首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 531 毫秒
1.
利用线式互相关PIV系统,采用轴编码器定位周期采样技术,在不同尖速比下对旋转水平轴风力机风轮不同子午面下游流场结构进行测量.分析得到不同条件下的瞬时图、时均图,重点对叶尖涡诱导效应区进行研究.实验结果表明:在风轮下游尾迹中可清晰看到叶轮近尾迹流场中的外部主流区、叶尖涡诱导效应区和中心尾迹区.其中风轮下游尾迹流管廓线是锥形螺旋体;叶尖涡核直径随轴向距离的增加而增大,随着测试方位角的增加,尾迹中各叶片产生的叶尖涡沿螺旋锥形廓线有序地向下游扩散流动;随着尖速比的增加,内部中心尾迹区轴向速度亏损值逐渐增加,并且中心尾迹区的范围逐渐扩大.  相似文献   

2.
水平轴风力机尾迹流场PIV实验研究   总被引:3,自引:0,他引:3  
在水平轴风力机模型不同尖速比条件下,利用PIV粒子图像测速技术对风轮尾迹流场进行了测量。采用锁相平均测量技术,获得了风轮尾迹流场的瞬时速度场、时均速度场、涡量场等有关定量信息,为准确计算风力机的流场、载荷和气动特性等提供了依据。实验结果表明:风轮叶片尾缘后侧的尾迹中存在轴向速度亏损区。尾迹在叶片尾缘生成后,随即发生膨胀。直到风轮下游2倍弦长以后,尾迹低速区逐渐衰减,轴向速度不断增加,尾迹区同时发生收缩现象。风轮尾迹涡从叶片尾缘脱落后,在向下游发展传播过程中,尾迹涡的涡心所形成的运动轨迹是与风轮叶片旋转方向相反的螺旋线,涡量数值随着螺旋线向风轮下游的延伸而减小。由于风力机叶片数少,相邻叶片之间的尾迹基本上不存在互相干扰的现象。  相似文献   

3.
利用线式互相关粒子图像测速(PIV)系统和轴编码器锁相技术,测量了不同尖速比下旋转水平轴风力机叶尖处的流场,获得了风轮叶尖处的瞬时速度场,并通过Tecplot软件处理得到了相应的时均速度场、速度云图及流线图.对瞬时图分析可知:在同一尖速比下,叶尖涡的出现使速度亏损值增加,风轮功率下降;在高尖速比下,流场中叶尖涡出现频率较高,有明显的旋涡结构出现,但随着尖速比的减小,叶尖涡出现的频率降低,只有形成的趋势,在尖速比λ-4的工况下,没有完整的涡出现.通过分析时均图表明:随着尖速比的增加,上游空气通过叶尖后的动能损失增大,叶尖尾迹区内的速度亏损范围增大.  相似文献   

4.
水平轴风力机尾迹流场试验   总被引:3,自引:0,他引:3  
在水平轴风力机模型不同尖速比条件下,利用旋转单斜丝热线在风轮下游进行尾迹流场速度测量。采用周期性采样和锁相平均技术热线测量技术,获得了风轮下游尾迹三维流场的定量信息,为准确计算风力机的流场、载荷和气动特性等提供了依据。试验结果表明:风轮下游尾迹区内气流存在明显的三维性。尾迹在向风轮下游的发展传播过程中,尾迹中心形成的运动轨迹是与风轮叶片旋转方向相反的螺旋线。尾迹区内的速度亏损随风轮下游轴向位置的增加而减弱,在气流向下游流动的过程中尾迹速度亏损值逐渐衰减,尾迹区的宽度不断扩大,并逐渐与主流掺混融合。尾迹区内相同轴向位置上不同叶高处的速度型相似。在叶片的尾迹区内,流动的紊流强度大大高于周围的非尾迹区,其中紊流强度径向、切向分量较大,轴向分量最小,尾迹区内的紊流具有高度不均匀性。  相似文献   

5.
应用PIV粒子图像测速技术,在风洞中测量水平轴风力机模型塔筒的近尾迹流场。通过对模型风力机在不同运行尖速比下、不同叶高平面内的塔筒近尾迹速度场和涡量场的分析,得到了塔筒近尾迹流场的结构特征,为水平轴风力机气动设计、性能预测及CFD数值模拟提供了依据。实验结果表明,受到风轮旋转效应的影响,在水平轴风力机塔筒下游轴向距离6倍当地弦长范围内,近尾迹在水平面内向一侧明显偏转,近尾迹流场相对塔筒中心轴面呈非对称分布。随着尖速比的减小,塔筒下游轴向距离6倍当地弦长范围内,近尾迹涡流宽度逐渐增大,且尾迹向一侧偏转的程度也越大。风力机叶片对塔筒近尾迹涡流的影响,在叶根部位高度平面内尤为显著,随着叶片高度的增加,叶片对塔筒近尾迹涡流的影响逐渐减弱。  相似文献   

6.
垂直轴风力机运行过程中,叶片上下表面边界层与剪切层的相互作用使风力机下游尾迹形成周期性涡结构,这种尾迹涡结构对风力机空气动力学特性具有重要影响。基于此,该文采用计算流体力学方法对不同工况下垂直轴风力机尾迹涡结构展开研究,利用快速傅里叶变换与相空间轨迹分析不同尖速比下风力机叶片涡脱落现象和尾迹涡结构,并通过分形维数研究转矩与尾迹流场速度变化。结果表明:风力机尾迹涡结构随尖速比变化呈现不同特征,当尖速比为3.6时,风力机尾迹两侧呈规则性反向脱落涡模态;低尖速比垂直轴风力机尾迹具有明显的混沌特性,且随尖速比的增加混沌特性逐渐减弱;随着尖速比的增加,风力机转矩与下游速度分形维数不断降低,且当尖速比为3.6时,风力机下游速度分形维数仅为1.07。  相似文献   

7.
韩玉霞  汪建文  李鑫  孙博  刘珍 《太阳能学报》2019,40(4):1179-1184
为研究湍流强度对风力机尾迹涡结构的影响规律,利用TR-PIV(time resolved-particle image velocimetry)对水平轴风力机模型在有、无格栅4.5D(D为风轮直径)范围内的尾流信息进行采集。通过定性及定量分析对比有、无格栅时尾迹流场瞬时涡量、平均涡量及湍动能的变化规律,再现了不同入流条件下尾迹涡形成、发展和湮灭的过程及尾迹涡系间能量传递特性。分析发现:自由流4.5D范围内均可见明显叶尖涡拟序结构,其衰减速度较慢。格栅入流时随湍流强度增加流层间的强剪切及径向掺混作用增强,使叶尖涡拟序结构失稳,2.5D时拟序结构消失;涡量集中区域较自由流明显扩张,叶尖涡诱导效应影响范围增加;尾迹涡系的湍动能较自由流明显增加,随着尾迹向下游发展叶尖涡、中心涡湍动能很快衰减,附着涡区湍动能却明显增强;附着涡区不再是隔离带,而是叶尖涡和中心涡的能量输送带,从而促进尾迹恢复。  相似文献   

8.
水平轴风力机尾迹的测量与分析   总被引:1,自引:0,他引:1  
胡丹梅  田杰  杜朝辉 《动力工程》2006,26(5):751-755,760
在水平轴风力机模型不同尖速比条件下,利用旋转单斜丝热线在风轮下游进行尾迹流场速度测量。采用周期性采样和锁相平均热线测量技术,获得了风轮下游尾迹三维流场的定量信息。实验结果表明:风轮下游尾迹区内气流存在明显的三维性。尾迹在向风轮下游的发展传播过程中,尾迹中心形成的运动轨迹是与风轮叶片旋转方向相反的螺旋线。尾迹区内在气流向下游流动的过程中尾迹速度亏损逐渐衰减,尾迹区的宽度不断扩大,并逐渐与主流掺混融合。在叶片的尾迹区内,流动的紊流强度大大高于周围的非尾迹区,其中径向、切向紊流强度分量较大,而轴向分量紊流强度最小,尾迹区内的紊流具有高度不均匀性。最后,利用CFD软件Fluent6.0对实验风力机的三维流场进行了数值模拟,实验与数值模拟得到了较为一致的结果。图11参10  相似文献   

9.
采用一种简单、有效的方法来改善风力机尾流效应,提升下游风力机功率。进行叶片旋向对风力机尾流特性的试验研究,利用低频粒子图像测速(PIV)系统对NACA4415翼型的叶片进行扰流流场测试并采集风力机的尾流数据。当2台串列排布的风力机旋向不同时,首先在下游风力机前1D(D为风轮直径)处,叶尖涡涡核位置向中央尾迹区偏移,而外部主流区的流体在叶尖涡诱导区的输运和卷吸作用下持续进入中央尾迹区并与之掺混使得轴向速度恢复得更佳;进而分析下游风力机后1D的流场数据,结果显示:虽然下游风力机叶尖涡几何结构被“打碎”,但涡核能量却未降低;最后探讨影响风力机功率特性的因素,下游风力机入流角的增大促使下游风力机捕获更多风能,在风轮间距为2D时,逆向旋转的功率比比同向旋转时高4.70%,且功率比随间距增加其增幅逐渐减小。  相似文献   

10.
文章对有无V型叶尖小翼的风力机尾迹流场进行了研究,重点对叶尖涡的产生与脱落、叶尖区域声辐射进行了分析。结果表明:风轮尾迹区分为主流区、中心尾迹区及叶尖涡诱导效应区;叶尖涡向下游有序流动,随着轴向距离的增加,叶尖涡向外扩展;声压脉动时均值的最大区域集中在风轮叶尖部位,叶尖区域噪声最大;加装V型小翼可以重整通过叶尖流场的气流,使叶尖涡的产生推迟、脱落提前,总声压级降低;数值模拟无小翼时,所选观测点频谱图中声压级总体处于50~70 dB,加装V型小翼后,频谱图中声压级处于45~65 dB,降低效果较明显。实验得到该点V型小翼风轮总声压级为82 dB,比无小翼风轮减少4 dB;风轮辐射声总声压级随着测点向风轮下游移动逐渐衰减,加装V型小翼后总声压级降低,降幅在1~5 dB。  相似文献   

11.
The performance characteristics and the near wake of a model wind turbine were investigated experimentally. The model tested is a three‐bladed horizontal axis type wind turbine with an upstream rotor of 0.90 m diameter. The performance measurements were conducted at various yaw angles, a freestream speed of about 10 m s ?1, and the tip speed ratio was varied from 0.5 to 12. The time‐averaged streamwise velocity field in the near wake of the turbine was measured at different tip speed ratios and downstream locations. As expected, it was found that power and thrust coefficients decrease with increasing yaw angle. The power loss is about 3% when the yaw angle is less than 10° and increases to more than 30% when the yaw angle is greater than 30°. The velocity distribution in the near wake was found to be strongly influenced by the tip speed ratio and the yaw angle. At the optimum tip speed ratio, the axial velocity was almost uniform within the midsection of the rotor wake, whereas two strong peaks are observed for high tip speed ratios when the yaw angle is 0°. As the yaw angle increases, the wake width was found to be reduced and skewed towards the yawed direction. With increasing downstream distance, the wake velocity field was observed to depend on the tip speed ratio and more pronounced at high tip speed ratio. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

12.
An experimental study is conducted to investigate the flow dynamics within the near‐wake region of a horizontal axis wind turbine using particle image velocimetry (PIV). Measurements were performed in the horizontal plane in a row of four radially distributed measurement windows (tiles), which are then patched together to obtain larger measurement field. The mean and turbulent components of the flow field were measured at various blade phase angles. The mean velocity and turbulence characteristics show high dependency on the blade phase angle in the near‐wake region closer to the blade tip and become phase independent further downstream at a distance of about one rotor diameter. In the near‐wake region, both the mean and turbulent characteristics show a systemic variation with the phase angle in the blade tip region, where the highest levels of turbulence are observed. The streamlines of the instantaneous velocity field at a given phase allowed to track a tip vortex which showed wandering trend. The tip vortices are mostly formed at r/R > 1, which indicates the wake expansion. Results also show the gradual movement of the vortex region in the axial direction, which can be attributed to the dynamics of the helical tip vortices which after being generated from the tip, rotate with respect to the blade and move in the axial direction because of the axial momentum of the flow. The axial velocity deficit was compared with other laboratory and field measurements. The comparison shows qualitative similarity. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

13.
偏航状态下风力机叶片与流场之间相互作用会导致风力机近尾迹流场的湍流特征变化,采用双向流固耦合对不同偏航工况下水平轴风力机近尾迹流场进行数值模拟研究,获得不同偏航角下尾迹湍流特征演化规律。结果表明:随着偏航角的增大,正偏航侧会出现“速度亏损圆环”,且此圆环的范围呈扩大趋势;偏航角的增大对叶根处速度亏损影响最大,对叶尖处速度亏损影响最小,与正偏航侧相比,负偏航侧的速度亏损值减为约1/2;随着偏航角的增大,正负偏航侧的湍流强度变化呈不对称性,正偏航侧对湍流耗散的影响程度较负偏航侧大;涡流黏度越来越小,且在偏航10°涡流黏度相对于偏航5°减小约1/2,沿着轴向叶尖涡的管状环涡结构变得不稳定,出现明显耗散,且在偏航15°之后涡结构的耗散破裂程度越来越剧烈,进而对风力机气动噪声产生较大影响。  相似文献   

14.
Fabio Pierella  Lars Sætran 《风能》2017,20(10):1753-1769
In wind farms, the wake of the upstream turbines becomes the inflow for the downstream machines. Ideally, the turbine wake is a stable vortex system. In reality, because of factors like background turbulence, mean flow shear, and tower‐wake interaction, the wake velocity deficit is not symmetric and is displaced away from its mean position. The irregular velocity profile leads to a decreased efficiency and increased blade stress levels for the downstream turbines. The object of this work is the experimental investigation of the effect of the wind turbine tower on the symmetry and displacement of the wake velocity deficit induced by one and two in‐line model wind turbines (,D= 0.9 m). The results of the experiments, performed in the closed‐loop wind tunnel of the Norwegian University of Science and Technology in Trondheim (Norway), showed that the wake of the single turbine expanded more in the horizontal direction (side‐wall normal) than in the vertical (floor normal) direction and that the center of the wake vortex had a tendency to move toward the wind tunnel floor as it was advected downstream from the rotor. The wake of the turbine tandem showed a similar behavior, with a larger degree of non‐symmetry. The analysis of the cross‐stream velocity profiles revealed that the non‐symmetries were caused by a different cross‐stream momentum transport in the top‐tip and bottom‐tip region, induced by the turbine tower wake. In fact, when a second additional turbine tower, mirroring the original one, was installed above the turbine nacelle, the wake recovered its symmetric structure. Copyright © 2017 John Wiley & Sons, Ltd.  相似文献   

15.
Large eddy simulations (LES) of the flow past a wind turbine with and without tower and nacelle have been performed at 2 tip speed ratios (TSR, ), λ=3 and 6, where the latter corresponds to design conditions. The turbine model is placed in a virtual wind tunnel to reproduce the “Blind test 1” experiment performed at the Norwegian University of Science and Technology (NTNU) closed‐loop wind tunnel. The wind turbine was modeled using the actuator line model for the rotor blades and the immersed boundary method for the tower and nacelle. The aim of the paper is to highlight the impact of tower and nacelle on the turbine wake. Therefore, a second set of simulations with the rotating blades only (neglecting the tower and nacelle) has been performed as reference. Present results are compared with the experimental measurements made at NTNU and numerical simulations available in the literature. The tower and nacelle not only produce a velocity deficit in the wake but they also affect the turbulent kinetic energy and the fluxes. The wake of the tower interacts with that generated by the turbine blades promoting the breakdown of the tip vortex and increasing the mean kinetic energy flux into the wake. When tower and nacelle are modeled in the numerical simulations, results improve significantly both in the near wake and in the far wake.  相似文献   

16.
A numerical framework for simulations of wake interactions associated with a wind turbine column is presented. A Reynolds‐averaged Navier‐Stokes (RANS) solver is developed for axisymmetric wake flows using parabolic and boundary‐layer approximations to reduce computational cost while capturing the essential wake physics. Turbulence effects on downstream evolution of the time‐averaged wake velocity field are taken into account through Boussinesq hypothesis and a mixing length model, which is only a function of the streamwise location. The calibration of the turbulence closure model is performed through wake turbulence statistics obtained from large‐eddy simulations of wind turbine wakes. This strategy ensures capturing the proper wake mixing level for a given incoming turbulence and turbine operating condition and, thus, accurately estimating the wake velocity field. The power capture from turbines is mimicked as a forcing in the RANS equations through the actuator disk model with rotation. The RANS simulations of the wake velocity field associated with an isolated 5‐MW NREL wind turbine operating with different tip speed ratios and turbulence intensity of the incoming wind agree well with the analogous velocity data obtained through high‐fidelity large‐eddy simulations. Furthermore, different cases of columns of wind turbines operating with different tip speed ratios and downstream spacing are also simulated with great accuracy. Therefore, the proposed RANS solver is a powerful tool for simulations of wind turbine wakes tailored for optimization problems, where a good trade‐off between accuracy and low‐computational cost is desirable.  相似文献   

17.
为获得风力机近尾流风速在垂直方向和水平方向的变化规律,提出一种测量风力机近尾流区风速的实测方法。针对某沿海滩涂风电场,采用2台搭载风速仪的无人机对近尾流区进行测量。结果表明:垂直方向,尾流和来流风速比值在1.0D~2.5D处(D为风轮直径)随着高度的增加呈先减小后增大的趋势,在轮毂中轴线处存在最小值0.53~0.68;风速比值沿轮毂中轴线呈非对称分布。0.5D处风速比值分别在上下风轮处存在2个极小值0.56和0.50。水平方向,风速比值在1.0D~3.0D处沿径向距离从左向右呈先减小后增大的趋势,在轮毂中轴线处存在最小值(0.54~0.78);风速比值沿轮毂中轴线呈对称分布,随着风轮下游距离的增加呈扩张趋势。最后给出用于A类风场风力机下游尾流风速剖面的预测公式。  相似文献   

18.
Wind measurements were performed with the UTD mobile LiDAR station for an onshore wind farm located in Texas with the aim of characterizing evolution of wind‐turbine wakes for different hub‐height wind speeds and regimes of the static atmospheric stability. The wind velocity field was measured by means of a scanning Doppler wind LiDAR, while atmospheric boundary layer and turbine parameters were monitored through a met‐tower and SCADA, respectively. The wake measurements are clustered and their ensemble statistics retrieved as functions of the hub‐height wind speed and the atmospheric stability regime, which is characterized either with the Bulk Richardson number or wind turbulence intensity at hub height. The cluster analysis of the LiDAR measurements has singled out that the turbine thrust coefficient is the main parameter driving the variability of the velocity deficit in the near wake. In contrast, atmospheric stability has negligible influence on the near‐wake velocity field, while it affects noticeably the far‐wake evolution and recovery. A secondary effect on wake‐recovery rate is observed as a function of the rotor thrust coefficient. For higher thrust coefficients, the enhanced wake‐generated turbulence fosters wake recovery. A semi‐empirical model is formulated to predict the maximum wake velocity deficit as a function of the downstream distance using the rotor thrust coefficient and the incoming turbulence intensity at hub height as input. The cluster analysis of the LiDAR measurements and the ensemble statistics calculated through the Barnes scheme have enabled to generate a valuable dataset for development and assessment of wind farm models.  相似文献   

19.
An equation is derived for the streamwise velocity of the tip vortex of a horizontal-axis wind turbine as the pitch of the vortex tends to zero. The equation is applicable at high tip speed ratios provided the vortex core remains of constant size and there is no flow along the vortex axis. Under these conditions, the vortex velocity is the average of the velocity in the wake and the external wind speed. This result appears to conflict with the computational need to have the vortex velocity approach the wind speed in the high thrust region. It is suggested that the conflict could be resolved by considering the axial flow within the tip vortex.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司    京ICP备09084417号-23

京公网安备 11010802026262号