diff --git a/多体+耦合求解器/叶片上坐标系变换.canvas b/多体+耦合求解器/叶片上坐标系变换.canvas
new file mode 100644
index 0000000..7f66e77
--- /dev/null
+++ b/多体+耦合求解器/叶片上坐标系变换.canvas
@@ -0,0 +1,27 @@
+{
+ "nodes":[
+ {"id":"6408e37f6227f511","type":"text","text":"coned坐标系\ni1_k\ni2_k\ni3_k","x":-160,"y":-200,"width":180,"height":140},
+ {"id":"c48d5ceaaf24dc21","type":"text","text":"pitch坐标系\nj1_k\nj2_k\nj3_k","x":180,"y":-200,"width":180,"height":140},
+ {"id":"05315d16dc63ca42","type":"text","text":"叶片刚性叶素节点坐标系\nlj1_k\nlj2_k\nlj3_k","x":520,"y":-200,"width":180,"height":140},
+ {"id":"89453e1b65c9e3ab","type":"text","text":"$$\nR_{lj \\to j} = \n\\begin{bmatrix}\n\\cos(\\text{twist}) & -\\sin(\\text{twist}) & 0 \\\\\n\\sin(\\text{twist}) & \\cos(\\text{twist}) & 0 \\\\\n0 & 0 & 1\n\\end{bmatrix}\n$$","x":270,"y":-380,"width":360,"height":120},
+ {"id":"8502d848516e0912","type":"text","text":"叶片柔性叶素节点坐标系\nn1_k\nn2_k\nn3_k","x":840,"y":-200,"width":180,"height":140},
+ {"id":"e4eb05643bb263da","type":"text","text":"$$\nR_{i \\to g_1'} = \n\\begin{bmatrix}\n\\cos(\\text{pre\\_cone}) & 0 & \\sin(\\text{pre\\_cone}) \\\\\n0 & 1 & 0 \\\\\n-\\sin(\\text{pre\\_cone}) & 0 & \\cos(\\text{pre\\_cone})\n\\end{bmatrix}\n$$","x":-390,"y":-500,"width":375,"height":120},
+ {"id":"8948f4893c073dee","type":"text","text":"叶根坐标系\ng1_prime_k、\ng2_prime_k、\ng3_prime_k","x":-480,"y":-200,"width":180,"height":140},
+ {"id":"19dc1f9aafe945e7","type":"text","text":"叶素节点orientation","x":1180,"y":-200,"width":180,"height":140},
+ {"id":"a3526395fd1e1879","type":"text","text":"力和力矩使用","x":1180,"y":10,"width":180,"height":140},
+ {"id":"5e8cd0e90b2ae11f","type":"text","text":"载荷加在此坐标系\n每个节点\nMx, My, Mxy, Mz\nFx, Fy, Fxy, , Fz","x":-495,"y":70,"width":210,"height":140},
+ {"id":"90d374e8834332b4","x":695,"y":165,"width":180,"height":90,"type":"text","text":"力矩变换从原点到叶素点需要平行移轴"},
+ {"id":"24c2775deb74b1f1","type":"text","text":"smll_rot_trans","x":660,"y":10,"width":250,"height":120},
+ {"id":"a7e732bf559a1c24","type":"text","text":"$$\nR_{j \\to i} = \n\\begin{bmatrix}\n\\cos(\\text{pitch}) & -\\sin(\\text{pitch}) & 0 \\\\\n\\sin(\\text{pitch}) & \\cos(\\text{pitch}) & 0 \\\\\n0 & 0 & 1\n\\end{bmatrix}\n$$","x":-70,"y":10,"width":360,"height":120}
+ ],
+ "edges":[
+ {"id":"625bb32862e524f0","fromNode":"6408e37f6227f511","fromSide":"right","toNode":"c48d5ceaaf24dc21","toSide":"left"},
+ {"id":"b94d6a9951731170","fromNode":"8948f4893c073dee","fromSide":"right","toNode":"6408e37f6227f511","toSide":"left"},
+ {"id":"6e88f9cb172ae6fa","fromNode":"c48d5ceaaf24dc21","fromSide":"right","toNode":"05315d16dc63ca42","toSide":"left"},
+ {"id":"c832651a1a42f7af","fromNode":"05315d16dc63ca42","fromSide":"right","toNode":"8502d848516e0912","toSide":"left"},
+ {"id":"b118c17041fd0b89","fromNode":"8502d848516e0912","fromSide":"right","toNode":"19dc1f9aafe945e7","toSide":"left"},
+ {"id":"6f47755fb87c14ef","fromNode":"6408e37f6227f511","fromSide":"top","toNode":"8948f4893c073dee","toSide":"top","label":"变形量"},
+ {"id":"efb288920b3d370a","fromNode":"5e8cd0e90b2ae11f","fromSide":"top","toNode":"8948f4893c073dee","toSide":"bottom","label":"当前版本暂不使用Mxy、Fxy"},
+ {"id":"52f4e8642aa2f565","fromNode":"a3526395fd1e1879","fromSide":"top","toNode":"19dc1f9aafe945e7","toSide":"bottom"}
+ ]
+}
\ No newline at end of file
diff --git a/工作OKRs/25.6-8 OKR.canvas b/工作OKRs/25.6-8 OKR.canvas
index 180cf60..a0443c5 100644
--- a/工作OKRs/25.6-8 OKR.canvas
+++ b/工作OKRs/25.6-8 OKR.canvas
@@ -5,7 +5,7 @@
{"id":"d2c5e076ba6cf7d7","type":"text","text":"# 推进计划\n未来四周计划推进的重要事情\n\n文献调研启动\n\n建模重新推导\n\n\n","x":-600,"y":-306,"width":456,"height":347},
{"id":"82708a439812fdc7","type":"text","text":"# 7月已完成\n\n","x":-220,"y":134,"width":440,"height":560},
{"id":"505acb3e6b119076","type":"text","text":"# 6月已完成\n\n\nP1 结果对比\n- Herowind 带3.5气动与fast3.5对比 相同\n- Herowind 带4.0气动与fast4.0对比 相同\n- Herowind 带hrl气动与fast对比 需气动支持15MW\n\nP1 Bladed交流问题汇总\n\nP1 模型线性化原理 done\n- Bladed 线性化理论手册 仔细阅读\n- multibody blade transform\n- fast线性化理论\n- 梳理Bladed线性化方法框架","x":-700,"y":134,"width":440,"height":560},
- {"id":"c18d25521d773705","type":"text","text":"# 计划\n这周要做的3~5件重要的事情,这些事情能有效推进实现OKR。\n\nP1 必须做。P2 应该做\n\n\nP1 柔性部件 叶片、塔架主动力惯性力算法 主线\n- 变形体动力学 简略看看ing\n- 柔性梁弯曲变形振动学习,主线 \n\t- 广义质量 刚度矩阵及含义\n- 如何静力学求解 \n\t- 基于本构方程 读孟的论文\n\t- normal mode shape 能否使用?\n\t- F = kx 外载与弹性势能相等\n\t\n- 梳理bladed动力学框架 this week\n\t- 子结构文献阅读\n\t- 叶片模型建模 done\n\n\nP1 结果对比\n- Bladed与FAST之间的对比 去掉角度 不能直接比较\n- 叶根坐标系转换 this week\n\nP1 编写线性化理论手册\nP1 上手Bladed \\ fast 线性化功能\n\nP2 如何优雅的存储、输出结果。\nP2 yaw 自由度再bug确认 已知原理了\n","x":-594,"y":-693,"width":450,"height":347},
+ {"id":"c18d25521d773705","type":"text","text":"# 计划\n这周要做的3~5件重要的事情,这些事情能有效推进实现OKR。\n\nP1 必须做。P2 应该做\n\n\nP1 柔性部件 叶片、塔架主动力惯性力算法 主线\n- 变形体动力学 简略看看ing\n- 柔性梁弯曲变形振动学习,主线 \n\t- 广义质量 刚度矩阵及含义\n- 如何静力学求解 \n\t- 基于本构方程 读孟的论文\n\t- normal mode shape 能否使用?\n\t- F = kx 外载与弹性势能相等\n\t\n- 梳理bladed动力学框架 this week\n\t- 子结构文献阅读\n\t- 叶片模型建模 done\n\n\nP1 结果对比\n- Bladed与FAST之间的对比 去掉角度 不能直接比较\n- 叶根坐标系转换 \n\t- 叶尖变形量 - 变形向量 dot product 叶根坐标系方向\n\t- 叶片载荷输入量呢 载荷传递在blade mesh.force moment,mesh.orientation = coord_sys.n\n\nP1 编写线性化理论手册\nP1 上手Bladed \\ fast 线性化功能\n\nP2 如何优雅的存储、输出结果。\nP2 yaw 自由度再bug确认 已知原理了\n","x":-594,"y":-693,"width":450,"height":347},
{"id":"30cb7486dc4e224c","type":"text","text":"# 8月已完成\n\n","x":260,"y":134,"width":440,"height":560}
],
"edges":[]
diff --git a/Pasted image 20250626080836.png b/线性化求解器/images/Pasted image 20250626080836.png
similarity index 100%
rename from Pasted image 20250626080836.png
rename to 线性化求解器/images/Pasted image 20250626080836.png
diff --git a/线性化求解器/images/Pasted image 20250626160523.png b/线性化求解器/images/Pasted image 20250626160523.png
new file mode 100644
index 0000000..a0563d9
Binary files /dev/null and b/线性化求解器/images/Pasted image 20250626160523.png differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/González-Assessment of linearization approache.md b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/González-Assessment of linearization approache.md
new file mode 100644
index 0000000..141cef8
--- /dev/null
+++ b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/González-Assessment of linearization approache.md
@@ -0,0 +1,364 @@
+Research Publications at Politecnico di Milano
+
+# Post-Print
+
+This is the accepted version of:
+
+F. Gonzalez, P. Masarati, J. Cuadrado, M.A. Naya
+Assessment of Linearization Approaches for Multibody Dynamics Formulations
+Journal of Computational and Nonlinear Dynamics, Vol. 12, N. 4, 2017, 041009 (7 pages)
+doi:10.1115/1.4035410
+
+The final publication is available at https://doi.org/10.1115/1.4035410
+
+Access to the published version may require subscription.
+
+# When citing this work, cite the original published paper.
+
+$\circledcirc$ 2017 by ASME. This manuscript version is made available under the CC-BY 4.0 license http://creativecommons.org/licenses/by/4.0/
+
+# ASME Accepted Manuscript Repository
+
+Institutional Repository Cover Sheet
+
+
Pierangelo | Masarati |
First | Last |
| ASME Paper Title:Assessment of Linearization Approaches for Multibody Dynamics Formulations |
| |
Authors: | |
| Gonzalez, Francisco; Masarati, Pierangelo; Cuadrado, Javier; Naya, Miguel A. |
| ASME Journal Title: Journal of Computational and Nonlinear Dynamics |
| |
Volume/Issue__12/4 | |
| |
| https://asmedigitalcollection.asme.org/computationalnonlinear/article/doi/10.1115/1 |
| ASMEDigital CollectionURL:2/Assessment-of-Linearization-Approaches-for |
| |
| |
| |
DOI: | |
| 10.1115/1.4035410 |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
+
+# Assessment of Linearization Approaches for Multibody Dynamics Formulations
+
+Francisco Gonz´alez∗
+Juan de la Cierva Post-Doctoral Fellow
+Laboratorio de Ingenierı´a Meca´nica
+Universidade da Coru˜na
+15403 Ferrol, Spain
+Email: f.gonzalez@udc.es
+Pierangelo Masarati
+Professor
+Dipartimento di Scienze e
+Tecnologie Aerospaziali
+Politecnico di Milano
+20156 Milano, Italy
+Email: pierangelo.masarati@polimi.it
+Javier Cuadrado
+Professor
+Laboratorio de Ingenierı´a Meca´nica
+Universidade da Coru˜na
+15403 Ferrol, Spain
+Email: javicuad@cdf.udc.es
+Miguel A. Naya
+Associate Professor
+Laboratorio de Ingenierı´a Meca´nica
+Universidade da Coru˜na
+15403 Ferrol, Spain
+Email: minaya@udc.es
+
+Formulating the dynamics equations of a mechanical system following a multibody dynamics approach often leads to a set of highly nonlinear Differential Algebraic Equations (DAEs). While this form of the equations of motion is suitable for a wide range of practical applications, in some cases it is necessary to have access to the linearized system dynamics. This is the case when stability and modal analysis are to be carried out; the definition of plant and system models for certain control algorithms and state estimators also requires a linear expression of the dynamics.
+
+A number of methods for the linearization of multibody dynamics can be found in the literature. They differ in both the approach that they follow to handle the equations of motion and the way in which they deliver their results, which in turn are determined by the selection of the generalized coordinates used to describe the mechanical system. This selection is closely related to the way in which the kinematic constraints of the system are treated. Three major approaches can be distinguished and used to categorize most of the linearization methods published so far. In this work, we demonstrate the properties of each approach in the linearization of systems in static equilibrium, illustrating them with the study of two representative examples.
+
+# 1 Introduction
+
+Multibody system (MBS) dynamics has experienced a significant growth over the last decades, boosted by advances in computer architecture and software. One of the most attractive features of MBS dynamics is the wide variety of modeling and formulation approaches from which researchers and analysts can choose when dealing with a particular problem [1]. Generally speaking, the system can be modeled with a set of independent generalized coordinates, which leads to a system of Ordinary Differential Equations (ODEs), or using dependent coordinates whose motion is related by one or more kinematic constraints. In the latter case the dynamics equations are expressed with a set of DAEs, which must be handled with appropriate algorithms, e.g., [2,3]. New formulations and methods are proposed and developed by the MBS research community, making it necessary to benchmark their performance and characterize their features. The ultimate goal of this effort is to provide guidelines that help the analyst select the most appropriate formalisms when dealing with a particular problem.
+
+Such guidelines are also necessary to select linearization methods for those applications that require a linear approximation of the equations of motion. This is the case of certain classical control algorithms, which need to represent the plant to be controlled with linear models [4], and some state and input estimators, like Kalman filters [5]. Even when a linear expression of the dynamics is not strictly required, its availability can make it simpler to gain insight into the system behavior. Using reduced order models, extracted from the linearization of the original differential-algebraic equations of motion, simplifies the analysis of structural and aeroelastic problems [6]; linearized models are also useful in stability analysis [7–9]. Modal analysis and the determination of natural frequencies and vibration modes of a system are another important application of linearized dynamics [10].
+
+Several methods have been proposed in the multibody literature to arrive at linearized forms of the dynamics equations. They present different features, depending mainly on the selection of the coordinates with which they describe the mechanical system. Some of them are recursive algorithms for robotic systems [11], others are built on Maggi-Kane’s coordinates [12]. Lagrange multipliers are a popular way to represent constrained systems and methods to linearize the resulting equations exist as well [6, 10, 13]. This paper categorizes linearization methods for multibody dynamics into three main groups, defined by the selection of generalized coordinates. Their properties are demonstrated in the linearization of the dynamics of mechanical systems in static equilibrium, for which Linear Time Invariant (LTI) problems are guaranteed to be obtained. Representative methods of each category were applied to the linearization of test problems and compared in terms of efficiency, ease of use, and accuracy. Moreover, in [14] two types of linearization problems were identified: heavily constrained systems in which the number of degrees of freedom is much lower than the number of kinematic variables, and systems (typically flexible mechanisms) in which both numbers are of similar magnitude. The behavior of the linearization methods was studied with an example of each type. Practical guidelines for the selection of the most appropriate technique for each case are presented to the reader as conclusions.
+
+# 2 Coordinates Selection
+
+A multibody system can be described with a set of $n$ generalized coordinates $\mathbf{q}=\{q_{1},\dots,q_{n}\}^{\mathrm{~T~}}$ related by $m$ kinematic constraints $\Phi(\mathbf{q},t)=\mathbf{0}$ . For the purposes of this paper, these are assumed to be holonomic and linearly independent; accordingly, the $m\times n$ Jacobian matrix of the constraints $\Phi_{\mathbf{q}}$ can be assumed to have full row rank $m$ . The equations of motion can be expressed as a nonlinear system of DAEs as
+
+$$
+\begin{array}{c}{\mathbf{M}\ddot{\mathbf{q}}=\mathbf{f}+\mathbf{f}_{c}}\\ {\Phi\left(\mathbf{q},t\right)=\mathbf{0}}\end{array}
+$$
+
+where $\mathbf{M}\left(\mathbf{q}\right)$ is the $n\times n$ mass matrix, $\mathbf{f}$ is the array of generalized applied and velocity-dependent forces, and $\mathbf{f}_{c}$ are the reaction forces introduced by the kinematic constraints. Equations (1) are often cast into one of the multibody formulations that exist in the literature, which can be classified into three main groups [14]:
+
+1. Minimal Coordinate Set (MCS), in which the formulation has as many coordinates as degrees of freedom the system has, $n-m$ ;
+2. Redundant Coordinate Set (RCS), consisting of the $n$ coordinates and a set of $m$ Lagrange multipliers; and
+3. Unconstrained Coordinate Set (UCS), in which the system is modeled with the original $n$ coordinates of the unconstrained problem and the constraints are embedded in the formulation.
+
+Velocity projection techniques, in which the system velocities are projected onto the subspace of admissible motion [15], fall into the MCS category. Graph-theory-based [16,17] and Transfer Matrix [18] multibody algorithms could also be included in this category.
+
+If a Lagrangian approach is followed, the constraint reactions are expressed in terms of the constraints Jacobian matrix as $\mathbf{f}_{c}=-\Phi_{\mathbf{q}}^{\mathrm{T}}\pmb{\lambda}$ , where $\lambda$ contains $m$ Lagrange multipliers. Using Baumgarte stabilization to remove constraints drift [19] one would directly arrive at a RCS method. Alternatively, eliminating the constraints delivers a UCS approach. UCS methods can also be obtained through penalty formulations [20] or with the force projection approach in [21].
+
+As mentioned, selecting one approach or another determines the output of the linearization methods, their efficiency and the way in which they convey information about the mechanical system. MCS formulations result in a linear problem with the exact spectrum of the constrained system; the eigenanalysis of this problem directly yields the $2\left(n-m\right)$ eigenvalues of the original system. RCS and UCS approaches may give rise to spurious or approximate eigenvalues; on the other hand, their use can be convenient for the sake of efficiency or ease of use. In the subsequent sections, representatives of each category will be compared in terms of effectiveness and efficiency, and their main features will be discussed. Two main factors, namely how easy it is to discriminate spurious eigenvalues and the simplicity of the linearized equations, also from the computational point of view, will be considered in the discussion.
+
+# 3 Linearization Methods
+
+In this section three methods to linearize Eqs. (1) are introduced. The first one is a MCS formulation obtained via velocity projections [22]. The second directly linearizes a RCS description of the dynamics [6]. The last one is a UCS formulation obtained replacing the kinematic constraints with penalty systems [20].
+
+# 3.1 MCS Formulation: Velocity Projection
+
+The system dynamics can be expressed in terms of a set of $n-m$ independent velocities $\dot{\mathbf{z}}$ using the transformation
+
+$$
+{\dot{\mathbf{q}}}=\mathbf{R}{\dot{\mathbf{z}}}
+$$
+
+where $\mathbf{R}\left(\mathbf{q}\right)$ is an $n\times(n-m)$ velocity transformation matrix that verifies $\mathbf{R}^{\mathrm{T}}\boldsymbol{\Phi}_{\mathbf{q}}^{\mathrm{T}}=\mathbf{0}$ . Matrix $\mathbf{R}$ can be determined via several methods: e.g., coordinate selection (splitting $\mathbf{q}$ into independent and dependent variables), the zero eigenvalue theorem [23], Singular Value Decomposition (SVD) [24,25], QR decomposition [26], and Gram-Schmidt orthonormalization [27, 28]. All these methods determine the subspace $\mathbf{R}$ of the velocities that complies with the constraints through operations on the Jacobian $\Phi_{\mathbf{q}}$ . Applying the transformation in Eq. (2), the system dynamics become a system of $(n-m)$ ODEs
+
+$$
+\mathbf{H}_{1}=\mathbf{R}^{\mathrm{T}}\mathbf{M}\mathbf{R}\ddot{\mathbf{z}}-\mathbf{R}^{\mathrm{T}}\left(\mathbf{f}-\mathbf{M}\dot{\mathbf{R}}\dot{\mathbf{z}}\right)=\mathbf{0}
+$$
+
+If the inputs are represented by $\mathbf{u}$ , the system can be linearized about an equilibrium configuration $\mathbf{z}_{0},\,\dot{\mathbf{z}}_{0},\,\ddot{\mathbf{z}}_{0},\,\mathbf{u}_{0}$ as follows
+
+$$
+\begin{array}{c}{{{\bf{H}}_{1}\left({{{\bf{z}}_{0}}+\delta{{\bf{z}},{{\dot{\bf{z}}}}_{0}}+\delta{{\dot{\bf{z}}},{{\dot{\bf{z}}}}_{0}}+\delta{{\ddot{\bf{z}}},{{\bf{u}}_{0}}}+\delta{\bf{u}}\right)\cong}}\\ {{\left.{\frac{{\partial{\bf{H}}_{1}}}{{\partial{\bf{z}}}}}\right|_{0}\delta{\bf z}+\frac{{\partial{\bf{H}}_{1}}}{{\partial{\dot{\bf{z}}}}}\bigg\vert_{0}\delta{\dot{\bf z}}+\frac{{\partial{\bf{H}}_{1}}}{{\partial{\dot{\bf z}}}}\bigg\vert_{0}\,\delta{\dot{\bf z}}+\frac{{\partial{\bf{H}}_{1}}}{{\partial{\bf{u}}}}\bigg\vert_{0}\,\delta{\bf u}}\end{array}
+$$
+
+The terms in Eq. (4) take the form
+
+$$
+\begin{array}{r l}&{\frac{\partial\mathbf{H}_{1}}{\partial\mathbf{z}}=\mathbf{K}_{r}=\left(\frac{\partial\mathbf{R}^{\mathsf{T}}}{\partial\mathbf{z}}\mathbf{M}\mathbf{R}+\mathbf{R}^{\mathsf{T}}\frac{\partial\mathbf{M}}{\partial\mathbf{z}}\mathbf{R}+\mathbf{R}^{\mathsf{T}}\mathbf{M}\frac{\partial\mathbf{R}}{\partial\mathbf{z}}\right)\hat{\boldsymbol{z}}}\\ &{\quad\quad-\frac{\partial\mathbf{R}^{\mathsf{T}}}{\partial\mathbf{z}}\left(\mathbf{f}-\mathbf{M}\mathbf{R}\hat{\boldsymbol{z}}\right)-\mathbf{R}^{\mathsf{T}}\left(\frac{\partial\mathbf{f}}{\partial\mathbf{z}}-\frac{\partial\mathbf{M}}{\partial\mathbf{z}}\mathbf{R}\hat{\boldsymbol{z}}-\mathbf{M}\frac{\partial\mathbf{R}\hat{\boldsymbol{z}}}{\partial\mathbf{z}}\right)}\\ &{\frac{\partial\mathbf{H}_{1}}{\partial\hat{\boldsymbol{z}}}=\mathbf{C}_{r}=-\mathbf{R}^{\mathsf{T}}\left(\frac{\partial\mathbf{f}}{\partial\hat{\boldsymbol{z}}}-\mathbf{M}\frac{\partial\mathbf{R}\hat{\boldsymbol{z}}}{\partial\hat{\boldsymbol{z}}}\right)}\\ &{\frac{\partial\mathbf{H}_{1}}{\partial\hat{\boldsymbol{z}}}=\mathbf{M}_{r}=\mathbf{R}^{\mathsf{T}}\mathbf{M}\mathbf{R}}\\ &{\frac{\partial\mathbf{H}_{1}}{\partial\mathbf{u}}=-\mathbf{F}_{r}=-\mathbf{R}^{\mathsf{T}}\frac{\partial\mathbf{f}}{\partial\mathbf{u}}}\end{array}
+$$
+
+The expressions for the numerical evaluation of the partial derivatives of $\mathbf{R}$ are detailed in [29]. The linearized dynamics can then be written as
+
+$$
+\mathbf{M}_{r}\delta{\ddot{\mathbf{z}}}+\mathbf{C}_{r}\delta{\dot{\mathbf{z}}}+\mathbf{K}_{r}\delta\mathbf{z}=\mathbf{F}_{r}\delta\mathbf{u}
+$$
+
+Equation (6) has $2\left(n-m\right)$ eigenvalues that represent the exact spectrum of the problem. Moreover, its leading matrix $\mathbf{M}_{r}$ is symmetric and positive-definite. As a consequence, Eq. (6) can be used in the solution of forward-dynamics problems with explicit integration schemes.
+
+# 3.2 RCS Formulation: Generalized Eigenanalysis
+
+Following a Lagrangian approach, the dynamics equations (1) can be expressed in the form
+
+$$
+\mathbf{H}_{2}=\left\{\begin{array}{c}{\mathbf{M}\ddot{\mathbf{q}}+\boldsymbol{\Phi}_{\mathbf{q}}^{\mathrm{T}}\boldsymbol{\lambda}-\mathbf{f}}\\ {\boldsymbol{\Phi}\left(\mathbf{q},t\right)}\end{array}\right\}=\mathbf{0}
+$$
+
+which can be linearized directly and cast in descriptor form
+
+$$
+\mathbf{E}_{q}\delta\dot{\mathbf{y}}=\mathbf{A}_{q}\delta\mathbf{y}+\mathbf{B}_{q}\delta\mathbf{u}
+$$
+
+$$
+\mathbf{E}_{q}=\left[\begin{array}{c c}{\mathbf{I}}&{\mathbf{0}}&{\mathbf{0}}\\ {\mathbf{0}}&{\mathbf{M}_{q}\mathbf{\Lambda}\mathbf{0}}\\ {\mathbf{0}}&{\mathbf{0}}&{\mathbf{0}}\end{array}\right];\qquad\mathbf{A}_{q}=\left[\begin{array}{c c}{\mathbf{0}}&{\mathbf{I}}&{\mathbf{0}}\\ {-\mathbf{K}_{q}-\mathbf{C}_{q}\mathbf{\Lambda}-\mathbf{\Phi}\mathbf{\Phi}\mathbf{\Phi}\mathbf{0}}\\ {-\mathbf{\Phi}\mathbf{\Phi}\mathbf{q}}&{\mathbf{0}}&{\mathbf{0}}\end{array}\right]
+$$
+
+with
+
+$$
+{\bf M}_{q}={\bf M}\,;\quad{\bf C}_{q}=-\frac{\partial{\bf f}}{\partial\dot{\bf q}}\,;\quad{\bf K}_{q}=\frac{\partial{\bf M}}{\partial{\bf q}}\ddot{\bf q}+\frac{\partial\Phi_{\bf q}^{\mathrm{T}}}{\partial{\bf q}}\lambda-\frac{\partial{\bf f}}{\partial{\bf q}}
+$$
+
+It must be noted that the evaluation of $\mathbf{K}_{q}$ requires the computation of the Lagrange multipliers $\lambda$ at equilibrium. Besides the $2\left(n-m\right)$ eigenvalues in the spectrum of the constrained problem, the generalized eigenanalysis of the $(2n+m)$ linearized equations (8) introduces $3m$ spurious eigenvalues, related to the constrained kinematic variables and the Lagrange multipliers. These are easily identifiable because their value is either infinity [6] or zero, if either or both $\Phi_{\mathbf{q}}$ and $\Phi_{\mathbf{q}}^{\mathrm{T}}$ are transferred from $\mathbf{A}_{q}$ to $\mathbf{E}_{q}$ . Because Eq. (7) represents a differential-algebraic problem, matrix $\mathbf{E}_{q}$ is structurally singular. Therefore, Eq. (8) cannot be used in forward-dynamics problems with explicit numerical integrators. It is possible though to overcome this issue applying a QZ decomposition to matrices $\mathbf{E}_{q}$ and $\mathbf{A}_{q}$ [30]. The details of this method are discussed in [6].
+
+# 3.3 UCS Formulation: Penalty Method
+
+Penalty-based relaxation of the constraints can be used to transform the system of DAEs in Eq. (1) into a set of $n$
+
+ODEs [20]. This approach makes the Lagrange multipliers $\lambda$ proportional to the constraints violation at the configuration, velocity, and acceleration levels
+
+$$
+\pmb{\lambda}=\Xi\left(\ddot{\Phi}+\Theta\dot{\Phi}+\Omega\Phi\right)
+$$
+
+where $\Xi$ is an $n\times n$ matrix that contains the penalty factors; $\Theta$ and $\pmb{\Omega}$ are also $n\times n$ terms with stabilization parameters that have a similar function to those used in Baumgarte stabilization [31]. The derivatives of the constraints with respect to time are $\dot{\Phi}\,{=}\,\Phi_{{\bf q}}\dot{\bf q}+\Phi_{t}$ and $\ddot{\Phi}=\Phi_{\mathbf{q}}\ddot{\mathbf{q}}+\dot{\Phi}_{\mathbf{q}}\dot{\mathbf{q}}+\dot{\Phi}_{t}$ , where $\Phi_{t}=\partial\Phi/\partial t$ . This approach is equivalent to replacing the constraints with mass-spring-damper systems. The resulting dynamics equations are
+
+$$
+\mathbf{H}_{3}=\mathbf{M}\ddot{\mathbf{q}}+\boldsymbol{\Phi}_{\mathbf{q}}^{\mathrm{T}}\boldsymbol{\Xi}\left(\ddot{\boldsymbol{\Phi}}+\boldsymbol{\Theta}\dot{\boldsymbol{\Phi}}+\boldsymbol{\Omega}\boldsymbol{\Phi}\right)-\mathbf{f}=\mathbf{0}
+$$
+
+It must be noted that the equilibrium configuration of Eq. (12) is not completely equivalent to that of Eq. (1). In most cases a certain violation of the kinematic constraints $\Phi={\bf0}$ will be required to balance the applied forces f. The linearized dynamics can be expressed as
+
+$$
+\begin{array}{l}{{\displaystyle{\bf H}_{3}\left({\bf q}_{0}+\delta{\bf q},{\dot{\bf q}}_{0}+\delta{\dot{\bf q}},{\ddot{\bf q}}_{0}+\delta{\dot{\bf q}},{\bf u}_{0}+\delta{\bf u}\right)\cong}}\\ {{\displaystyle\left.\frac{\partial{\bf H}_{3}}{\partial{\bf q}}\right\rvert_{0}\delta{\bf q}+\left.\frac{\partial{\bf H}_{3}}{\partial{\dot{\bf q}}}\right\rvert_{0}\delta{\dot{\bf q}}+\left.\frac{\partial{\bf H}_{3}}{\partial{\dot{\bf q}}}\right\rvert_{0}\delta{\ddot{\bf q}}+\left.\frac{\partial{\bf H}_{3}}{\partial{\bf u}}\right\rvert_{0}\delta{\bf u}}}\end{array}
+$$
+
+or in the more compact form
+
+$$
+\mathbf{M}_{p}\delta\ddot{\mathbf{q}}+\mathbf{C}_{p}\delta\dot{\mathbf{q}}+\mathbf{K}_{p}\delta\mathbf{q}=\mathbf{F}_{p}\delta\mathbf{u}
+$$
+
+using the terms
+
+$$
+\begin{array}{l}{\displaystyle\frac{{\partial{\bf{H}}_{3}}}{{\partial{\bf{q}}}}={\bf{K}}_{p}=\frac{{\partial{\bf{M}}}}{{\partial{\bf{q}}}}{\dot{\bf{q}}}+\frac{{\partial\Phi_{\bf{q}}^{\mathrm{T}}}}{{\partial{\bf{q}}}}\Xi\left({\ddot{\Phi}}+\Theta{\dot{\Phi}}+\Omega{\Phi}\right)\quad}\\ {\displaystyle\quad\quad+\ \Phi_{\bf{q}}^{\mathrm{T}}\Xi\left({\Omega{\Phi_{\bf{q}}}+\frac{{\partial{\Phi_{\bf{q}}}}}{{\partial{\bf{q}}}}\left({\ddot{\bf{q}}}+{\bf{\Theta}}{\dot{\bf{q}}}\right)+\frac{{\partial{\dot{\Phi}}_{\bf{q}}}}{{\partial{\bf{q}}}}{\dot{\bf{q}}}\right)\quad}\\ {\displaystyle\quad\quad+\ \Phi_{\bf{q}}^{\mathrm{T}}\Xi\left({\Theta\frac{{\partial{\Phi_{t}}}}{{\partial{\bf{q}}}}+\frac{{\partial{\dot{\Phi}}_{t}}}{{\partial{\bf{q}}}}}\right)-\frac{{\partial{\bf{f}}}}{{\partial{\bf{q}}}}}\end{array}
+$$
+
+
+Fig. 1: An $N$ -loop four-bar linkage with spring elements
+
+together with $2m$ spurious eigenvalues related to the constrained coordinates. True and spurious eigenvalues can be told apart by checking if their associated eigenvectors v verify the velocity-level kinematic constraints Φ˙ΦΦ = 0. The violation of such constraints remains close to zero for true eigenvalues, while it is not negligible for the spurious ones.
+
+As it happened in Section 3.1, if the penalty matrix $\Xi$ is correctly chosen, $\mathbf{M}_{p}$ is symmetric and positive-definite, so Eq. (14) can be directly used in forward-dynamics applications.
+
+# 4 Numerical Examples
+
+The methods described in Section 3 were applied to the linearization of mechanical systems about static equilibrium configurations. Two examples were selected: the first was an $N$ -loop four-bar linkage with spring elements along the diagonals. The second was a flexible double pendulum. The four-bar linkage is heavily constrained and only has one degree of freedom; it is representative of mechanical systems in which the number of kinematic constraints is similar to the number of generalized coordinates $\left(n-m\ll n\right)$ . The double pendulum, on the other hand, includes degrees of freedom associated with flexibility among the generalized coordinates $\mathbf{q}$ . In this case, $n-m\approx n$ .
+
+$$
+\begin{array}{r l}&{\frac{\partial{\bf H}_{3}}{\partial{\bf\dot{q}}}={\bf C}_{p}}\\ &{\quad\quad=\Phi_{{\bf q}}^{\mathrm{T}}\Xi\left(\frac{\partial{\bf\dot{\Phi}}_{{\bf q}}}{\partial{\bf\dot{q}}}{\bf\dot{q}}+\frac{\partial{\bf\dot{\Phi}}_{t}}{\partial{\bf\dot{q}}}+\dot{\bf\Phi}_{{\bf q}}+\Theta{\bf\Phi}_{{\bf q}}\right)-\frac{\partial{\bf f}}{\partial{\bf q}}}\\ &{\quad\quad\frac{\partial{\bf H}_{3}}{\partial{\bf\dot{q}}}={\bf M}_{p}={\bf M}+\Phi_{{\bf q}}^{\mathrm{T}}\Xi\Phi_{{\bf q}}}\\ &{\quad\quad\frac{\partial{\bf H}_{3}}{\partial{\bf u}}=-{\bf F}_{p}=-\frac{\partial{\bf f}}{\partial{\bf u}}}\end{array}
+$$
+
+# 4.1 Multiple Loop Four-Bar Linkage
+
+Figure 1 shows an $N$ -loop four-bar linkage made up of equal rods of length $l_{f}=1\;\mathrm{m}$ and uniformly distributed mass $m_{f}=1~\mathrm{kg}$ . It moves under gravity effects with $g=9.81$ $\mathrm{m}/\mathrm{s}^{2}$ . Each loop $i$ in the linkage has a spring connecting points $B_{i}$ and $A_{i-1}$ of stiffness $k_{f}\,{=}\,25\,\mathrm{N/m}$ and natural length $l_{f0}=\sqrt{2}\,\mathrm{m}$ . If $N_{\mathrm{}}=1$ , the system is equivalent to the test case discussed in [14].
+
+The linkage is modeled with the $x$ and $y$ coordinates of points $B_{0}–B_{N}$ as variables plus the $\boldsymbol{\upvarphi}$ angle from the global $x_{\mathrm{{}}}$ - axis to the rod that connects points $A_{N}$ and $B_{N}$ $\begin{array}{r}{{n=2}N+3}\end{array}$ ). The kinematic constraints term $\Phi$ is composed by equations that enforce that the distances between the tips of each rod remain constant during motion, plus one extra equation that relates the value of $\boldsymbol{\upvarphi}$ to $x_{N}$ and $y_{N}$ $\!\!\!/m=2N+2\!\!\!$ ).
+
+Equation (14) is a system of $n$ ODEs. The method delivers an approximation of the $2\left(n-m\right)$ true system eigenvalues,
+
+# 4.2 Flexible Double Pendulum
+
+Figure 2 shows a planar flexible double pendulum. It is composed of two links with length $l_{p}=4~\mathrm{m}$ and uniformly
+
+the velocity projection become
+
+$$
+\begin{array}{l l}{{\displaystyle{\bf K}_{r}=-\frac{\partial{\bf R}^{\mathrm{T}}}{\partial{\bf z}}{\bf f}-{\bf R}^{\mathrm{T}}\frac{\partial{\bf f}}{\partial{\bf z}}\;;\;\;\;\;\;\;}}&{{\displaystyle{\bf C}_{r}=-{\bf R}^{\mathrm{T}}\frac{\partial{\bf f}}{\partial\dot{\bf z}}}}\\ {~~}&{{\displaystyle{\bf M}_{r}={\bf R}^{\mathrm{T}}{\bf M}{\bf R}\;;\;\;\;}}&{{\displaystyle{\bf F}_{r}={\bf R}^{\mathrm{T}}\frac{\partial{\bf f}}{\partial{\bf u}}}}\end{array}
+$$
+
+
+Fig. 2: A flexible double pendulum
+
+The terms for the linearization of the Lagrangian formulation in Section 3.2 are
+
+$$
+\mathbf{M}_{q}=\mathbf{M}\,;\quad\mathbf{C}_{q}=-{\frac{\partial\mathbf{f}}{\partial{\dot{\mathbf{q}}}}}\,;\quad\mathbf{K}_{q}={\frac{\partial\Phi_{\mathbf{q}}^{\mathrm{T}}}{\partial\mathbf{q}}}{\boldsymbol{\lambda}}-{\frac{\partial\mathbf{f}}{\partial\mathbf{q}}}
+$$
+
+
+Fig. 3: Coordinates used to describe the motion of each link
+
+distributed mass $m_{p}$ , density $\uprho=1500\ \mathrm{kg}/\mathrm{m}^{3}$ , and Young modulus $E=10^{8}\,\,\mathrm{\dot{N}}/\mathrm{m}^{2}$ . The rods have a circular section with radius $r_{p}=0.15\;\mathrm{m}$ . Rod flexibility is modeled following the finite element approach described in [32]. Each link has a tangent floating frame of reference attached to its first articulation, point $o$ for the first link and point $P_{1}$ for the second one, and is discretized with $n_{p}$ finite elements.
+
+The motion of each link is represented through coordinates $x,y$ , and θ, which describe the translation and rotation of the body-fixed reference frame with respect to the global one, and nodal coordinates $q_{f x}^{i},q_{f y}^{i}$ , and $q_{f\Theta}^{i}$ , which represent the elastic deformations at node $i$ , as shown in Fig. 3. The total number of generalized coordinates is $n=6n_{p}+12$ .
+
+Regarding kinematic constraints, four equations are used to fix point $o$ to the ground and to ensure that both links remain connected at point $P_{1}$ during motion. Three more constraints per link enforce that the body-fixed reference frame remains attached to the first node of the rod and tangent to its center line, by making q1fx = q1fy $q_{f x}^{1}=q_{f y}^{1}=q_{f\theta}^{1}=0$ Thus, the total number of constraint equations is $m=10$ .
+
+# 5 Results and Discussion
+
+The linearization of the examples in Section 4 takes place about static equilibrium configurations, in which ${\dot{\mathbf{q}}}=\mathbf{0}$ and $\ddot{\bf q}={\bf0}$ . Under these conditions, terms in Eqs. (5), (10), and (15) adopt simpler expressions. The terms required by
+
+The terms required by the penalty formulation in Section 3.3 become
+
+$$
+\begin{array}{l}{{{\bf{K}}_{p}=\displaystyle\frac{\partial\hat{\bf{\Phi}}_{\bf{q}}^{\mathrm{T}}}{\partial{\bf q}}\Xi\Omega\Phi+\Phi_{\bf{q}}^{\mathrm{T}}\Xi\Omega\Phi_{\bf{q}}-\frac{\partial\mathrm{f}}{\partial{\bf q}}}}\\ {~~}\\ {{{\bf{C}}_{p}=\displaystyle\Phi_{\bf{q}}^{\mathrm{T}}\Xi\Theta\Phi_{\bf{q}}-\frac{\partial\mathrm{f}}{\partial\hat{\bf{q}}}}}\\ {~~}\\ {{{\bf{M}}_{p}={\bf{M}}+\Phi_{\bf{q}}^{\mathrm{T}}\Xi\Phi_{\bf{q}}~;~~~~~~~~~~~~~~~{\bf{F}}_{p}=\frac{\partial\mathrm{f}}{\partial{\bf u}}}}\end{array}
+$$
+
+The linearization methods were implemented in MATLAB; the computation times reported in the following were obtained running the code in an Intel i7-4790K processor at $4.0\ \mathrm{GHz}$ with 8 GB RAM under Windows $10~\mathrm{Pro}$ . These times include the evaluation of the linearization terms in Eqs. (16)–(18) and the solution of the corresponding eigenvalue problem, but not the computation of the equilibrium configuration. In the case of the penalty formulation, they also include the time elapsed in the discrimination of the spurious eigenvalues.
+
+In all the numerical experiments reported in this section, the penalty and stabilization terms required by the penalty formulation were diagonal matrices, proportional to a single scalar value: $\Xi=\alpha\mathbf{I}$ , $\Theta=2\xi\omega\mathbf{I}$ , and $\Omega=\omega^{2}\mathbf{I}$ , where I is the $n\times n$ identity matrix.
+
+# 5.1 Multiple Loop Four-Bar Linkage
+
+Regardless of $N$ , the exact spectrum of the problem is composed of a double complex eigenvalue $s_{1}$ . Angle $\boldsymbol{\upvarphi}$ was selected as independent coordinate with the velocity transformation method. The values selected for the parameters of the penalty method were $\upalpha=10^{6}$ , $\xi=1$ , and ${\mathfrak{o}}=10$ . The MCS and RCS methods delivered the exact value of $s_{1}$ , while the penalty formulation introduced some error in its value as a consequence of the modification of the original system. Table 1 shows the system properties and the time elapsed in linearization with each method.
+
+In order to evaluate the suitability of the methods when dealing with systems with dissipative elements, dampers with $c_{f}=1~\mathrm{Ns/m}$ were introduced alongside the four-bar
+
+Table 1: Properties and spectrum of the $N$ -loop four-bar linkage, and elapsed times in linearization, for different values of $N$
+
+
+ | | | | elapsed (ms) |
N | n | d | S1 | MCS | RCS | UCS |
1 | 5 | 2.2343 | ±2.1477i | 0.115 | 0.108 | 0.185 |
5 | 13 | 1.8922 | ±1.5455i | 0.188 | 0.356 | 0.548 |
10 | 23 | 1.8454 | ±1.4352i | 0.313 | 0.954 | 1.684 |
15 | 33 | 1.8230 | ±1.3955i | 0.556 | 1.872 | 3.612 |
20 | 43 | 1.8217 | ±1.3750i | 0.752 | 4.517 | 6.152 |
+
+Table 3: First eigenvalues of the flexible double pendulum discretized with $n_{p}=50$
+
+S1 | S2 | S3 | S4 |
±1.3357i | ±3.5244i | ±9.4175i | ±22.4054i |
+
+Table 2: Eigenvalues of the damped $N_{}$ -loop four-bar linkage with $c_{f}=1\:\mathrm{Ns/m}$
+
+
+N S1 |
1 -0.2424±2.1340i |
5 -0.2350±1.5276i |
10 -0.2325±1.4162i |
15 -0.2316±1.3761i |
20 -0.2312±1.3554i |
+
+The penalty formulation did not experience any problems in the singular configuration and correctly identified the two double eigenvalues $s_{1}=\pm3.4310i$ and $s_{2}=\pm4.4294\,i$ of the system without any special provisions. The generalized eigenanalysis in Section 3.2 also found the correct eigenvalues. The determination of the static solution required pseudo-inversion, since $\Phi_{\mathbf{q}}$ is no longer full-rank. Indeed, the multipliers $\pmb{\lambda}$ are now underdetermined; nonetheless, the solution remains determined. The MCS formulation, however, failed unless a new selection of independent coordinates was carried out at the singular configuration. The increase in the degree of freedom means that two independent velocities instead of one are necessary to fully define the system motion at this point.
+
+# 5.2 Flexible Double Pendulum
+
+In the flexible double pendulum example, the number of eigenvalues that composes the spectrum of the problem increases with the number of finite elements $n_{p}$ used in the discretization. The dynamics of this system was linearized at the equilibrium configuration in which both rods were aligned along the negative $y$ -axis, i.e., $\theta_{1}=\theta_{2}=-\pi/2$ . Table 3 shows the first four system eigenvalues computed using the MCS method.
+
+
+Fig. 4: A singular static equilibrium configuration for the one-loop four-bar linkage with $k_{f}=0$
+
+The RCS and UCS methods introduced some deviations from the eigenvalues reported in Table 3; both yielded eigenvalues with a small real component. The values of the parameters of the penalty method were $\upalpha=10^{7}$ , $\xi=1$ , and ${\mathfrak{o}}=10$ . Tables 4 and 5 show the elapsed times in the linearization of the dynamics and the maximum differences between the first four eigenvalues computed with the MCS approach $\left({{s}_{i}}\right)$ and the other methods $(s_{i}^{*})$ . It should be mentioned that, while the differences yielded by the UCS method are a consequence of the modification of the original system, the RCS should deliver exact eigenvalues. The errors reported in Table 5 are quite small and likely the result of numerical tolerances in the QZ algorithm used by MATLAB to carry out the eigenanalysis, and in the determination of $\pmb{\lambda}$ required by Eq. (17).
+
+springs in a second set of numerical simulations. All the methods correctly evaluated the system eigenvalues for the damped case, shown in Table 2; the elapsed times were similar to those reported in Table 1.
+
+In all cases the error in the eigenvalues introduced by the penalty formulation remained below $3\cdot10^{-6}$ rad/s. This error depends on the value of the penalty factor $\upalpha$ and the stabilization parameters $\mathbf{\omega}_{\infty}$ and $\xi$ [14].
+
+A particular case of the static equilibrium of the four-bar linkage takes place when $k_{f}=0$ and gravity acts along the global $x_{\mathrm{{}}}$ -axis. In this scenario, the equilibrium configuration occurs at $\upvarphi=0$ , as illustrated in Fig. 4 for $N=1$ . This is a singular configuration at which the Jacobian $\Phi_{\mathbf{q}}$ instantaneously loses rank, leading to the sudden introduction of one extra degree of freedom [33].
+
+# 5.3 Methods Assessment
+
+The study of the two examples in Sections 5.1 and 5.2 highlighted the features of the three linearization approaches considered in this research.
+
+The MCS velocity projection approach in Section 3.1 delivered the exact spectrum of the problems. This method requires selecting a set of independent velocities $\dot{\bf z}$ and the evaluation of the transformation matrix R, which in the general case is accomplished numerically. In problems with $n\approx m$ like the $N$ -loop four-bar linkage, composed mostly of rigid links, the method significantly reduces the number of coordinates and is computationally advantageous. However, if $n\gg m$ , as in the case of the flexible pendulum, the reduction in problem size is not so significant and may be outweighed by the evaluation of R. The method will fail if the selection of independent velocities does not remain valid during the whole motion, as in the case of systems with singular configurations.
+
+Table 4: Linearization of the flexible double pendulum: elapsed times (s)
+
+
+np | 10 | 25 | 50 | 75 | 100 |
MCS | 0.015 | 0.185 | 1.338 | 6.411 | 17.940 |
RCS | 0.020 | 0.176 | 1.229 | 5.800 | 16.072 |
UCS | 0.019 | 0.170 | 1.219 | 5.388 | 14.228 |
+
+Table 5: Flexible double pendulum: maximum differences with respect to the eigenvalues evaluated with the MCS method
+
+
+np | max |Re (si)-Re(s*)Il RCS UCS | max Im (si) - Im (s*)Il RCS UCS |
10 | 7.6.10-11 1.4·10-4 | 6.7.10-11 3.6·10-5 |
25 | 4.0·10-9 1.4·10-4 | 3.9·10-9 3.6·10-5 |
50 | 5.6·10-8 1.4·10-4 | 8.2·10-9 3.6·10-5 |
75 | 9.6·10-8 1.4·10-4 | 9.9·10-7 3.7·10-5 |
100 | 1.5.10-7 1.4·10-4 | 2.5·10-6 3.8·10-5 |
+
+The RCS and UCS methods added spurious eigenvalues to the problems spectra. The direct linearization described in Section 3.2 is much simpler than evaluating the terms required by the penalty and velocity transformation methods in Eqs. (5) and (15). Moreover, the spurious terms are easily identifiable. However, the linearized dynamics obtained this way cannot be directly used in explicit timeintegration schemes. The decomposition process required to circumvent this drawback can be computationally expensive, although subspace methods, e.g., the Implicitly Restarted Arnoldi Method in ARPACK [34], can be used to alleviate this issue. Indeed, in practical applications only a subspace of the entire eigenspace is usually needed. Furthermore, without even considering the spurious eigenvalues, computing the entire spectrum in practical problems with thousands to millions degrees of freedom would be too computationally expensive and of little use, if at all feasible.
+
+The penalty method in Section 3.3 delivers approximated eigenvalues, as a consequence of relaxing the original kinematic constraints and replacing them with penalty systems, thus modifying the equilibrium point of the system. It also introduces spurious eigenvalues in the spectrum. The parameters of the method must be adjusted [33,35] to ensure that the linearized dynamics in Eq. (14) can be considered an adequate approximation of the constrained system’s exact one.
+
+Table 6: Summary of method properties
+
+
+Feature | MCS | RCS | UCS |
Efficiency (n ≈ m) | | | |
Efficiency (n 》m) | | | |
Forward dynamics with explicit integration | Yes | No | Yes |
Exact eigenvalues | Yes | Yes | No |
Spurious eigenvalues | No | Yes | Yes |
Handles singular configurations | No | Yes | Yes |
Degrades sparsity | Yes | No | Moderately |
+
+Each approach also affects the sparsity pattern of the system matrices in a different manner. The sparsity patterns of matrices $\mathbf{M}_{r}$ , $\mathbf{M}_{q}$ , and $\mathbf{M}_{p}$ in the double pendulum example with $n_{p}=5$ are compared in Fig. 5, both in their original form and after a Cholesky factorization. Direct eigenanalysis preserves matrix sparsity. The velocity transformation of the MCS method reduces the system size, but at the cost of increasing the fill-in of the system matrix after factorization. The penalty method introduces some additional structural non-zeros, but its effect on sparsity is less severe. In practice, both the RCS and the UCS approaches preserve the structure of the mass matrix in terms of separation between the inertia coefficients of each structural component. As a consequence, the Cholesky factors of Figs. 5d and 5f produce a characteristic block-diagonal structure. It must be noted that in systems with a more complicated, non openloop topology, the constraints responsible for loop closure would introduce some coupling terms off the main block diagonal in matrix $\Phi_{\mathbf{q}}^{\mathrm{T}}\Xi\Phi_{\mathbf{q}}$ and, consequently, in matrix $\mathbf{M}_{p}$ . On the contrary, the MCS structurally couples the degrees of freedom of the components, thus completely filling the Cholesky factor of Fig. 5b.
+
+The features of each method are summarized in Table 6.
+
+# 6 Conclusions
+
+When dependent coordinates are used to model a multibody system, kinematic constraints can be handled with a wide variety of methods, which can be grouped into three major categories: Minimal Coordinate Set (MCS) methods, in which the dynamics are expressed in terms of a set of independent velocities; Redundant Coordinate Set (RCS) methods, in which the set of coordinates is enlarged with as many Lagrange multipliers as constraints the system has; and Unconstrained Coordinate Set (UCS) methods, which embed the constraints in the dynamics equations. This work assessed the properties of each family of methods in the linearization of two examples in static equilibrium configurations: an $N$ -loop four-bar linkage with spring elements and a flexible double pendulum modeled with finite elements.
+
+
+Fig. 5: Sparsity pattern of the mass matrix of the flexible double pendulum with $n_{p}=5$
+
+Results confirmed that the method choice affects the linearization process in terms of the way in which information about the linearized system behavior is delivered, accuracy, and computational efficiency. While MCS methods obtained the exact problem spectrum, RCS and UCS introduced spurious eigenvalues in the spectra that needed to be properly identified and dealt with. The RCS approach featured a simple expression of the linearization terms, while the resultant descriptor form equations required additional processing before they could be used in forward-dynamics settings with explicit integration. MCS and UCS methods, on the other hand, provided compact expressions that could be readily used to this end. Regarding efficiency, the MCS method proved to be advantageous in the linearization of heavily constrained problems. Conversely, if the number of constraints was small compared to the total number of variables used in the modeling, RCS and UCS methods showed a superior performance. As a consequence, the linearization method selection must be carried out taking into consideration the characteristics of the mechanical system and the intended use of the linearized equations of motion.
+
+# Acknowledgements
+
+The first author would like to acknowledge the support of the Spanish Ministry of Economy through its postdoctoral research program Juan de la Cierva, contract No. JCI-2012-12376.
+
+# References
+
+[1] Bauchau, O. A., 2011. Flexible Multibody Dynamics. Springer.
+[2] Mariti, L., Belfiore, N. P., Pennestr\`ı, E., and Valentini, P. P., 2011. “Comparison of solution strategies for multibody dynamics equations”. International Journal for Numerical Methods in Engineering, 88(7), pp. 637– 656.
+[3] Marques, F., Souto, A., and Flores, P., 2016. “On the constraints violation in forward dynamics of multibody systems”. Multibody System Dynamics, Online First, pp. 1–35.
+[4] Ogata, K., 2001. Modern Control Engineering, 4th ed. Prentice Hall.
+[5] Grewal, M. S., and Andrews, A. P., 2015. Kalman Filtering: Theory and Practice with MATLAB, 4th ed. Wiley-IEEE Press.
+[6] Ripepi, M., and Masarati, P., 2011. “Reduced order models using generalized eigenanalysis”. Proceedings of the Institution of Mechanical Engineers, Part K: Journal of Multi-body Dynamics, 225(1), pp. 52–65.
+[7] Escalona, J. L., and Chamorro, R., 2008. “Stability analysis of vehicles on circular motions using multibody dynamics”. Nonlinear Dynamics, 53(3), pp. 237– 250.
+[8] Masarati, P., 2013. “Estimation of Lyapunov exponents from multibody dynamics in differentialalgebraic form”. Proceedings of the Institution of Mechanical Engineers, Part K: Journal of Multi-body Dynamics, 227(4), pp. 23–33.
+[9] Masarati, P., and Tamer, A., 2015. “Sensitivity of trajectory stability estimated by Lyapunov characteristic exponents”. Aerospace Science and Technology, 47, pp. 501–510.
+[10] Masarati, P., 2009. “Direct eigenanalysis of constrained system dynamics”. Proceedings of the Institution of Mechanical Engineers, Part K: Journal of Multi-body Dynamics, 223(4), pp. 335–342.
+[11] Gontier, C., and Li, Y., 1995. “Lagrangian formulation and linearization of multibody system equations”. Computers & Structures, 57(2), pp. 317–331.
+[12] Peterson, D. L., Gede, G., and Hubbard, M., 2015. “Symbolic linearization of equations of motion of constrained multibody systems”. Multibody System Dynamics, 33(2), pp. 143–161.
+[13] Negrut, D., and Ortiz, J. L., 2006. “A practical approach for the linearization of the constrained multibody dynamics equations”. Journal of Computational and Nonlinear Dynamics, 1(3), pp. 230–239.
+[14] Gonz´alez, F., Masarati, P., and Cuadrado, J., 2016. “On the linearization of multibody dynamics formulations”. In Proceedings of the ASME 2016 International Design Engineering Technical Conferences & Computers and Information in Engineering Conference, no. DETC2016-59227.
+[15] Serna, M. A., Avil´es, R., and Garc´ıa de Jal´on, J., 1982. “Dynamic analysis of plane mechanisms with lower pairs in basic coordinates”. Mechanism and Machine Theory, 17(6), pp. 397–403.
+[16] Jain, A., 2011. “Graph theoretic foundations of multibody dynamics. Part II: analysis and algorithms”. Multibody System Dynamics, 26(3), pp. 307–333.
+[17] Jain, A., 2012. “Multibody graph transformations and analysis – Part II: Closed-chain constraint embedding”. Nonlinear Dynamics, 67(3), pp. 2153–2170.
+[18] Rong, B., Rui, X., and Wang, G., 2010. “Modified finite element transfer matrix method for eigenvalue problem of flexible structures”. Journal of Applied Mechanics, 78(2), p. 021016.
+[19] Baumgarte, J., 1972. “Stabilization of constraints and integrals of motion in dynamical systems”. Computer Methods in Applied Mechanics and Engineering, 1(1), pp. 1–16.
+[20] Bayo, E., Garc´ıa de Jal´on, J., and Serna, M. A., 1988. “A modified Lagrangian formulation for the dynamic analysis of constrained mechanical systems”. Computer Methods in Applied Mechanics and Engineering, 71(2), pp. 183–195.
+[21] Masarati, P., 2011. “Adding kinematic constraints to purely differential dynamics”. Computational Mechanics, 47(2), pp. 187–203.
+[22] Garc´ıa de Jalo´n, J., and Bayo, E., 1994. Kinematic and Dynamic Simulation of Multibody Systems. The Real– Time Challenge. Springer–Verlag.
+[23] Kamman, J. W., and Huston, R. L., 1984. “Dynamics of constrained multibody systems”. Journal of Applied Mechanics, 51(4), December, pp. 899–903.
+[24] Singh, R. P., and Likins, P. W., 1985. “Singular value decomposition for constrained dynamical systems”. Journal of Applied Mechanics, 52(4), pp. 943– 948.
+[25] Mani, N. K., Haug, E. J., and Atkinson, K. E., 1985. “Application of singular value decomposition for analysis of mechanical system dynamics”. Journal of Mechanisms, Transmissions, and Automation in Design, 107(1), pp. 82–87.
+[26] Kim, S. S., and Vanderploeg, M. J., 1986. “QR decomposition for state space representation of constrained mechanical dynamic systems”. Journal of Mechanisms, Transmissions, and Automation in Design, 108(2), pp. 183–188.
+[27] Liang, C. G., and Lance, G. M., 1987. “A differentiable null space method for constrained dynamic analysis”. Journal of Mechanisms, Transmissions, and Automation in Design, 109(3), pp. 405–411.
+[28] Agrawal, O. P., and Saigal, S., 1989. “Dynamic analysis of multi-body systems using tangent coordinates”. Computers & Structures, 31(3), pp. 349–355.
+[29] Dopico, D., Zhu, Y., Sandu, A., and Sandu, C., 2014. “Direct and adjoint sensitivity analysis of ordinary differential equation multibody formulations”. Journal of Computational and Nonlinear Dynamics, 10(1), p. paper 011012.
+[30] Moler, C. B., and Stewart, G. W., 1973. “An algorithm for generalized matrix eigenvalue problems”. SIAM Journal on Numerical Analysis, 10(2), pp. 241–256.
+[31] Gonza´lez, F., and K¨ovecses, J., 2013. “Use of penalty formulations in dynamic simulation and analysis of redundantly constrained multibody systems”. Multibody System Dynamics, 29(1), pp. 57–76.
+[32] Shabana, A. A., 1998. Dynamics of Multibody Systems, 2nd ed. Cambridge University Press.
+[33] Gonza´lez, F., Dopico, D., Pastorino, R., and Cuadrado, J., 2016. “Behaviour of augmented Lagrangian and Hamiltonian methods for multibody dynamics in the proximity of singular configurations”. Nonlinear Dynamics, 85(3), pp. 1491–1508.
+[34] Lehoucq, B., Sorensen, D. C., and Vu, P., 1995. ARPACK: An implementation of the Implicitly Restarted Arnoldi Iteration that computes some of the eigenvalues and eigenvectors of a large sparse matrix. Tech. rep. Available from netlib $@$ ornl.gov under the directory scalapack.
+[35] Flores, P., Machado, M., Seabra, E., and Tavares da Silva, M., 2011. “A parametric study on the Baumgarte stabilization method for forward dynamics of constrained multibody systems”. Journal of Computational and Nonlinear Dynamics, 6(1), p. paper 011019.
\ No newline at end of file
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/González-Assessment of linearization approache_origin.pdf b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/González-Assessment of linearization approache_origin.pdf
new file mode 100644
index 0000000..7d1ba09
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/González-Assessment of linearization approache_origin.pdf differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/10fed2e277063cbd855a276a052553425c3d37db4417e6aa393ee43f483ee171.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/10fed2e277063cbd855a276a052553425c3d37db4417e6aa393ee43f483ee171.jpg
new file mode 100644
index 0000000..849c25e
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/10fed2e277063cbd855a276a052553425c3d37db4417e6aa393ee43f483ee171.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/18bf589199ee47e50e2141234453ecf60fee110883d93fb0c773aa49a23802db.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/18bf589199ee47e50e2141234453ecf60fee110883d93fb0c773aa49a23802db.jpg
new file mode 100644
index 0000000..b059145
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/18bf589199ee47e50e2141234453ecf60fee110883d93fb0c773aa49a23802db.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/24f97cf7bdba31ee1f5f6d667bc24f48331904db335b81dec3482dad0854c06b.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/24f97cf7bdba31ee1f5f6d667bc24f48331904db335b81dec3482dad0854c06b.jpg
new file mode 100644
index 0000000..422c302
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/24f97cf7bdba31ee1f5f6d667bc24f48331904db335b81dec3482dad0854c06b.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/44082296835398cefecdd2667ae0537958bc10385ea8c96db6249570ada4107f.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/44082296835398cefecdd2667ae0537958bc10385ea8c96db6249570ada4107f.jpg
new file mode 100644
index 0000000..a21da2d
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/44082296835398cefecdd2667ae0537958bc10385ea8c96db6249570ada4107f.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/4dbaa92b67aa376a9c626b28d98d1f4d1a3a9ee5c377f60bc1a0d6a3d7c22fac.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/4dbaa92b67aa376a9c626b28d98d1f4d1a3a9ee5c377f60bc1a0d6a3d7c22fac.jpg
new file mode 100644
index 0000000..ea08f95
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/4dbaa92b67aa376a9c626b28d98d1f4d1a3a9ee5c377f60bc1a0d6a3d7c22fac.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/7399fa4b679180befc20e9c026e0f78ad0a0b1b0185c1509a99665b835fc1eec.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/7399fa4b679180befc20e9c026e0f78ad0a0b1b0185c1509a99665b835fc1eec.jpg
new file mode 100644
index 0000000..00cce3a
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/7399fa4b679180befc20e9c026e0f78ad0a0b1b0185c1509a99665b835fc1eec.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/a9a611fe433398bd8f21584f0a401a475bced6da4af426ee22c8fd7121d9d59e.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/a9a611fe433398bd8f21584f0a401a475bced6da4af426ee22c8fd7121d9d59e.jpg
new file mode 100644
index 0000000..d095d1e
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/a9a611fe433398bd8f21584f0a401a475bced6da4af426ee22c8fd7121d9d59e.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/b7c62f9b940c0204f1d552509ca0ed24d79ea86f8bba56142f4ebe17201cb819.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/b7c62f9b940c0204f1d552509ca0ed24d79ea86f8bba56142f4ebe17201cb819.jpg
new file mode 100644
index 0000000..45ca45f
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/b7c62f9b940c0204f1d552509ca0ed24d79ea86f8bba56142f4ebe17201cb819.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/ce1c48625d6a75b393e72dafab788aea716830d364e810bc9e34d64ce2979baa.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/ce1c48625d6a75b393e72dafab788aea716830d364e810bc9e34d64ce2979baa.jpg
new file mode 100644
index 0000000..cf9b5e9
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/ce1c48625d6a75b393e72dafab788aea716830d364e810bc9e34d64ce2979baa.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/dd0d84c05f31f9e3a688e4a9c7eee85af6568099ad0d1df0bdcffc364325970e.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/dd0d84c05f31f9e3a688e4a9c7eee85af6568099ad0d1df0bdcffc364325970e.jpg
new file mode 100644
index 0000000..7113168
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/dd0d84c05f31f9e3a688e4a9c7eee85af6568099ad0d1df0bdcffc364325970e.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/e31eba505a8a9f8ed7b341ad196ce6cbf1dde26927ccc4a5a81b942b376560bc.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/e31eba505a8a9f8ed7b341ad196ce6cbf1dde26927ccc4a5a81b942b376560bc.jpg
new file mode 100644
index 0000000..8203476
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/e31eba505a8a9f8ed7b341ad196ce6cbf1dde26927ccc4a5a81b942b376560bc.jpg differ
diff --git a/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/e92f05eb9bcb58fc666cf85f41b0dbbe3d185ae21ade1ae75ec99981d807659e.jpg b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/e92f05eb9bcb58fc666cf85f41b0dbbe3d185ae21ade1ae75ec99981d807659e.jpg
new file mode 100644
index 0000000..2adc776
Binary files /dev/null and b/线性化求解器/参考文献/González-Assessment of linearization approache/auto/images/e92f05eb9bcb58fc666cf85f41b0dbbe3d185ae21ade1ae75ec99981d807659e.jpg differ
diff --git a/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/Jonkman-Full-system linearization for floating.md b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/Jonkman-Full-system linearization for floating.md
new file mode 100644
index 0000000..0dee657
--- /dev/null
+++ b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/Jonkman-Full-system linearization for floating.md
@@ -0,0 +1,246 @@
+
+
+# Full-System Linearization for Floating Offshore Wind Turbines in OpenFAST
+
+Preprint
+
+Jason M. Jonkman, Alan D. Wright, Greg J. Hayman, and Amy N. Robertson
+
+National Renewable Energy Laboratory
+
+Presented at the ASME 2018 $7^{s t}$ International
+Offshore Wind Technical Conference
+San Francisco, California
+November 4-7, 2018
+
+# Full-System Linearization for Floating Offshore Wind Turbines in OpenFAST
+
+Preprint
+
+Jason M. Jonkman, Alan D. Wright, Greg J. Hayman, and Amy N. Robertson
+
+National Renewable Energy Laboratory
+
+# Suggested Citation
+
+Jonkman, Jason, Alan Wright, Greg Hayman, and Amy Robertson. 2018. Full-System Linearization for Floating Offshore Wind Turbines in OpenFAST: Preprint. Golden, CO: National Renewable Energy Laboratory. NREL/CP-5000-71865. https://www.nrel.gov/docs/fy19osti/71865.pdf
+
+This work was authored by the National Renewable Energy Laboratory, operated by Alliance for Sustainable Energy, LLC, for the U.S. Department of Energy (DOE) under Contract No. DE-AC36-08GO28308. Funding provided by U.S. Department of Energy Office of Energy Efficiency and Renewable Energy Wind Energy Technologies Office. The views expressed herein do not necessarily represent the views of the DOE or the U.S. Government. The U.S. Government retains and the publisher, by accepting the article for publication, acknowledges that the U.S. Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for U.S. Government purposes.
+
+This report is available at no cost from the National Renewable Energy Laboratory (NREL) at www.nrel.gov/publications.
+
+U.S. Department of Energy (DOE) reports produced after 1991 and a growing number of pre-1991 documents are available free via www.OSTI.gov.
+
+Cover Photos by Dennis Schroeder: (clockwise, left to right) NREL 51934, NREL 45897, NREL 42160, NREL 45891, NREL 48097, NREL 46526.
+
+# FULL-SYSTEM LINEARIZATION FOR FLOATING OFFSHORE WIND TURBINES IN OPENFAST
+
+Jason M Jonkman, Alan D Wright, Greg J Hayman, Amy N Robertson National Renewable Energy Laboratory 15013 Denver West Parkway Golden, Colorado 80401 (USA)
+
+# ABSTRACT
+
+The wind engineering community relies on multiphysics engineering software to run nonlinear time-domain simulations (e.g., for design-standards-based loads analysis). Although most physics involved in wind energy are nonlinear, linearization of the underlying nonlinear system equations is often advantageous to understand the system properties and exploit well-established methods and tools for analyzing linear systems. This paper presents the development of the new linearization functionality of the open-source engineering tool OpenFAST for floating offshore wind turbines, as well as the concepts and mathematical background needed to understand and apply it.
+
+# INTRODUCTION
+
+To support design and analysis—so that wind turbines are innovative, optimized, reliable, and cost-effective—the wind industry and research communities rely on engineering software (i.e., design tools) capable of predicting the coupled dynamic loads and responses of the wind system. OpenFAST (formerly known as FAST), developed by the National Renewable Energy Laboratory (NREL) via support from the U.S. Department of Energy, is an open-source multiphysics tool that enables the design of wind turbines [1]. OpenFAST models the important physical phenomena and system couplings, including the environmental excitation (wind, waves, and current) and full-system dynamic response (rotor, drivetrain, nacelle, support structure, and controller) under both normal (for fatigue) and extreme (for ultimate) loading conditions. The OpenFAST model enables the analysis of a range of wind turbine configurations, including two- or three-bladed horizontal-axis rotors, pitch or stall regulation, rigid or teetering hub, upwind or downwind rotor, and lattice or tubular towers. The wind turbine can be modeled on land or offshore on bottom-fixed or floating substructures.
+
+The primary use of OpenFAST is to run nonlinear timedomain simulations (e.g., for design-standards-based loads analysis). Although most physics involved in wind energy are nonlinear, linearization of the underlying nonlinear system equations is often advantageous to understand the system response and exploit well-established methods and tools for analyzing linear systems. For example, linear state-space models can be transformed to transfer functions, impulseresponse functions, or frequency-response functions. The ability to generate linearized models is important for eigenanalysis (to derive structural natural frequencies, damping ratios, and mode shapes), controls design (based on linear statespace models), stability analysis, gradients for optimization problems, and support for the development of reduced-order models.
+
+Previous FAST linearization work focused on 1) structuring the FAST v8 source code to enable linearization (FAST v8 is the predecessor of OpenFAST); 2) developing the general approach to linearizing the mesh mapping within the module-to-module, input-output coupling relationships, including rotations; 3) linearizing core (but not all) features of the land-based modules of FAST (InflowWind, AeroDyn, ServoDyn, and ElastoDyn) and their coupling; and 4) verifying this implementation by applying the tool to sample cases [2]. More recently, the ability to linearize OpenFAST models with advanced structural dynamics for aeroelastically tailored blades using the BeamDyn module was introduced and verified [3]. This work extends these efforts to linearizing core (but not all) features of the floating offshore wind turbine (FOWT) modules in OpenFAST and their coupling, including the hydrodynamics in HydroDyn and moorings in MAP $^{++}.$ , and verifying the updated implementation.
+
+The HydroDyn hydrodynamics module of OpenFAST [1] allows for multiple approaches for calculating the hydrodynamic loads on floating platforms, including a potential-flow theory solution, a strip-theory solution, or a hybrid combination of the two. Waves may be regular (periodic) or irregular (stochastic), and long crested (unidirectional) or short crested (with wave energy spread across a range of directions). Second-order terms in the wave kinematics and/or potential-flow solution are available options. The $\mathrm{MAP++}$ module of OpenFAST [1] models mooring systems quasi-statically, considering geometric nonlinearities of catenary or taut lines, elastic stretching, line weight, and buoyancy as well as clump weights and buoyancy tanks, lineto-line interconnections, and seabed friction.
+
+The overall linearization approach that the FAST modularization framework was designed to support is explained in [4] and is consistent with the present implementation. Details on the linearization of the land-based modules and their coupling, including mesh mapping, is explained in [2,3]. Without replicating most of the information, this paper uses the same approach and nomenclature of [2,3,4], adding details about the linearization of the floating offshore modules HydroDyn and $\mathrm{MAP++}$ and their coupling, including mesh mapping.
+
+The mathematical background presented in this paper is currently being implemented in the OpenFAST source code. Unfortunately, the implementation at the time of this writing has not yet been completed enough to produce results. Results will be presented in future work to highlight the functionality and verify the implementation (e.g., for a sample case whereby the natural frequencies and damping of the NREL 5-MW baseline wind turbine atop a spar buoy will be calculated as a function of rotational speed and presented in Campbell-diagram form).
+
+The linearization of OpenFAST involves: 1) finding an operating point (OP), 2) linearizing the underlying nonlinear equations of each module about the OP, 3) linearizing the module-to-module input-output coupling relationships in the OpenFAST glue code about the OP, and 4) combining all linearized matrices into the full-system linear state-space model and exporting those matrices and the OP to a file. Each step is highlighted in the following sections.
+
+# OPERATING-POINT DETERMINATION
+
+OP (or fixed-point) determination is an important first step in the linearization process because a linear representation of a nonlinear system is only valid for small deviations (perturbations) from an OP. In the current release of OpenFAST, an OP can be defined by given initial conditions (time zero) or a given time (or times) in the nonlinear timemarching process. It is usually important for the OP to be a static-equilibrium condition (for parked/idling turbines) or steady-state condition (for operating turbines); otherwise, it may have an undesirable effect on the linear system matrices.
+
+An OP is defined by given values for the continuous-time states, $x\big|_{o p}$ , discrete-time states, $\left.x^{d}\right|_{o p}$ , inputs, $u\Big|_{o p}$ , and time, $t\Big|_{o p}$ , for each module. Equations (1a), (1c), and (1d) from [4] can then be used to calculate the OP values of the first time derivative of the continuous-time states, $\dot{x}\big|_{o p}$ , constraint (algebraic) states, $z{\Big|}_{o p}$ , and outputs, ${y\Big|}_{o p}$ , for each module. Each of these variables can be perturbed (represented by $\Delta$ ) about their respective OP values as given by Eq. (11) from [4] (e.g., for module inputs $\boldsymbol{u}=\boldsymbol{u}\Big|_{o p}+\Delta\boldsymbol{u}\,\Big)$ . References [2,3] clarify how this operation is extended to rotations (orientations) in three dimensions (3D), which do not reside in a linear space.
+
+The number of states, inputs, and outputs (i.e., the size of the vectors $x\;,\;x^{d}\;,\;z\;,\;u$ , and $y$ ) depend on the features enabled in OpenFAST.
+
+# MODULE LINEARIZATION
+
+As explained in [4], the FAST modularization framework supports a very general (need not be linear) state-space formulation, with any combination of continuous-time-state, discrete-time-state, constraint- (algebraic-) state, other- (e.g., logical) state, and output equations. However, for a module to support linearization, the formulation is limited to a hybrid semiexplicit differential-algebraic equation (DAE) of index 1,2 which has the following limitations: 1) the continuous-timestate derivatives and discrete-time-state updates must be written as an explicit function of the states, inputs, and parameters; 2) the constraints must be of index 1; and 3) other states are used only for time-integration or when acting as parameters in the linearization process.
+
+To support linearization, a module must also be able to export Jacobian matrices for the state and output equations with respect to the states and inputs. The OpenFAST module states, inputs, and outputs kept in the linearization process for the features linearized to date are summarized in Table 1. The OpenFAST module features linearized to date include only continuous-time and constraint states (no features with discretetime states have yet been linearized).
+
+The linearized form of a general module is given by Eqs. (12) and (13) from [4]; the simplified forms for the land-based modules InflowWind, AeroDyn, ServoDyn, ElastoDyn, and BeamDyn are given in [2,3]. The simplified forms for the newly linearized offshore modules HydroDyn and $\mathrm{MAP++}$ are given next, along with a description of how each module is linearized.
+
+TABLE 1. MODULE STATES, INPUTS, AND OUTPUTS KEPT IN THE OPENFAST LINEARIZATION PROCESS TO DATE.
+
+
+Module | States | Inputs | Outputs |
InflowWind (IfW) | None | Positions where the undisturbed (inflow) wind will be indino Disturbances of horizontal wind speed, power-law shear exponent,andwind-propagation direction | Undisturbed (inflow) wind velocity at input positions User-selected wind-inflow outputs |
ServoDyn (SrvD) | None | Nacelle-yaw angle and rate Generator speed | Blade-pitch-angle command (independent) Nacelle-yaw moment Generator torque and electrical power User-selected control and electrical-drive outputs Translational displacements, orientations, |
ElastoDyn (ED) | Position and orientation and translational and rotational velocities of the floating platform (continuous states) Relative bending-mode degrees of freedom of the bladesandtowerandtheir first-time derivatives (continuous states) Relative rotational displacement and velocity off the nacelle-yaw, generator azimuth, shaft torsion, and rotor-teeter (continuous states) | Applied point forces and moments distributed along the blades and tower Applied point forces and moments lumped on the hub, nacelle, and platform Blade-pitch-angle command (both independent and rotor- collective) Nacelle-yaw moment Generator torque | translational androtationalvelocities,and translational and rotation accelerations of analysis nodes along the blades and tower Translational displacements, orientations, translational androtationalvelocities,and translationalandrotationaccelerationsoftheblade- root, nacelle, and platform reference points Translational displacement,orientation,and rotational velocity of the hub reference point Nacelle-yaw angle and rate Generator speed User-selected structural outputs |
BeamDyn (BD) | Positions and orientations and translational and rotational velocitiesofnodes distributed along eachblade (continuous states) | Translational displacements, orientations, translational androtationalvelocities,andtranslationalandrotational accelerations of eachbladeroot Applied point forces and moments distributed along each blade Applied line (per-unit length) forces and moments distributedalongeachblade | Reaction point forces and moments lumped at each bladeroot Translational displacements,orientations, translational and rotational velocities, and translationalandrotationaccelerationsofnodes along each blade User-selected structural outputs |
AeroDyn (AD) | Inflow angle at each blade analysis node/airfoil (constraint states) | Translational displacements, orientations, and translational velocities of analysis nodes along the blades and tower Orientation of the blade-root reference point Translational displacement, orientation, and rotational velocity of the hub reference point Undisturbed (inflow) wind velocities at analysis nodes along the blades and tower | Aerodynamic-applied line(per-unit length) forces and moments distributed along the blades and tower User-selectedaerodynamicoutputs |
HydroDyn (HD) | State-space-based wave- excitation states (continuous states) State-space-based wave- radiation states (continuous states) | Translational displacements,orientations, translational and rotational velocities, and translational and rotational accelerations of analysis nodes distributed along the floating platform Translational displacement,orientation,translationand rotational velocities, and translational and rotational accelerations of the platform reference point Disturbanceofwaveelevationattheplatformreference point | Hydrodynamic-applied point forces and moments distributed along the floating platform Hydrodynamic-applied line (per-unit length) forces and moments distributed along the floating platform Hydrodynamic-applied point force and moment at the platform reference point User-selected hydrodynamic outputs |
MAP++ (MAP) | Horizontal and vertical tensions at thefairleadof each mooring line (constraint states) Positions of each connect node (constraint states for | Translational displacements of each fairlead | Reaction point forces (tensions) lumped at each fairlead User-selected mooring outputs |
+
+The linearization of the HydroDyn (HD) module applies to both the strip-theory solution, potential-flow solution, or a hybrid combination of the two. Linear state-space-based waveexcitation and wave-radiation models have been added to HydroDyn to enable linearization of the potential-flow solution, including wave-excitation loads (with diffraction) from disturbances of wave elevation and wave-radiation loads with free-surface memory effects.
+
+The linear state-space-based wave-radiation model was previously available in HydroDyn within FAST v8. As shown in [5], the model approximates the convolution term in the Cummins equation as shown in Eq. (1), with a linear statespace model as shown in Eq. (2), where $F_{R d t n}$ is the waveradiation loads; $K_{R d t n}$ is the wave-radiation-retardation kernels; $\dot{q}_{P t f m}$ is the floating platform translation and rotational velocities; $x_{R d t n}$ are the wave-radiation states; and $A_{R d t n}$ ,
+
+$B_{R d t n}$ , and $C_{R d t n}$ are the matrices of the linear state-space wave-radiation model. The more wave-radiation states there are, the better the state-space model can approximate the convolution. Within HydroDyn, the user can select either the convolution-based wave-radiation model, which is implemented via numerical convolution using discrete-time states (which, if linearized, would overly complicate application of the linear state-space model), or the linear statespace-based wave-radiation model, which is implemented using matrices derived from the SS_Fitting preprocessor [6] or equivalent.
+
+$$
+\begin{array}{c}{{F_{R d t n}\left(t\right)=-\displaystyle\int_{}^{t}{{K_{R d t n}}\left(t-\tau\right){{\dot{q}}_{P t f m}}\left(\tau\right)d\tau}}}\\ {{\dot{x}_{R d t n}={A_{R d t n}}x_{R d t n}+{B_{R d t n}}{{\dot{q}}_{P t f m}}}}\\ {{F_{R d t n}\cong{C_{R d t n}}x_{R d t n}}}\end{array}
+$$
+
+The linear state-space-based wave-excitation model was newly introduced in HydroDyn within OpenFAST with the addition of the linearization functionality for FOWTs. Similar to the linear state-space-based wave-radiation model, the model approximates the infinite integral shown in Eq. (3) with a linear state-space model as shown in Eq. (4), where $F_{E x c t n}$ is the wave-excitation loads under unidirectional waves; $K_{E x c t n}$ is the incident wave-excitation kernels; $\zeta$ is the wave elevation; $\zeta_{c}$ is a time-shifted wave elevation; $x_{E x c t n}$ is the waveexcitation states; and $A_{E x c t n}$ , $B_{E x c t n}$ , and $C_{E x c t n}$ are the matrices of the linear state-space wave-excitation model. Further details on this new linear state-space-based waveexcitation model are provided in Annex A. Within HydroDyn, the user can now select either a Fourier-transform-based waveexcitation model, which is implemented with discrete Fourier transforms (which is not conducive to linearization), or the new linear state-space-based wave-excitation model.
+
+$$
+\begin{array}{c}{{\displaystyle F_{E x c t n}\left(t\right)=\int_{\phantom{\bigg|}E_{E x c t n}}\left(t-\tau\right)\zeta\left(\tau\right)d\tau}}\\ {{\displaystyle\dot{x}_{E x c t n}=A_{E x c t n}x_{E x c t n}+B_{E x c t n}\zeta_{c}\vphantom{\bigg|}}}\\ {{\displaystyle F_{E x c t n}\cong C_{E x c t n}x_{E x c t n}}}\end{array}
+$$
+
+When linear state-space-based wave excitation or wave radiation are enabled in the potential-flow solution, continuoustime states associated with the excitation and radiation damping are included in the linearized HydroDyn model. Linearization is permitted only in still water $\left(\left.\zeta_{c}\right|_{o p}=x_{E x c t n}\right|_{o p}=F_{E x c t n}\right|_{o p}=0\:)^{3}$ and is not permitted when
+
+Fourier-transform-based wave-excitation, convolution-based wave-radiation, wave directional spreading, second-order wave kinematics, or second-order potential-flow loads are enabled. The linearized form of the HydroDyn state and output equations is given by Eq. (5), where ∆x(HD)= $\varDelta x^{(H D)}=\left\{{\varDelta x_{E x c t n}}\atop{\varDelta x_{R d t n}}\right\},$ $\varDelta u^{(H D)}$ includes contributions from $\varDelta\zeta$ and $\varDelta\dot{q}_{P t f m}$ , and ∆y(HD) includes contributions from ∆FExctn and ∆FRdtn.
+
+$$
+\begin{array}{r c l}{{}}&{{\Delta\dot{x}^{(H D)}=A^{(H D)}A x^{(H D)}+B^{(H D)}A u^{(H D)}}}\\ {{}}&{{}}&{{A y^{(H D)}=C^{(H D)}A x^{(H D)}+D^{(H D)}A u^{(H D)}}}\end{array}
+$$
+
+The continuous-state matrix, input matrix, continuous-state output matrix, and the input-transmission matrix are the Jacobians of the state and output equations relative to the states and inputs about the OP, computed analytically for the Jacobians of the state equations and numerically via a centraldifference perturbation technique for the Jacobians of the output equations as shown in Eq. (6), where $X^{(H D)}$ is the continuous-state functions and $Y^{(H D)}$ is the output functions of HydroDyn. For inputs that are rotations in 3D (i.e., $u=A$ and $\varDelta u=\varDelta\vec{\theta}$ ), it is implied that $u\big|_{o p}^{\left(H D\right)}+\varDelta u^{\left(H D\right)}$ uop , $u\big|_{o p}^{\left(H D\right)}-\varDelta u^{\left(H D\right)}$ , and $2{\varDelta}u^{\left(H D\right)}\;\;\;\;\;\;\mathrm{are}\;\;\;\;\;\mathrm{written}\;\;\;\;\;\mathrm{as}\;\;\;\;\;\;{\varLambda}_{\left\vert{\vphantom{\Biggl(}}^{H D\right)}\right.}^{\left(H D\right)}\;f_{A A}\left(A{\vec{\theta}}^{\left(H D\right)}\right),$ $A\big|_{o p}^{(H D)}\,f_{A A}\big(-\varDelta\vec{\theta}^{(H D)}\big)$ , and $2\varDelta\vec{\theta}^{(H D)}$ , respectively. In Eq. (6b), the column order is dictated by the order of HydroDyn inputs, and $\boldsymbol{\theta}$ are appropriately sized zero matrices. For the numerically computed Jacobians, the default perturbation sizes are hard-coded within HydroDyn (but can be customized by recompiling) and are ${\approx}0.035\mathrm{~m~}$ for translational states, $2^{\circ}$ for rotational states and inputs, and fractions of the water depth for translational inputs. The Jacobians contain the linearized contributions from 1) state-space-based wave excitation, 2) hydrodynamic added mass, 3) state-space-based wave-radiation damping, 4) hydrostatic restoring, and 5) linearized viscous drag. The latter is found by numerically linearizing the viscousdrag term from the relative form of Morison’s equation in the strip-theory solution (plus member-end effects) as in Eq. (6), not based on a stochastic linearization method (e.g., [7]).
+
+The $\mathrm{MAP++}$ (MAP) module has constraint (algebraic) states, there are no restrictions to linearization, and the linearized form is given by $\varDelta y^{\left(M A P\right)}=D^{\left(M A P\right)}\varDelta u^{\left(M A P\right)}$ . The input-transmission matrix is the Jacobian of the output equations relative to the inputs about the OP, computed numerically via a central-difference perturbation technique as shown in Eq. (7), where $Y^{(M A P)}$ is the output functions of $\mathrm{MAP++}$ . For inputs that are rotations in 3D (i.e., $u=A$ and $\begin{array}{r l r}&{\begin{array}{r l r}{\boldsymbol{\varDelta}u=\boldsymbol{\varDelta}\vec{\theta}\,\,\mathrm{)},\quad\mathrm{~it}\quad\mathrm{~is~}\quad\mathrm{~implied~}\quad\mathrm{~that~}\quad u\bigl|_{o p}^{\big(M A P\big)}+\varDelta u^{\big(M A P\big)}\,,}\end{array}}\\ &{\begin{array}{r l r}{u_{o p}^{\big(M A P\big)}-\varDelta u^{\big(M A P\big)}\,,\quad}&{\mathrm{and}\quad}&{\mathrm{2}\varDelta u^{\big(M A P\big)}\quad}&{\mathrm{are~}\quad\mathrm{~written~}\quad\mathrm{~as~}}\end{array}}\\ &{\begin{array}{r l r}{\boldsymbol{\varLambda}_{o p}^{\big(M A P\big)}\,f_{\_{\Delta\boldsymbol{A}}}\left(\boldsymbol{\varDelta}\vec{\theta}^{\big(M A P\big)}\right),\quad}&{\boldsymbol{\varLambda}_{\big|o p}^{\big(M A P\big)}\,f_{\_{\Delta\boldsymbol{A}}}\left(-\varDelta\vec{\theta}^{\big(M A P\big)}\right),\quad}&{\mathrm{and}}\end{array}}\end{array}$ shown in Eqs. (6) and (7) of [2]. This difference in implementation was needed in $\mathrm{MAP++}$ because the residual of constraint equations, $Z^{(M A P)}$ , is not readily available internally within MAP $^{++}$ . The default perturbation sizes are hard-coded within $\mathrm{MAP++}-$ but can be customized by recompiling—and are fractions of the water depth for the inputs. To obtain an accurate Jacobian, it is important that the perturbation sizes are not too large to make a taut mooring line go slack or a slack line go taut, but not too small to prevent the tolerance used to solve the implicit constraint equations from effecting the numerically computed derivative.
+
+$2\varDelta\vec{\theta}^{(M A P)}$ , respectively. The Jacobian contains the linearized contribution from mooring restoring.
+
+# MODULE-TO-MODULE, INPUT-OUTPUT COUPLING RELATIONSHIPS LINEARIZATION
+
+$$
+\begin{array}{r l}&{\mathcal{A}^{(l t D)}=\frac{\hat{\mathcal{A}}\mathcal{X}}{\hat{\mathcal{Z}}\mathcal{X}}\Bigg|_{\varphi=\ D}^{(l t D)}=\Bigg[\frac{\mathcal{A}_{\hat{X}\kappa\sin\varphi}}{\mathcal{A}_{\hat{X}\kappa\sin\varphi}}\Bigg]}\\ &{\mathcal{B}^{(l t D)}=\frac{\hat{\mathcal{B}}\mathcal{X}}{\hat{\mathcal{C}}\mathcal{M}}\Bigg|_{\varphi=\ D}^{(l t D)}=\Bigg[\theta\quad\theta\quad\theta\quad\theta\quad\beta_{E\kappa\omega}\Bigg]}\\ &{\mathcal{C}^{(l t D)}=\frac{\hat{\mathcal{A}}\mathcal{Y}}{\hat{\mathcal{Z}}\mathcal{X}}\Bigg|_{\varphi=\ }^{(l t D)}=\frac{\displaystyle{\cal Y}\left(x\right)_{\varphi}+\varDelta x\,u_{[l]_{\varphi}}\,,I[\mathrm{i}_{\varphi}]}{24\Delta\pi}-\gamma\left(x\right)_{\varphi}-\varDelta x\,u_{[l_{\varphi}}\,,I_{\log}^{\prime}\right)\Bigg|^{(l t D)}}\\ &{\mathcal{D}^{(l t D)}=\frac{\hat{\mathcal{B}}\mathcal{Y}}{\hat{\mathcal{C}}\mathcal{M}}\Bigg|_{\varphi=\varphi}^{(l t D)}=\frac{\displaystyle{\cal Y}\left(x\right)_{\varphi}\,,\,u_{[l_{\varphi}}\,,\beta_{\textup{d}^{\prime}}\,+\,\mathrm{d}u_{t},I_{\varphi}]}{24\Delta\mu}\gamma\left(x\right)_{\varphi}\cdot u_{[l_{\varphi}}^{\prime}\,,-\,\mathrm{d}u_{t},I_{\varphi]}\Bigg|^{(l t D)}}\end{array}
+$$
+
+Most module inputs and outputs in OpenFAST reside on spatial boundaries, which in the FAST modularization framework are defined in terms of a mesh. The module-tomodule, input-output coupling relationships in the OpenFAST glue code are algebraic, and include spatial mesh-to-mesh mapping. The general approach to linearizing the mesh mapping within the module-to-module, input-output coupling relationships within OpenFAST, including rotations, is detailed in [2].
+
+The linearized input-output transformation functions, $U$ , are given by Eq. (8) from [2] or Eq. (15) from [4], repeated in Eq. (8) for convenience.
+
+$$
+D^{(M P)}=\frac{d Y}{d u}\Bigg|_{o p}^{\left(M A P\right)}=\frac{Y\left(z\big|_{o p}+{\varDelta z},u\big|_{o p}+{\varDelta u},t\big|_{o p}\right)-Y\left(z\big|_{o p}-{\varDelta z},u\big|_{o p}-{\varDelta u},t\big|_{o p}\right)}{2{\varDelta u}}\Bigg|^{\left(M P\right)}
+$$
+
+While $\mathrm{MAP++}$ and AeroDyn are both modules with constraint states, their numerical linearization approach differs. Equation (7) involves the total derivative of the $\mathrm{MAP++}$ output equations relative to the inputs about the OP, where it is implied that the constraint perturbations, ∆z(MAP) and $-\varDelta z^{(M A P)}$ , are tied to the input perturbations, $\varDelta u^{(M A P)}$ and $-\varDelta u^{\left(M A P\right)}$ , by solving the implicit constraint equations, $0=Z\Bigl(z\big|_{o p}+{\cal A}z,u\big|_{o p}+{\cal A}u,t\big|_{o p}\Bigr)\Bigr|^{(M A P)}$ and $0=Z\Bigl(z\big|_{o p}-varDelta z,u\big|_{o p}-\varDelta u,t\big|_{o p}\Bigr)\Bigr|^{(M A P)}$ , respectively, essentially internally eliminating the constraint states within $\mathrm{MAP++}$ during the linearization process. This is in contrast to the approach taken to linearize AeroDyn, whereby the Jacobians $\left.\frac{\partial Z}{\partial z}\right|_{o p}^{\left(A D\right)},\;\;\left.\frac{\partial Z}{\partial u}\right|_{o p}^{\left(A D\right)},\;\;\left.\frac{\partial Y}{\partial z}\right|_{o p}^{\left(A D\right)},\;\;\mathrm{and}\;\;\left.\frac{\partial Y}{\partial u}\right|_{o p}^{\left(A D\right)}$ are computed separately and algebraically manipulated to compute D as
+
+$$
+{\cal{O}}=\frac{\partial U}{\partial\tilde{u}}\Bigg\vert_{o p}\,\varDelta u+\frac{\partial U}{\partial y}\Bigg\vert_{o p}\,\varDelta y\,\,\mathrm{with}\,\left\vert\frac{\partial U}{\partial\tilde{u}}\right\vert_{o p}\Bigg\vert\neq0
+$$
+
+As is evident from Table 1, the InflowWind, ServoDyn, ElastoDyn, AeroDyn, BeamDyn, HydroDyn, and $\mathrm{MAP++}$ modules were developed so that for the most part—other than mapping between independent spatial discretizations—the input of one module equals the output of another. It follows
+
+that
+
+$$
+U=\left\{U^{\left(J W\right)}\right\},\qquad\quad A u=\left[\begin{array}{c}{{\left(U^{\left(J W\right)}\right)}}\\ {{U^{\left(S\times J D\right)}}}\\ {{U^{\left(E D\right)}}}\\ {{U^{\left(B D\right)}}}\\ {{\left(U^{\left(J D\right)}\right)}}\\ {{U^{\left(4D\right)}}}\\ {{U^{\left(H D\right)}}}\\ {{U^{\left(M D\right)}}}\\ {{U^{\left(M A P\right)}}}\end{array}\right],\qquad\begin{array}{c}{{\left[A u^{\left(J W\right)}\right]}}\\ {{\left(A u^{\left(S r D\right)}\right)}}\\ {{A u^{\left(B D\right)}}}\\ {{A u^{\left(B D\right)}}}\\ {{\left(A u^{\left(A D\right)}\right)}}\\ {{\left(A u^{\left(A D\right)}\right)}}\\ {{\left(A u^{\left(A D\right)}\right)}}\end{array}\right],
+$$
+
+
+
+from Eq. (8) for these seven modules are given by Eq. (9), where $I$ and $\boldsymbol{\theta}$ are appropriately sized identity and zero matrices and the sub-Jacobian matrices are composed of $I\mathrm{~s~}$ , $\theta\;\mathbf{s}.$ , and the linearized matrices from the mapping transfers given in [2].
+
+
+
+Because of their large size, the sub-Jacobian matrices are not shown here, but are described qualitatively instead:
+
+The first, second, fourth, and fifth equations of Eq. (8) for InflowWind inputs, ${\boldsymbol{\theta}}={\boldsymbol{U}}^{(I J W)}$ , ServoDyn inputs, $\boldsymbol{\theta}=\boldsymbol{U}^{(S r v D)}$ , BeamDyn inputs, $\boldsymbol{{\cal O}}=\boldsymbol{{U}}^{(B D)}$ , and AeroDyn inputs, $0=U^{(A D)}$ , are described in [2,3]. The third equation of Eq. (8) for ElastoDyn inputs, $0=U^{(E D)}$ , are also described in [2,3], except that the new floating offshore linearization functionality now also includes terms expressing the contributions of HydroDyn and $\mathrm{MAP++}$ . That is, this equation expresses that the applied point force and moment perturbations distributed along the blades and tower as input to ElastoDyn are derived from the aerodynamicapplied line (per-unit length) force and moment perturbations distributed along the blades and tower as output from AeroDyn. This linearized load-mapping transfer also depends on the translational-displacement perturbations of analysis nodes along the blades and tower output from ElastoDyn. And when BeamDyn is enabled, point force and moment perturbations on the hub as input to ElastoDyn are derived from the bladeroot reaction point force and moment output from BeamDyn, which also depends on the translationaldisplacement perturbation of the hub reference point output from ElastoDyn. Additionally, the blade-pitchangle-commands, nacelle-yaw-moment, and generator-torque perturbations as input to ElastoDyn are derived from the equivalent outputs from ServoDyn. For the floating offshore functionality, point force and moment perturbations on the platform as input to ElastoDyn are derived from the hydrodynamic-applied line (per-unit length) and point force and moment perturbations distributed along the floating platform as output from HydroDyn and the reaction point forces (tensions) lumped at each fairlead as output from $\mathrm{MAP++}$ , which also depends on the translational-displacement perturbation of the platform reference point output from ElastoDyn.
+
+The sixth equation of Eq. (8) for HydroDyn inputs, $O=U^{(H D)}$ expresses that the translational displacement, orientation, translational and rotational velocity, and translational and rotational acceleration perturbations of analysis nodes distributed along the floating platform as input to HydroDyn are derived from motion outputs from ElastoDyn at the platform reference point. The seventh equation of Eq. (8) for $\mathrm{MAP++}$ inputs, $\boldsymbol{\theta}=\boldsymbol{U}^{(M A P)}$ expresses that the translational displacement perturbations of each fairlead as input to $\mathrm{MAP++}$ are derived from the motion outputs from ElastoDyn at the platform reference point.
+
+The Jacobian, $\left.\frac{\partial{\cal U}}{\partial\tilde{u}}\right|_{o p}$ , has ones along its entire diagonal and it is easily shown that its determinant from Eq. (8) is nonzero.
+
+# FINAL MATRIX ASSEMBLY
+
+Once all individual modules and input-output relationships are linearized about the OP, the linearized model of the complete coupled system can be assembled. Linearization of the full-system model produces a linear state-space model representation of the complete nonlinear system about the OP, including the influence of system state and input perturbations on the system response and outputs. The general linearized form of the complete coupled system is given by Eqs. (18) and form for the OpenFAST features linearized to date (without discrete-time states and with ElastoDyn, BeamDyn, and HydroDyn as the only modules with continuous-time states) is given by Eq. (10), where $\boldsymbol{\varDelta u}^{+}$ is the additional input perturbations (explained further in [4]).
+
+$$
+\begin{array}{r}{\varDelta\dot{x}=A A x+B\varDelta u^{+}}\\ {\varDelta y=C\varDelta x+D\varDelta u^{+}}\end{array}
+$$
+
+The full-system state-space matrices are given in Eq. (11), where $G\bigr|_{o p}$ —explained more in [4] for the general case—for the OpenFAST linearization to date is given by Eq. (12). The matrix $G\Big\vert_{o p}$ has ones along its entire diagonal and it is easily shown that its determinant from Eq. (12) is nonzero, which means that the matrix inverse, $\left[G\big|_{o p}\right]^{-I}$ from Eq. (11), exists and is bounded in the neighborhood around the OP.
+
+The input-transmission matrices impact all matrices of the linearized coupled system, highlighting the important role played by direct feedthrough of input to output in the coupled system response. For example, while the continuous-state matrix of ElastoDyn, $\boldsymbol{A}^{(E D)}$ , contains mass, stiffness, and damping only directly associated with the structural model, the full-system continuous state matrix, $A$ , contains mass, stiffness, and damping associated with coupled aero-hydroservo-elastics, including hydrodynamic added mass, waveradiation and viscous damping, and hydrostatic and mooring restoring for FOWTs.
+
+When the linearized full-system matrices $A\,,\,\,B\,,\,\,C$ , and $D$ are exported to a file by OpenFAST, the additional input perturbations, ${\varDelta}{u}^{+}$ , can be chosen by the user to be: 1) the inputs of all modules, 2) none of the module inputs (removing $B$ and $D$ from the file), or 3) a standard subset of these inputs, which include the standard wind turbine control inputs of nacelle-yaw moment, generator torque, and blade-pitchangle commands (both independent and rotor-collective); the standard wind-inflow disturbances of horizontal wind speed, power-law shear exponent, and wind-propagation direction; and the standard incident-wave disturbance of wave elevation. Likewise, the output perturbations, $_{\varDelta y}$ , can be chosen by the user to be: 1) the outputs of all modules, 2) none of the module outputs (removing $C$ and $D$ from the file), or 3) only the subset of output variables selected by the user through the OpenFAST module input file(s). Regardless of what the user selects to be exported to a file, all of the module inputs and outputs are used to form the linearized full-system matrices in Eq. (11), but only a subset of these matrices are exported based on the user selection.
+
+$$
+{\left[\begin{array}{l l l l l l l l}{\left|{\begin{array}{l l l l l l l l l}{\alpha}&&{\beta^{\operatorname{aff}}}&&&{0}&&{0}&{\ddots}&{0}\\ {\alpha}&{\beta^{\operatorname{aff}}}&{0}&&{0}&&{\beta^{\operatorname{aff}}\left|{\vphantom{\begin{array}{l l l l l l l l}{0}}&&{0}&&{0}&&{0}\\ {0}&&{\gamma^{\operatorname{aff}}}&{0}&&{0}&&{0}&{0}\\ {0}&{0}&&{\gamma^{\operatorname{aff}}}&{\left|{\begin{array}{l l l l l l l l}{\alpha}&{0}&&{\beta^{\operatorname{aff}}}&&{0}&&{0}\\ {0}&&{0}&&{0}&&{\theta^{\operatorname{aff}}}&{\left|{\vphantom{\begin{array}{l l l l l l l l}{{\ddots}}&{0}&&{0}&{0}\\ {0}&&{0}&&{0}&{0}&{\ddots}&&{0}\\ {0}&&{0}&&{0}&&{\gamma^{\operatorname{aff}}}&{\left|{\vphantom{\begin{array}{l l l l l l l l l}{0}}&&{0}&&{0}&{0}\\ {0}&&{0}&&{\gamma^{\operatorname{aff}}}&&{\left|{\begin{array}{l l l l l l l l l}{0}&&{0}&&{0}&{0}\\ {0}&&{0}&&{0}&{\ddots}&&{0}\\ {0}&&{0}&&{0}&{0}&&{0}&{\gamma^{\operatorname{aff}}}&{\left|{\vphantom{\begin{array}{l l l l l l l l l}{0}&{0}&{0}&{0}&{0}\\ {0}&&{0}&{0}&{\ddots}&&{0}\\ {0}&&{0}&{0}&&{0}&{0}&{\gamma^{\operatorname{aff}}}&{\left|{\vphantom{\begin{array}{l l l l l l l l l}{0}&{0}&{0}&{0}&{0}\\ {0}&&{0}&{0}&{0}&{\gamma^{\operatorname{aff}}}&&{\left|{\begin{array}{l l l l l l l l l}{0}&&{0}&{0}&{0}\\ {0}&&{0}&{0}&{0}&{0}\\ {0}&&{0}&{\gamma^{\operatorname{aff}}}&{0}&&{0}&{0}\\ {0}&{0}&&{0}&{0}&{
+$$
+
+
+
+# CONCLUSIONS
+
+Linearization of the underlying nonlinear wind-system equations is often important for understanding the system response and exploiting well-established methods and tools for analyzing linear systems. This paper has presented the development of the new linearization functionality of the opensource engineering tool OpenFAST for FOWTs, as well as the concepts and mathematical background needed to understand and apply it. Details have been presented on finding an OP, linearizing the underlying nonlinear equations of each module about the OP, linearizing the module-to-module input-output coupling relationships about the OP, and combining all linearized matrices into the full-system linear state-space model. The new linearization functionality enables the contributions from state-space-based wave excitation, hydrodynamic added mass, state-space-based wave-radiation damping, hydrostatic restoring, linearized viscous drag, and linearized mooring restoring for FOWTs to be included in the full linearized system.
+
+Unfortunately, the implementation at the time of this writing has not yet been completed enough to produce results. Results will be presented in future work to highlight the functionality and verify the implementation (e.g., for a sample case whereby the natural frequencies and damping of the NREL 5-MW baseline wind turbine atop the OC3-Hywind spar buoy will be calculated as a function of rotational speed and presented in Campbell-diagram form).
+
+Beyond the developments implemented here, potential future OpenFAST linearization enhancements include: 1) introducing routines to find a static equilibrium, steady-state, or periodic steady-state condition for improved OP determination; 2) enabling the linearization of the tuned-mass-damper (TMD)
+
+models and external user-specified controllers within ServoDyn for improved controls development and analysis; 3) implementing a form of unsteady airfoil aerodynamics, including dynamic stall, within AeroDyn amendable to linearization (e.g., from [8]) and including that model within the linearization of AeroDyn for advanced aerodynamic and stability analysis; 4) linearizing the offshore functionality of OpenFAST for bottom-fixed offshore wind systems; and 5) linearizing new features as the OpenFAST models are further developed in the future going forward.
+
+# ACKNOWLEDGMENTS
+
+The authors would like to thank Bonnie Jonkman of Envision Energy USA, Ltd. for her contributions to the development of the FAST modularization framework and to implementation of the land-based linearization functionality.
+
+This work was authored by the National Renewable Energy Laboratory, operated by the Alliance for Sustainable Energy, LLC, for the U. S. Department of Energy (DOE) under Contract No. DE-AC36-08GO28308. Funding was provided by the U.S. Department of Energy Office of Energy Efficiency and Renewable Energy. The views expressed in the article do not necessarily represent the views of the U.S. Department of Energy or the U.S. government. The U.S. government retains and the publisher, by accepting the article for publication, acknowledges that the U.S. government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for U.S. government purposes.
+
+# REFERENCES
+
+[1] Web page https://nwtc.nrel.gov/OpenFAST (accessed May 7, 2018).
+
+[2] Jonkman, J. M. and Jonkman, B. J. “FAST modularization framework for wind turbine simulation: full-system linearization.” Journal of Physics: Conference Series, The Science of Making Torque from Wind (TORQUE 2016), 5–7 October 2016, Munich, Germany [online journal]. 082010. Vol. 753, 2016. URL: http://iopscience.iop.org/article/10.1088/1742- 6596/753/8/082010/pdf; NREL/CP-5000-67015. Golden, CO: National Renewable Energy Laboratory.
+[3] Jonkman, J. M.; Jonkman, B. J.; and Platt, A. “FullSystem Linearization for Wind Turbines with Aeroelastically Tailored Rotor Blades in OpenFAST.” (in preparation).
+[4] Jonkman, J. M. “The New Modularization Framework for the FAST Wind Turbine CAE Tool.” $5l^{s t}$ AIAA Aerospace Sciences Meeting including the New Horizons Forum and Aerospace Exposition, 7–10 January 2013, Grapevine (Dallas/Ft. Worth Region), TX [online proceedings]. URL: http://arc.aiaa.org/doi/pdf/10.2514/6.2013-202. AIAA-2013-0202. Reston, VA: American Institute of Aeronautics and Astronautics, January 2013; NREL/CP5000-57228. Golden, CO: National Renewable Energy Laboratory.
+[5] Duarte, T.; Alves, M.; Jonkman, J.; and Sarmento, A. “State-Space Realization of the Wave-Radiation Force within FAST.” $32^{n d}$ International Conference on Ocean, Offshore, and Arctic Engineering (OMAE2013), 9–14 June 2013, Nantes, France [DVD-ROM]. OMAE2013- 10375. Houston, TX: The American Society of Mechanical Engineers (ASME International) Ocean, Offshore and Arctic Engineering (OOAE) Division, June 2013; ISBN 978-0-7918-5542-3; NREL/CP-5000-58099. Golden, CO: National Renewable Energy Laboratory.
+[6] Web page https://nwtc.nrel.gov/SS_Fitting (accessed May 7, 2018).
+[7] Borgman, L. E. “Spectral Analysis of Ocean Wave Forces on Piling.” Journal of the Waterways and Harbors Division, Vol. 93, Issue 2, pp. 129-156, 1967.
+[8] Hansen, M. H.; Guanaa, M.; and Madsen, H. A. A Beddoes-Leishman type dynamic stall model in statespace and indicial formulations. Risø-R-1354 (EN). Roskilde, Denmark: Risø National Laboratory. 2004.
+
+# ANNEX A
+
+# LINEAR STATE-SPACE-BASED WAVE-EXCITATION MODEL
+
+In the frequency domain, the potential-flow-based waveexcitation loads (three forces and three moments) can be expressed for unidirectional waves by Eq. (A.1), where $X\big(\omega,\beta\big)$ is the frequency- $\left(\,\omega\,\right)$ and direction- $\left(\,\beta\,\right)$ dependent wave-excitation loads on the floating platform normalized per unit wave amplitude (derived, e.g., from a potential-flow-based wave-body interaction solver WAMIT) and $\zeta(\omega)$ is the waveelevation spectrum for waves propagating in direction $\beta$ (both are complex-valued to include phases). For time-domain simulations, the spectral wave data and wave direction are model inputs and the phases of the various wave components are selected randomly. The time-domain wave-excitation loads, $F_{E x c t n}\left(t\right)$ , are usually precomputed at model initialization by calculating the inverse Fourier transform shown in Eq. (A.2), where $j=\sqrt{-I}$ is the imaginary number (HydroDyn uses a discrete Fourier transform in place of the continuous-time transform shown).
+
+$$
+F_{E x c t n}\left(\omega\right)=X\big(\omega,\beta\big)\zeta\big(\omega\big)
+$$
+
+$$
+F_{E x c t n}\left(t\right)=\frac{I}{2\pi}\int_{-\infty}^{\infty}X\big(\omega,\beta\big)\zeta\left(\omega\right)\!e^{j\omega t}d\omega
+$$
+
+Equivalently, as shown in [A.1], the wave-excitation load components can be calculated through the infinite convolution integral given by Eq. (3), where the incident wave-excitation kernels derived from Eq. (A.3), represent the incident waveexcitation impulse-response functions (IRF) that depend only on platform parameters. The goal is to use a linear state-space equation to approximate Eq. (3), with the time-domain wave elevation, $\zeta\!\left(t\right)$ , acting as the input, and the time-domain wave-excitation loads, $F_{E x c t n}\left(t\right)$ , as the output.
+
+$$
+K_{E x c t n}\left(t\right)=\frac{I}{2\pi}\int_{-\infty}^{\infty}X\big(\omega,\beta\big)e^{j\omega t}d\omega
+$$
+
+Fitting a linear state-space model to the infinite convolution integral is complicated by the noncausality of this system. This issue has been studied extensively in [A.2] and [A.3], where a time-shifting method is used to arrive at a causal transfer function between wave elevation and wave-excitation loads. We follow the same method used in [A.2] and [A.3] here. Figure A.1 shows an example platform-pitch response (blue curve) resulting from a wave-elevation impulse at time $t=\partial$ s. Clearly, the system is noncausal because its response is nonzero for negative time even though the impulse acts at $t=\partial$ s. The system response for negative time is as large as the response for positive time, so it cannot be ignored.
+
+
+FIGURE A.1. NONCAUSAL AND CAUSAL IMPULSERESPONSE FUNCTIONS.
+
+An approximately causal IRF is obtained through a timeshifting method. Figure A.1 shows that for all times less than $-t_{c}$ ( $t_{c}\approx I2$ s for this example), the IRF is approximately zero. A time delay, $t_{c}$ , is introduced to shift the original IRF to obtain a “causalised” IRF, $K_{E x c t n_{c}}\left(t\right)=K_{E x c t n_{c}}\left(t-t_{c}\right).$ . Figure A.1 shows that this time-shifted IRF (red) is approximately zero for negative time. Equation (A.4), in terms of $K_{E x c t n_{c}}\left(t\right)$ and $\zeta_{c}\left(t\right)$ , follows directly from Eq. (3) and the time-shift properties of Fourier transforms, where $\zeta_{c}\left(t\right)=\zeta\left(t+t_{c}\right)$ is the wave elevation predicted at time $t_{c}$ into the future. To implement this approach, a wave-prediction model must be used to predict the wave elevation for future times. Within OpenFAST simulations, this is not an issue because the wave elevation is precalculated at model initialization, so the future wave elevations are already available. In practice, the future wave evaluation could be estimated (e.g., by measuring the wave elevation with a buoy and approximating the wave evolution).
+
+$$
+F_{E x c t n}\left(t\right)=\int_{-\infty}^{\infty}K_{E x c t n_{c}}\left(t-\tau\right)\zeta_{c}\left(\tau\right)d\tau
+$$
+
+Now the goal is to fit a linear state-space model to the causalised IRF, $K_{E x c t n_{c}}\left(t\right)$ , based on input $\zeta_{c}\left(t\right).$ . For a statespace system of the form shown in Eq. (4), it can be shown that the IRF is $C_{E x c t n}\ e^{A_{E x c t n}t}\ B_{E x c t n}$ , where $e^{\b{A}_{E x c t n}t}$ is the matrixexponential function [A.4]. The goal is to determine $A_{E x c t n}$ , $B_{E x c t n}$ , and $C_{E x c t n}$ so that $C_{E x c t n}\ e^{A_{E x c t n}t}\ B_{E x c t n}$ approximates $K_{E x c t n_{c}}\left(t\right)$ in terms of some error norm. This is usually done by minimizing a cost function, such as given by Eq. (A.5), where $G\big(t_{k}\big)$ is a weighting function to be chosen, evaluated at specific sampling times, $t_{k}$ , for $k=I,2,\dotsc,m$ . At least $m$ discrete values of KExctn a re assumed to be known either from theory or experiment.
+
+$$
+Q=\sum_{k=I}^{m}G\!\left(t_{k}\right)\!\left[K_{E x c t n_{c}}\left(t_{k}\right)\!-C_{E x c t n}~e^{A_{E x c t n}t}~B_{E x c t n}\right]^{2}
+$$
+
+Several methods exist for performing this systemidentification step. For this work, we rely on the use of a linear state-space realization of the impulse response using Hankel singular value decomposition and MATLAB’s SystemIdentification-Toolbox command imp2ss [A.5]. The state-space model structure is selected so that the infinite-frequency limit of the identified model in the frequency domain is zero, forcing the output transmission $(\,D\,)$ matrix in the general state-space description to be equal to the zero matrix [A.5]. A balanced model-truncation method is used to arrive at a low-order statespace approximation for the wave-excitation loads.
+
+This general method for generating a linear state-space model of the wave excitation for inclusion into OpenFAST is being developed and programmed into a MATLAB script. Further details, with example cases and a user’s guide for this script will be described in a future publication.
+
+[A.1] Jonkman, J. M. Dynamics Modeling and Loads Analysis of an Offshore Floating Wind Turbine. Ph.D. Thesis. Department of Aerospace Engineering Sciences, University of Colorado, Boulder, CO, 2007; NREL/TP500-41958. Golden, CO: National Renewable Energy Laboratory.
+[A.2] Yu, Z. and Falnes, J. “State-space Modeling of a Vertical Cylinder in Heave.” Applied Ocean Research, Vol. 17, Issue 5, pp. 265-275, 1995.
+[A.3] Lemmer, F., Raach, S., Schlipf, D., and Cheng, P. “Parametric Wave Excitation Model for Floating Wind Turbines.” Energy Procedia, Vol. 94, 2016, pp. 290-305; $I3^{t h}$ Deep Sea Offshore Wind R&D Conference, EERA DeepWind’2016, 20-22 January 2016, Trondheim, Norway; DOI: 10.1016/j.egypro.2016.09.186.
+[A.4] Ogata, K. Modern Control Engineering, Englewood Cliffs, NJ: Prentice-Hall Inc., 1990.
+[A.5] Ljung, L. and Singh, R. “Version 8 of the MATLAB System Identification Toolbox.” $I6^{t h}$ IFAC Symposium on System Identification, Brussels, Belgium. ISBN 9783902823069; pp. 1826-1832, 2012; doi:10.3182/20120711-3-BE-2027.00061.
\ No newline at end of file
diff --git a/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/Jonkman-Full-system linearization for floating_origin.pdf b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/Jonkman-Full-system linearization for floating_origin.pdf
new file mode 100644
index 0000000..75a799a
Binary files /dev/null and b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/Jonkman-Full-system linearization for floating_origin.pdf differ
diff --git a/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/228dd88a114ae163dad8ffb4be4a2a74d82c5bee4a868bbce9f447707580824f.jpg b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/228dd88a114ae163dad8ffb4be4a2a74d82c5bee4a868bbce9f447707580824f.jpg
new file mode 100644
index 0000000..5264513
Binary files /dev/null and b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/228dd88a114ae163dad8ffb4be4a2a74d82c5bee4a868bbce9f447707580824f.jpg differ
diff --git a/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/22dc794cd6fdc6791a3412b34d67e99749275b20bb63eee47ea19d5908a94158.jpg b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/22dc794cd6fdc6791a3412b34d67e99749275b20bb63eee47ea19d5908a94158.jpg
new file mode 100644
index 0000000..e5a4948
Binary files /dev/null and b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/22dc794cd6fdc6791a3412b34d67e99749275b20bb63eee47ea19d5908a94158.jpg differ
diff --git a/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/63a387a5c4d629058ef1038d42302921b057bc3c4140965b126080002098ca7c.jpg b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/63a387a5c4d629058ef1038d42302921b057bc3c4140965b126080002098ca7c.jpg
new file mode 100644
index 0000000..e1733be
Binary files /dev/null and b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/63a387a5c4d629058ef1038d42302921b057bc3c4140965b126080002098ca7c.jpg differ
diff --git a/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/a831b7ac45c7ef3e2ffe228fb1164fc895012b24d942fac4f79d60ba618f2485.jpg b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/a831b7ac45c7ef3e2ffe228fb1164fc895012b24d942fac4f79d60ba618f2485.jpg
new file mode 100644
index 0000000..393f154
Binary files /dev/null and b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/a831b7ac45c7ef3e2ffe228fb1164fc895012b24d942fac4f79d60ba618f2485.jpg differ
diff --git a/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/e3cf5199761175d226cc33794ee9ebb7a89b2f9cf0f39528b2eaeef3d2ad016d.jpg b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/e3cf5199761175d226cc33794ee9ebb7a89b2f9cf0f39528b2eaeef3d2ad016d.jpg
new file mode 100644
index 0000000..00b03a5
Binary files /dev/null and b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/e3cf5199761175d226cc33794ee9ebb7a89b2f9cf0f39528b2eaeef3d2ad016d.jpg differ
diff --git a/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/f47cb8194eff840ddeca7a7167915ca9ff6fca5b6c65e89e27039e48f46917cc.jpg b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/f47cb8194eff840ddeca7a7167915ca9ff6fca5b6c65e89e27039e48f46917cc.jpg
new file mode 100644
index 0000000..9c23f64
Binary files /dev/null and b/线性化求解器/参考文献/Jonkman-Full-system linearization for floating/auto/images/f47cb8194eff840ddeca7a7167915ca9ff6fca5b6c65e89e27039e48f46917cc.jpg differ
diff --git a/线性化求解器/参考文献/Skjoldan-Implicit Floquet analysis of wind tur/auto/Skjoldan-Implicit Floquet analysis of wind tur.md b/线性化求解器/参考文献/Skjoldan-Implicit Floquet analysis of wind tur/auto/Skjoldan-Implicit Floquet analysis of wind tur.md
new file mode 100644
index 0000000..9c5c302
--- /dev/null
+++ b/线性化求解器/参考文献/Skjoldan-Implicit Floquet analysis of wind tur/auto/Skjoldan-Implicit Floquet analysis of wind tur.md
@@ -0,0 +1,377 @@
+# Implicit Floquet analysis of wind turbines using tangent matrices of a non-linear aeroelastic code
+
+P. F. Skjoldan1 and M. H. Hansen2
+
+1 Loads, Aerodynamic and Control, Siemens Wind Power A/S, DK-2630 Taastrup, Denmark
+2 Wind Energy Division, National Laboratory for Sustainable Energy, Risø DTU, DK-4000 Roskilde, Denmark
+
+# ABSTRACT
+
+The aeroelastic code BHawC for calculation of the dynamic response of a wind turbine uses a non-linear finite element formulation. Most wind turbine stability tools for calculation of the aeroelastic modes are, however, based on separate linearized models. This paper presents an approach to modal analysis where the linear structural model is extracted directly from BHawC using the tangent system matrices when the turbine is in a steady state. A purely structural modal analysis of the periodic system for an isotropic rotor operating at a stationary steady state was performed by eigenvalue analysis after describing the rotor degrees of freedom in the inertial frame with the Coleman transformation. For general anisotropic systems, implicit Floquet analysis, which is less computationally intensive than classical Floquet analysis, was used to extract the least damped modes. Both methods were applied to a model of a three-bladed $2.3\;\mathrm{MW}$ Siemens wind turbine model. Frequencies matched individually and with a modal identification on time simulations with the non-linear model. The implicit Floquet analysis performed for an anisotropic system in a periodic steady state showed that the response of a single mode contains multiple harmonic components differing in frequency by the rotor speed. Copyright $\copyright$ 2011 John Wiley & Sons, Ltd.
+
+# KEYWORDS
+
+modal analysis; Floquet analysis; rotor dynamics
+
+# Correspondence
+
+P. F. Skjoldan, Loads, Aerodynamic and Control, Siemens Wind Power A/S, Dybendalsvænget 3, DK-2630 Taastrup, Denmark. E-mail: peter.skjoldan@siemens.com
+
+Received 26 June 2010; Revised 7 October 2010; Accepted 12 February 2011
+
+# 1. INTRODUCTION
+
+Today, advanced non-linear finite element codes1–3 are routinely used for load calculations on wind turbines. Most wind turbine stability tools for calculation of the aeroelastic modes are, however, based on separate linearized models. Stability analysis can be divided into three steps: first, a calculation of the steady state; then, a linearization of the equations of motion about the steady state and last, a modal analysis to extract modal frequencies, damping and mode shapes. This paper presents an approach to structural modal analysis applicable to any periodic steady state where the linearization is obtained directly from the non-linear wind turbine aeroelastic code BHawC.3
+
+The equations of motion for a wind turbine operating at a constant mean rotor speed contain periodic coefficients, preventing direct eigenvalue analysis of the system. Most recent wind turbine stability tools $4\mathrm{-}7$ incorporate the Coleman transformation, also known as the multiblade coordinate transformation, which describes the rotor degrees of freedom in the inertial frame. This transformation eliminates the periodic coefficients if the system is isotropic, i.e. the rotor consists of identical symmetrically mounted blades, and the environment conditions are symmetric. Floquet analysis is, however, applicable to anisotropic systems and any periodic steady state. It requires integration of the equations of motion over a period of rotor rotation, as many times as there are state variables in the system. Because of the computational burden of this approach, it has only been applied to reduce or simplify wind turbine models with a limited number of degrees of freedom.8–10 One way to reduce the computation time is to use the Fast Floquet Theory11 where only one third of the integrations are necessary for a three-bladed isotropic rotor. Another way is to use implicit Floquet analysis12 where the least damped modes can be extracted after a limited number of integrations.
+
+Stol et al.13 compare the Floquet analysis with the Coleman transformation approach applied to a periodic steady state, where the remaining periodic coefficients are eliminated by averaging and find small differences in modal frequencies and damping, concluding that it is not necessary to use Floquet analysis.
+
+Another approach to modal analysis is system identification,14–16 which operates on the response from numerical simulations or measurements, and no knowledge of the system equations is needed to extract the modal properties. The accuracy of the methods is, however, limited and depends on the chosen excitation.
+
+In this paper, tangent matrices for mass, damping and stiffness are extracted from the aeroelastic code BHawC. If the system is isotropic and the steady state is stationary, the Coleman transformation is applied before extracting the modal parameters by eigenvalue analysis. For an anisotropic system, implicit Floquet analysis is used for the modal analysis. When the system is isotropic, the response of a single mode contains a single harmonic component for tower degrees of freedom and up to three components for the blades. The response of a single mode in the anisotropic system on both blades and tower contains multiple harmonic components differing in frequency by the rotor speed.
+
+Section 2 of this paper describes the BHawC model, and Section 3 explains the methods for modal analysis, the Coleman transformation approach, the implicit Floquet analysis and also the partial Floquet analysis, a system identification technique. In Section 4, the methods are applied to a model of a wind turbine. Section 5 discusses the approaches, and Section 6 concludes the paper.
+
+# 2. STRUCTURAL MODEL
+
+The BHawC wind turbine aeroelastic code3 is based on a structural finite element model sketched in Figure 1, where the main structural parts, tower, nacelle, shaft, hub and blades, are modelled as two-node 12-degrees of freedom Timoshenko beam elements. The code uses a corotational formulation, where each element has its own coordinate system that rotates with the element. The elastic deformation is described in the element frame, whereas the movement of the element coordinate system accounts for rigid body motion. In this way, a geometrically non-linear model is obtained using linear finite elements.
+
+The configuration of the system, defined by nodal positions $\pmb{p}$ and orientations $\pmb q$ , nodal velocities $\dot{\pmb u}$ (of both positions and orientations) and nodal accelerations $\ddot{u}$ , must satisfy the equilibrium equation given in global coordinates as
+
+$$
+f_{\mathrm{iner}}(\boldsymbol{p},\boldsymbol{q},\dot{\boldsymbol{u}},\ddot{\boldsymbol{u}})+f_{\mathrm{damp}}(\boldsymbol{q},\dot{\boldsymbol{u}})+f_{\mathrm{int}}(\boldsymbol{p},\boldsymbol{q})=f_{\mathrm{ext}}
+$$
+
+where $f_{\mathrm{iner}},f_{\mathrm{damp}},f_{\mathrm{int}}$ and $\pmb{f}_{\mathrm{ext}}$ are the inertial, damping, internal and external force vectors, respectively, and $\dot{\mathrm{O}}=\mathrm{d}/\mathrm{d}t$ denotes a time derivative. The inertial forces depend on the acceleration of the masses, the damping forces are given by viscous damping, the internal forces are due to elastic forces and the external forces contain the aerodynamic forces.17 To find
+
+
+Figure 1. Sketch of the BHawC model substructures.
+
+this equilibrium configuration, increments of the positions and the orientations $\delta\pmb{u}$ , the velocities $\delta\dot{\pmb{u}}$ and the accelerations $\delta\ddot{\pmb{u}}$ are obtained using Newton–Raphson iteration with the tangent relation obtained from the variation of Equation (1) as
+
+$$
+\mathbf{M}(q)\delta{\ddot{u}}+\mathbf{C}(q,{\dot{u}})\delta{\dot{u}}+\mathbf{K}(p,q,{\dot{u}},{\ddot{u}})\delta u=r
+$$
+
+where M, C and $\mathbf{K}$ are the tangent mass, damping/gyroscopic and stiffness matrices, respectively, and $r=f_{\mathrm{ext}}+f_{\mathrm{iner}}+$ $f_{\mathrm{damp}}\!-\!f_{\mathrm{int}}$ is the residual. The stiffness matrix is composed of constitutive, geometric and inertial stiffness. The orientation $\pmb q$ of the nodes is described by quaternions, also known as the Euler parameters,18 a general four-parameter representation equivalent to a triad, which for node number $i$ is updated as
+
+$$
+\pmb q_{i}:=q u a t(\delta\pmb u_{i,\mathrm{rot}})*\pmb q_{i}
+$$
+
+where $\delta\mathbf{\boldsymbol{u}}_{i,\mathrm{rot}}$ contains three rotations that are assumed infinitesimal and thus commute and where this rotation pseudovector is transformed by the function termed quat into a quaternion, which is used to update the nodal quaternion $\pmb q_{i}$ employing the special quaternion product denoted by $^*$ , which maintains the unity of the quaternion. The nodal positions $\pmb{p}$ , the nodal velocities $\dot{\pmb u}$ and the accelerations $\ddot{u}$ are updated by regular addition of the positional part of $\delta\pmb{u},\,\delta\dot{\pmb{u}}$ and $\delta\ddot{\pmb{u}}$ , respectively. All components in $\pmb{p}$ , $\pmb q$ and $\delta\pmb{u}$ are absolute and described in a global frame.
+
+The present work considers small perturbations in position and orientation ${\bf\delta y}$ , velocity $\dot{\mathbf{y}}$ and acceleration $\ddot{\mathbf{y}}$ to a steady state with constant mean rotor speed $\varOmega$ defined by $(p_{\mathrm{ss}},q_{\mathrm{ss}},\dot{\pmb u}_{\mathrm{ss}},\ddot{\pmb u}_{\mathrm{ss}})$ , the steady state positions, orientations, velocities and accelerations, respectively, all periodic with the rotor period $T=2\pi/\varOmega$ . The linearized equations of motion are obtained from equation (2) at $r\approx\theta$ as
+
+$$
+{\bf M}({q}_{\mathrm{ss}})\ddot{\boldsymbol{y}}+{\bf C}({q}_{\mathrm{ss}},\dot{\boldsymbol{u}}_{\mathrm{ss}})\dot{\boldsymbol{y}}+{\bf K}({p}_{\mathrm{ss}},{q}_{\mathrm{ss}},\dot{\boldsymbol{u}}_{\mathrm{ss}},\ddot{\boldsymbol{u}}_{\mathrm{ss}}){\boldsymbol{y}}=\boldsymbol{\theta}
+$$
+
+where the matrices $\mathbf{M}$ , $\mathbf{C}$ and $\mathbf{K}$ are the $T$ -periodic tangent system matrices that are employed in the modal analysis described in the next section.
+
+# 3. METHODS
+
+In this section, the four methods for modal analysis of structures with rotors are presented.
+
+# 3.1. Coleman approach
+
+The Coleman transformation requires identical degrees of freedom on each blade, and therefore, the equations of motion (equation (4)) in global coordinates were first transformed into substructure coordinates $y_{\mathrm{T}}$ . The transformation is
+
+$$
+\begin{array}{r l}&{\boldsymbol{y}=\mathrm{\mathbf{T}}\boldsymbol{y}_{\mathrm{T}}}\\ &{\mathbf{T}=\mathbf{diag}(\mathbf{I}_{N_{s}},\mathbf{T}_{\mathrm{r}},\mathbf{T}_{\mathrm{b1}},\mathbf{T}_{\mathrm{b2}},\mathbf{T}_{\mathrm{b3}})}\end{array}
+$$
+
+where $\mathbf{T}$ is a block diagonal time-variant matrix composed of the identity matrix $\mathbf{I}_{N_{\mathrm{s}}}$ sized by the number of degrees of freedom of the tower, the nacelle and the drivetrain, $\mathbf{T_{r}}$ transforms the degrees of freedom on the shaft and the hub into a hub centre frame and $\mathrm{T}_{\mathfrak{b}j}$ transforms the degrees of freedom on blade number $j=1,2,3$ into a local frame for blade $j$ . The triads were obtained in the periodic steady state, and thus, $\mathbf{T}$ is $T$ -periodic.
+
+The time-variant transformation into inertial frame coordinates $z$ is
+
+$$
+\begin{array}{r l}&{{\mathbf y}_{\mathrm{T}}={\mathbf B}\,z}\\ &{{\mathbf B}={\textbf d i a g}({\mathbf I}_{N_{\mathrm{s}}},{\mathbf B}_{\mathrm{r}},{\mathbf B}_{\mathrm{b}})}\end{array}
+$$
+
+where $\mathbf{B}_{\mathrm{r}}$ is a simple rotational transformation of the shaft and the hub and $\mathbf{B}_{\mathrm{b}}$ is the Coleman transformation introducing multiblade coordinates for a three-bladed rotor11,19 as
+
+$$
+\mathbf{B}_{\mathrm{b}}=\left[\begin{array}{l l l}{\mathbf{I}_{N_{\mathrm{b}}}}&{\mathbf{I}_{N_{\mathrm{b}}}\cos\psi_{1}}&{\mathbf{I}_{N_{\mathrm{b}}}\sin\psi_{1}}\\ {\mathbf{I}_{N_{\mathrm{b}}}}&{\mathbf{I}_{N_{\mathrm{b}}}\cos\psi_{2}}&{\mathbf{I}_{N_{\mathrm{b}}}\sin\psi_{2}}\\ {\mathbf{I}_{N_{\mathrm{b}}}}&{\mathbf{I}_{N_{\mathrm{b}}}\cos\psi_{3}}&{\mathbf{I}_{N_{\mathrm{b}}}\sin\psi_{3}}\end{array}\right]
+$$
+
+where $\psi_{j}=\varOmega t+2\pi(j-1)/3$ is the mean azimuth angle to blade number $j$ and $N_{\mathrm{b}}$ is the number of degrees of freedom on each blade. The inertial frame coordinate vector
+
+$$
+\boldsymbol{z}=\{y_{\mathrm{s}}^{\mathrm{T}}\,z_{\mathrm{r}}^{\mathrm{T}}\,a_{0}^{\mathrm{T}}\,a_{1}^{\mathrm{T}}\,b_{1}^{\mathrm{T}}\}^{\mathrm{T}}
+$$
+
+contains the untransformed coordinates for tower, nacelle and drivetrain ${\mathfrak{y}}_{\mathrm{s}}$ , the coordinates for shaft and hub $z_{\mathrm{r}}$ measured in a non-rotating frame aligned with the hub and the multiblade symmetric coordinates $\pmb{a}_{0}$ , cosine coordinates $\pmb{a}_{1}$ and sine coordinates $\pmb{b}_{1}$ . The details on how multiblade coordinates describe the motion of a wind turbine rotor in the inertial frame are discussed by Hansen.20,21
+
+The Coleman transformed equations were obtained by first inserting equation (5) into equation (4), then converting it to first order form and lastly introducing the inertial frame transformation in equation (6) as ${\displaystyle y_{\mathrm{T}2}=\mathrm{diag}({\bf B},{\bf B})z_{2}}$ where ${y}_{\mathrm{T}2}=\{{y}^{\mathrm{T}}\,{\dot{{y}}}^{\mathrm{T}}\}^{\mathrm{T}}$ and $z_{2}=\dot{\{z^{\mathrm{T}}\,\tilde{z}^{\mathrm{T}}\}}^{\mathrm{T}}$ are the state v ectors in substructure and ine rtial frames, respectively, with $\tilde{z}=\dot{z}+\bar{\omega}z$ and the constant matrix $\bar{\boldsymbol{\omega}}=\mathbf{B}^{-1}\mathbf{B}$ . The result is
+
+$$
+\begin{array}{r l}&{\dot{z}_{2}=\mathbf{A}_{\mathrm{B}}z_{2}}\\ &{\mathbf{A}_{\mathrm{B}}=\left[\mathbf{-}\mathbf{\bar{\omega}}-\mathbf{\bar{\omega}}\mathbf{\bar{\omega}}_{\mathrm{KB}}\quad\mathbf{-M}_{\mathrm{B}}^{-1}\mathbf{C}_{\mathrm{B}}-\mathbf{\bar{\omega}}\right]}\end{array}
+$$
+
+where $\mathbf{A}_{\mathrm{B}}$ is the Coleman transformed system matrix and
+
+$$
+\begin{array}{r l}&{\mathbf{M}_{\mathrm{B}}=\mathbf{B}^{-1}\mathbf{T}^{\mathrm{T}}\mathbf{M}\mathbf{T}\,\mathbf{B}}\\ &{\mathbf{C}_{\mathrm{B}}=\mathbf{B}^{-1}\mathbf{T}^{\mathrm{T}}(\mathbf{C}\,\mathbf{T}+2\,\mathbf{M}\,\dot{\mathbf{T}})\mathbf{B}}\\ &{\mathbf{K}_{\mathrm{B}}=\mathbf{B}^{-1}\mathbf{T}^{\mathrm{T}}(\mathbf{K}\,\mathbf{T}+\mathbf{C}\,\dot{\mathbf{T}}+\mathbf{M}\,\ddot{\mathbf{T}})\mathbf{B}}\end{array}
+$$
+
+are the Coleman transformed mass, damping/gyroscopic and stiffness matrices, respectively. If the system is isotropic, then $\mathbf{A}_{\mathrm{B}}$ is time-invariant, and a transient solution of equation (9) is
+
+$$
+z_{2}=\mathrm{e}^{\mathbf{A}_{\mathrm{B}}t}z_{2}(0)=\mathbf{V}\mathrm{e}^{\mathbf{A}t}q(0)
+$$
+
+where $\mathbf{A}$ is a diagonal matrix containing the eigenvalues of $\mathbf{A}_{\mathrm{B}}$ , $\mathbf{V}$ contains the corresponding eigenvectors as columns and $\pmb q(0)=\mathbf V^{-1}z_{2}(0)$ are the initial conditions in modal c oordinates. It is assumed th at all eigenvect ors are linearly independen t .
+
+The blade motion given in the inertial frame in equation (11) can be transformed back into the rotating frame using equation (6) as21
+
+$$
+\Gamma,i k=\mathrm{e}^{\sigma_{k}t}\left(A_{0,i k}\cos(\omega_{k}t+\varphi_{0,i k})+A_{\mathrm{BW},i k}\cos\left((\omega_{k}+\Omega)t+\varphi_{j}+\varphi_{\mathrm{BW},i k}\right)+A_{\mathrm{FW},i k}\cos\left((\omega_{k}-\Omega)t-\varphi_{j}+\varphi_{\mathrm{GW},i k}\right)\right),
+$$
+
+where $\varphi_{j}=2\pi(j-1)/3$ and $\sigma_{k}$ and $\omega_{k}$ pare the modal damping and frequency of mode number $k$ , respectively, given by the eigenvalue $\lambda_{k}=\sigma_{k}+\mathrm{i}\omega_{k}$ with $\mathrm{i}=\sqrt{-1}$ . The amplitudes for degree of freedom number $i$ were determined from the components of the eigenvector $\nu_{k}$ gi ven in multiblade coordinates of equation (8) as $A_{0,i k}=|a_{0,i k}|$ and
+
+$$
+\begin{array}{r l}&{A_{\mathrm{BW},i k}=\frac{1}{2}\big((\mathrm{Re}\,(a_{1,i k})+\mathrm{Im}\,(b_{1,i k}))^{2}+(\mathrm{Re}\,(b_{1,i k})-\mathrm{Im}\,(a_{1,i k}))^{2}\big)^{1/2}}\\ &{A_{\mathrm{FW},i k}=\frac{1}{2}\big((\mathrm{Re}\,(a_{1,i k})-\mathrm{Im}\,(b_{1,i k}))^{2}+(\mathrm{Re}\,(b_{1,i k})+\mathrm{Im}\,(a_{1,i k}))^{2}\big)^{1/2}}\end{array}
+$$
+
+where the subscripts 0, BW and FW denote symmetric, backward whirling and forward whirling motion, respectively.
+
+# 3.2. Classical Floquet analysis
+
+Floquet analysis enables the solution of the periodic equations of motion directly without an explicit transformation. Equation (4) is written in first order form
+
+$$
+\begin{array}{r l}&{\dot{\boldsymbol{y}}_{2}=\mathbf{A}\boldsymbol{y}_{2}}\\ &{\mathbf{A}=\left[\mathbf{-M}^{-1}\mathbf{K}\quad\mathbf{-M}^{-1}\mathbf{C}\right]}\end{array}
+$$
+
+where ${\mathbf y}_{2}=\{{\mathbf y}^{{\mathrm T}}\,\dot{{\mathbf y}}^{{\mathrm T}}\}^{{\mathrm T}}$ is the state vector and $\mathbf{A}$ is the $T$ -periodic system matrix.
+
+Floquet theory22 states that the solution to equation (15) is of the form
+
+$$
+\mathbf{\boldsymbol{y}}_{2}=\mathbf{\boldsymbol{U}}\mathbf{\boldsymbol{e}}^{\mathbf{\boldsymbol{\Lambda}}t}\mathbf{\boldsymbol{U}}^{-1}(0)\mathbf{\boldsymbol{y}}_{2}(0)
+$$
+
+where $\mathbf{U}$ is a $T$ -periodic matrix and $\mathbf{A}$ is a diagonal matrix. One way to construct this solution is to form a fundamental solution to equation (15) as
+
+$$
+\displaystyle\varphi=\bigl[\varphi_{1}\quad\varphi_{2}\quad.\ .\quad\varphi_{N}\bigr]
+$$
+
+over one period, $t\ \in\ [0;T]$ , where $N$ is the number of state variables, such that $\dot{\varphi}\;=\;{\bf A}\varphi$ . The monodromy matrix defined as
+
+$$
+\mathbf{C}=\boldsymbol{\varphi}^{-1}(0)\boldsymbol{\varphi}(T)
+$$
+
+contains all modal properties, which can be extracted from the eigenvalue decomposition
+
+$$
+\mathbf{C}=\mathbf{V}\mathbf{J}\mathbf{V}^{-1}
+$$
+
+where $\mathbf{V}$ contains the column eigenvectors $\nu_{k}$ of $\mathbf{C}$ , which are all assumed to be linearly independent and $\mathbf{J}$ is a diagonal matrix containing the eigenvalues $\rho_{k}$ of $\mathbf{C}$ , called the characteristic multipliers. The characteristic exponents $\lambda_{k}=\sigma_{k}+\mathrm{i}\omega_{k}$ contain the frequency $\omega_{k}$ and damping $\sigma_{k}$ and are related to the characteristic multipliers as $\rho_{k}=\exp(\lambda_{k}T)$ . Because the complex logarithm is not unique, the frequency is not determined uniquely, and the principal frequency $\omega_{\mathrm{p},k}$ and the damping $\sigma_{k}$ are defined from the characteristic multipliers as
+
+$$
+\begin{array}{c}{\displaystyle\sigma_{k}=\frac{1}{T}\ln(\vert\rho_{k}\vert)}\\ {\displaystyle\omega_{\mathrm{p},k}=\frac{1}{T}\arg(\rho_{k})}\end{array}
+$$
+
+where $\arg(\rho_{k})\in]-\pi;\pi]$ is implied, resulting in $\omega_{\mathrm{p},k}\in\big[-\Omega/2;\Omega/2\big]$ . Any integer multiple of the rotor speed can be added to the principal frequency to obtain a more physically meaningful frequency23,24
+
+$$
+\omega_{k}=\omega_{\mathrm{p},k}+j_{k}\Omega
+$$
+
+a choice that also affects the periodic modal matrix $\mathbf{U}$ in equation (16). This matrix $\mathbf{U}$ contains the periodic mode shapes uk and is given as24
+
+$$
+\pmb{u}_{k}=\varphi\nu_{k}\mathrm{e}^{-\lambda_{k}t}
+$$
+
+where the real part of $\lambda_{k}$ is given by equation (20) and the imaginary part of $\lambda_{k}$ is defined by equation (21) by selecting $j_{k}$ such that $\pmb{u}_{k}$ is as constant as possible for degrees of freedXom measured in the inertial frame.
+
+Introducing the Fourier transform of the periodic mode shape
+
+$$
+{\pmb u}_{k}=\sum_{j=-\infty}^{\infty}u_{j k}\mathrm{e}^{\mathrm{i}j\Omega t}
+$$
+
+the transient solution in equation (16) can be written as a sum of harmonic components
+
+$$
+y_{2}=\sum_{k=1}^{N}\sum_{j=-\infty}^{\infty}\mathcal{U}_{j k}\mathrm{e}^{(\sigma_{k}+\mathrm{i}(\omega_{k}+j\Omega))t}q_{k}(0)
+$$
+
+where $\begin{array}{r}{\pmb q(0)=\mathbf{U}^{-1}(0)\mathbf{y}_{2}(0).}\end{array}$ . Note that equation (12) is a special case of this expression for $j=-1,0,1$
+
+# 3.3. Implicit Floquet analysis
+
+The implicit Floquet method is here described based on the detailed description in Bauchau and Nikishkov,12 which focuses on computation of the characteristic multipliers from the state transition matrix $\Phi(T,0)$ . It can be defined in classical Floquet theory as
+
+$$
+\boldsymbol{\varphi}(T)=\boldsymbol{\Phi}(T,0)\,\boldsymbol{\varphi}(0)
+$$
+
+Using equation (18), the relationship between the state transition and monodromy matrices is derived as
+
+$$
+\Phi(T,0)=\varphi(0){\bf C}\,\varphi^{-1}(0)
+$$
+
+showing that $\Phi(T,0)$ and C have identical eigenvalues (characteristic multipliers), and their eigenvectors are related as $\pmb{\nu}_{k}=\pmb{\varphi}^{-1}(0)\pmb{w}_{k}$ , where $w_{k}$ represents the eigenvectors of $\Phi(T,0)$ .
+
+The key feature of the state transition matrix is that it defines the solution $y_{2}(T)=\Phi(T,0)y_{2}(0)$ for a time integration of the system equations (equation (15)) over one period $T$ with initial conditions ${\mathfrak{y}}_{2}(0)$ . Hence, without knowing the state transition matrix, it is possible to obtain the product of it with an arbitrary vector (the initial state vector) by integration of
+
+equation (15) over one period. The Arnoldi algorithm25 is a method to approximate the eigenvalues and the eigenvectors of a matrix, say $\Phi(T,0)$ , using only the matrix multiplication with $\Phi(T,0)$ to construct an $m$ -sized subspace
+
+$$
+\mathbf{P}=\left[p_{1}\quad p_{2}\quad\ldots\quad p_{m}\right]
+$$
+
+that satisfies the orthonormality condition
+
+$$
+\mathbf{P}^{\mathrm{T}}\mathbf{P}=\mathbf{I}
+$$
+
+and where the eigenvalues $\tilde{\rho}_{k}$ of the subspace projected state transition matrix
+
+$$
+\mathbf{H}=\mathbf{P}^{\mathrm{T}}\Phi(T,0)\mathbf{P}
+$$
+
+converge towards the eigenvalues $\rho_{k}$ of $\Phi(T,0)$ with the largest modulus as the size $m$ of the subspace increases. The subspace eigenvectors $\tilde{w}_{k}$ of $\mathbf{H}$ projected back to the full state space converge towards the eigenvectors $w_{k}$ of $\Phi(T,0)$ , i.e. $w_{k}\approx\mathbf{P}\tilde{w}_{k}$ . The Arnoldi algorithm proceeds as follows:
+
+Choose an arbitrary vector $\pmb{p}_{1}$ with $|p_{1}|=1$ for $n=1,2,\ldots,m$ $\pmb{a}:=\Phi(T,0)p_{n}$ (integration of equation (15) over $t\in[0;T])$ $\begin{array}{l}{b:=a}\\ {\mathrm{for~}j=1,2,\dotsc,n}\\ {\quad h_{j,n}:=p_{j}^{\operatorname{T}}a}\\ {\quad b:=b-h_{j,n}p_{j}}\end{array}$ end if $n=t_0$的值,则系统任何一个变量在$t>=t_0$时的运动行为可以被完全确定。
状态变量组的最小性体现在:
-状态变量${x_1}(t)、{x_2}(t),...,{x_n}(t)$为完全表征系统行为所必须的系统变量的最少个数,减少变量数将破坏表征的完全性,而增加变量数将是完全表征系统行为所不需要的。
\ No newline at end of file
+状态变量${x_1}(t)、{x_2}(t),...,{x_n}(t)$为完全表征系统行为所必须的系统变量的最少个数,减少变量数将破坏表征的完全性,而增加变量数将是完全表征系统行为所不需要的。
+
+状态变量值的不唯一性:
+由于系统中变量的个数必大于n,而其中仅有n个是线性无关的,因此决定了状态变量组在选取上的不唯一性。
+
+系统任意两个状态变量组之间的关系:
+系统的任意选取的两个状态变量组之间为线性非奇异变换的关系。
+
+
+# 状态空间描述的形式
+
+1、状态方程:描述系统**状态变量**与**输入变量**之间关系的一阶微分方程组(连续时间系统)或一阶差分方程组(离散时间系统)
+
+**表征系统中输入所引起的内部状态变化的过程。** 输入引起的--->状态变化
+
+![[Pasted image 20250626160523.png]]
\ No newline at end of file