography,ERT)[1-2]、电磁层析成像(ElectromagneticTomography,EMT)[3-4]、电容层析成像(Electr ...
    本站小编 Free考研考试 2021-12-25