力学进展, 2020, 50(1): 202001-202001 DOI: 10.6052/1000-0992-19-009

高雷诺数壁湍流的研究进展及挑战

郑晓静,1,2,, 王国华1

1 兰州大学湍流-颗粒研究中心, 西部灾害与环境力学重点实验室,兰州 730000

2 西安电子科技大学机电工程学院, 西安 710071

Progresses and challenges of high Reynolds number wall-bounded turbulence

ZHENG Xiaojing,1,2,, WANG Guohua1

1 Center for Particle-laden Turbulence, Lanzhou University, Key Laboratory of Mechanics on Disaster and Environment in Western China, The Ministry of Education of China, Lanzhou 730000, China

2 School of Mechano-Electronic Engineering, Xidian University, Xi'an 710071, China

通讯作者: E-mail:mengsh@hit.edu.cn

责任编辑: 许春晓

收稿日期: 2019-06-4   接受日期: 2019-10-6   网络出版日期: 2019-11-04

Corresponding authors: E-mail:mengsh@hit.edu.cn

Received: 2019-06-4   Accepted: 2019-10-6   Online: 2019-11-04

作者简介 About authors

郑晓静,1958年5月生,浙江乐清人.1982年在华中科技大学获工学学士学位,1987年在兰州大学获理学博士学位,1992年起任兰州大学教授.现兼任中国科学技术协会副主席,国际理论与应用力学联合会执委会委员,中国科学院数理学部副主任,开云棋牌官方 副理事长,"湍流与复杂系统国家重点实验室"和"甘肃省荒漠化与风沙灾害防治"国家重点实验室(筹)学术委员会主任等.主要研究领域:高雷诺数含沙壁湍流,电磁固体力学、板壳非线性力学.获首届"中国青年科技奖"和"做出突出贡献的中国博士学位获得者"称号,1997年获国家自然基金委员会"杰出青年科学基金",获国家自然科学二等奖2项、国家科技进步二等奖1项.2009年当选中国科学院数理学部院士,2010年当选发展中国家科学院院士并任工程科学部评奖委员会主席.2014年获何梁何利科学与技术进步奖,2017年获"周培源力学奖"等.

摘要

高雷诺数壁湍流(high Reynolds number wall-bounded turbulence,HRNWT)是目前湍流科学研究的一个热点也是一个难点,对其现象、规律及机制的认知不足,理论体系远未建立而且研究手段受到各种限制.本文基于对HRNWT主要研究手段的介绍,针对HRNWT中的湍流统计量、超大尺度结构(very large scale motions,VLSMs)的尺度和形态以及起源和影响及其与颗粒的相互作用,总结了HRNWT的研究现状和最新进展,特别梳理了近年来本文作者团队在HRNWT特别是高雷诺数颗粒两相壁湍流方面的研究成果,并对HRNWT的进一步研究给出了建议及展望.

关键词: 高雷诺数壁湍流 ; 超大尺度结构 ; 两相流 ; 实验观测 ; 计算模拟

Abstract

High Reynolds number wall-bounded turbulence (HRNWT) has been a hotspot and difficult issue in the field of turbulence research in recent years. The understanding of flow phenomena, laws of physics and mechanisms are insufficient and studies on HRNWT are restricted by various deficiencies of research techniques. There is still a long way to go to establish a thorough theoretical framework for HRNWT. Based on the introductions of the leading research techniques, this paper summarizes and reviews the research progresses in statistical characteristics of the HRNWT and very large scale motions (VLSMs) including their morphologies, originations, influences on the turbulent flows, as well as the interactions between turbulence and moving particles in the HRNWT. The contributions of the author's research team on these issues, especially the turbulence-particle interactions are combined. Finally, we highlight the prospects of direction, trends and remaining challenges of further researches.

Keywords: high Reynolds number wall-bounded turbulence ; very large scale motions ; two phase flows ; experimental measurement ; numerical simulation

PDF (9761KB) 元数据 多维度评价 相关文章 导出 EndNote| Ris| Bibtex  收藏本文

本文引用格式

郑晓静, 王国华. 高雷诺数壁湍流的研究进展及挑战. 力学进展[J], 2020, 50(1): 202001-202001 DOI:10.6052/1000-0992-19-009

ZHENG Xiaojing, WANG Guohua. Progresses and challenges of high Reynolds number wall-bounded turbulence. Advances in Mechanics[J], 2020, 50(1): 202001-202001 DOI:10.6052/1000-0992-19-009

1 引言

湍流是流体流动速度在空间和时间上都具有急剧不规则和高度随机性脉动的一种流动状态. 这种流动状态在自然界和工程界以及日常生活中很普遍, 如:自然界中的河流、瀑布、海洋、大气的流动和工业界中发动机的油气混合、燃烧等, "没有湍流我们在地球上将无法生活" (Oertel et al. 2005). 尽管早在欧洲文艺复兴时期, 达芬奇就通过在水流中放入一个方柱状的阻碍物直观地看到流体流动中所出现的大大小小的涡团并绘制了著名的钝体绕流素描图(Frisch 1995), 但真正意义上的实验研究却始于3个世纪后德国流体力学家Hagen (1839)在管道流中以木屑为示踪粒子所进行的流动可视化实验. Hagen的实验定性地告诉人们层流会随着雷诺数的增加转捩为湍流, 而对这种转捩的定量研究可以追溯到英国力学家、物理学家Osborne Reynolds在1883年做的圆管流动实验. 在这个著名的实验中, 他通过观测不同流速、不同直径和不同流体黏度的层流向湍流的转捩, 发现转捩发生时流体流动的特征速度和特征尺度与运动黏度的比值几乎相同. Prandtl (1910)把这个无量纲的比值称之为雷诺数$Re$, 而摩擦雷诺数$Re_\tau=u_\tau\delta /v$ (这里$u_\tau$为流体的壁面摩擦速度, $v$为运动黏度, $\delta$在边界层流动中为边界层厚度, 管道流中为管道半径, 槽道流动中为半槽高度)是壁湍流研究的一个非常重要的参数.

HRNWT是一类具有典型特征的流动形态. 由于壁面的无滑移约束, 流体的黏性作用使得壁湍流在边界层内区(即靠近壁面其分子黏性有重要作用且黏性应力与雷诺应力之和基本不变的区域)和外区(即惯性占主导且分子黏性可以忽略的区域)呈现出不同的流动现象和规律以及内外区间的复杂的相互作用, 成为壁湍流研究的关键和难点. HRNWT广为存在, 如: 行进中的飞机和舰船的湍流$(Re_\tau \sim O (10^4\sim 10^5))$、大气边界层风场$(Re_\tau \sim O (10^6))$等. 尽管迄今为止并没有明确界定高雷诺数的具体范围, 但对于HRNWT的特性已有一些定性的共识. 如: HRNWT中存在显著的流动尺度分离, 使得平均速度可以清晰地显示出经典的对数标度律(Nagib et al. 2007); 又如: 除了存在平均速度的对数区外, 高雷诺数效应使得湍流含能涡与耗散涡的尺度充分分离, 表现为能谱中存在服从$-5/3$标度的惯性子区(McKeon & Morrison 2007); 再如: 把湍动能产生率$p=-\overline{uw}^+{\rm d}U^+/{\rm d}z^+$ (其中$U^+$为无量纲流向平均速度, $\overline{uw}^+$为无量纲雷诺切应力) 在对数区高于在黏性层(Marusic et al. 2010a)或者将平均速度剖面标度律中的尾流因子这一边界层特征参数趋于平稳作为高雷诺数流动的判据(Smits et al. 2011). 自20世纪90年代后, 更多的研究者是把在壁湍流中是否出现VLSMs作为HRNWT的特有标志(Hutchins & Marusic 2007a). 这种VLSMs是一种流向尺度大于$3\delta$的湍流拟序结构, 其在流向速度预乘谱上表现为在低波数段出现一新的峰值. 本文综述的HRNWT将采用VLSMs的存在作为判据.

HRNWT已经逐渐成为流体力学研究的一个活跃领域. 这主要是因为不断发现了一些与基于低雷诺数壁湍流研究得到并形成基本共识的理论、标度律以及所理解的物理过程等有所不同的新现象. 这些新现象包括: 壁湍流在低雷诺数流动时, 其对数区下边界通常被认为是一固定值, 但在高雷诺数情形则发现对数区下边界具有雷诺数依赖性(Klewicki et al. 2009, Marusic et al. 2013); 平均速度对数标度律中的卡门常数$\kappa$在各种类型流动中的随雷诺数的增加逐渐趋于各自的一个不同于低雷诺数流动情形的常数(Nagib & Chauhan 2008) 壁湍流在低雷诺数情形的流向湍流强度沿高度的分布是单峰状的, 但在高雷诺数时则发现在外区出现了第二峰值(Fernholz et al. 1995), 进一步的研究揭示出这一峰值与外区的VLSMs密切相关. 这些HRNWT中的雷诺数效应会引发一些对已有应用研究的挑战. 如: 目前大量的湍流减阻策略是建立在对近壁条带的调控上, 这是因为低雷诺数条件下内区结构生成演化被认为是自维持的, 可以忽略外区的影响(Panton 2001). 而随着雷诺数的增加, 内外区间被证实存在显著的相互作用, 尤其是外区的VLSMs对内区湍流脉动存在显著的调制作用(Mathis et al. 2009, Hutchins et al. 2011), 这样会使得基于原有减阻策略的减阻效率在高雷诺数情形中显著降低(许春晓 2015). 因此, HRNWT的这些较之于低雷诺数情形的新现象不仅说明在现有壁湍流研究中需要深化对雷诺数效应的研究以全面准确地认知壁湍流, 而且还具有重要的应用需求.

本文将从HRNWT研究的实验技术和数值模拟、统计特性、VLSMs的尺度和形态以及起源和影响、与颗粒的相互作用等方面, 在回顾总结已有研究进展的基础上, 重点介绍本文作者团队在国家自然科学基金委的重点项目和重大项目资助下, 针对HRNWT研究的最新工作, 并提出后续值得进一步关注的若干关键问题.

2 主要研究手段

除理论分析包括湍流模式研究外, 实验观测和数值模拟是目前壁湍流研究的两个主要且广泛采用的手段. 自1883年的雷诺实验后, 研究者们针对壁湍流开展了一系列的单点接触式和非接触式测量, 以得到不同情形壁湍流的平均速度剖面、雷诺应力(包括雷诺正应力$u^2$, $v^2$, $w^2$, 即湍动能或湍流强度, 以及雷诺切应力$uv$, $uw$, $vw$, 其中u, v, w分别为流向、展向、垂向脉动速度)、湍流脉动能谱等. 采用的测量仪器主要有皮托管(Pitot tube)、热线风速仪(hot wire anemometer, HWA)、基于示踪粒子测速的粒子图像测速仪(particle image velocimetry, PIV)、激光多普勒测速仪(laser Doppler anemometry, LDA)等. 随着人们逐渐认识到: 湍流不仅仅是随机的, 还是各种尺度涡旋和结构时空演化过程的一种表现后, 为了揭示湍流中的涡旋结构特征和规律及其影响, 发展了一系列可以测量湍流结构的方法, 如: 氢气泡法流动显示(hydrogen bubble flow visualization)、层析粒子图像测速仪(tomographic PIV)、立体粒子图像测速仪(stereo PIV)、热线空间多点同步测量等. 与此同时, 对壁湍流的数值模拟, 特别是高精度直接数值模拟(direct numerical simulation, DNS), 由于其具有能全尺度解析湍流运动的优越性, 也成为对壁湍流实验的有效补充甚至直接检验. 然而, 由于HRNWT中存在显著的流动尺度分离并出现大于$3\delta$的VLSMs, 这就对相应的实验设备及其观测和数值模拟提出了更高要求甚至是巨大挑战.

在实验室中发现了HRNWT的一系列新现象. Kim 和 Adrian (1999)最先在实验室研究中发现VLSMs. 这除了得益于他们精细的实验设计和测量以及深刻分析外, 还得益于他们的实验是在普林斯顿大学的超级管(Superpipe,其直径0.129 m, 长26 m, 使用压缩空气得到的最高$Re_\tau=5.0\times 10^5)$中进行的. 利用这一装置, 通过识别预乘能谱的峰值尺度, 他们发现当$Re_\tau=3000$左右时会出现流向尺度最大可达圆管半径14倍的湍流结构. 这一新现象引发了研究者们对HRNWT的浓厚兴趣, 一个直接的驱动是: 这种VLSMs是否还会出现在其他流动形式中? 随着雷诺数的提高, 壁湍流还会出现什么有别于中低雷诺数流动的新现象和新规律? 于是, Österlund (2000)Nagib 等(2007)分别在瑞典皇家理工学院的最小湍流度风洞(Minimum Turbulence Level, MTL, 其截面尺寸1.2 m $\times$ 0.8 m、长7.0 m, 最高$Re_\tau=1.4\times 10^4)$和美国伊利诺伊理工大学的国家诊断设施风洞(National Diagnostic Facility, NDF, 其截面尺寸1.52 m $\times$ 1.22 m、长10.3 m, 最高$Re_\tau=2.2\times 10^4$)开展了零压力梯度下湍流边界层对数区范围以及若干标度参数的雷诺数效应研究. 他们发现: 壁湍流内区和外区的重叠区域, 即重叠区, 下限$z^+=zu_\tau/v\approx 200$远高于低雷诺数情形中的$z^+=30$; 相应的尾流因子和形状因子呈现出与低雷诺数流动情形不同的雷诺数渐进特征. 继20世纪建造的上述3个装置后, 本世纪新建成的专用于HRNWT研究的设施主要有: 澳大利亚墨尔本大学的高雷诺数边界层风洞(High Reynolds Number Boundary Layer Wind Tunnel, HRNBLWT, 其截面为1.89 m $\times$ 0.92 m、长27 m、最高$Re_\tau=3.2\times 10^4$), 美国新罕布什尔大学的流体物理设施风洞(Flow Physics Facility, FPF, 其截面为2.5 m $\times$ 6 m, 长72 m, 最高$Re_\tau=5.0\times 10^4$), 以及位于意大利普雷达皮奥一座山体内的于2006年设计建造的长风管(Center for International Cooperation in Long Pipe Experiments, CICLoPE, 其直径0.9 m, 长111.5 m, 最高$Re_\tau=4.0\times 10^4$). 这些后续建成的风洞最高雷诺数虽然比Superpipe的低, 但由于所使用的是常压、常密度气体, 且测量段截面积比Superpipe的要大得多, 这就在一定程度上降低了对测速仪器分辨率的要求. 类似于Superpipe, 同样基于压缩空气的原理, 普林斯顿大学空气动力实验室还建成了一座用于湍流边界层测量的高雷诺数风洞(High Reynolds Number Test Facility, HRTF, 其直径0.46 m、长4.8 m, 最高$Re_\tau=8.9\times 10^4$). 借助这些实验装置, 研究者获得了HRNWT的宝贵数据并得到了一些反映雷诺数对壁湍流影响的重要成果. 如: Nickels 等(2007)在HRNBLWT进行的$Re_\tau=2.3\times 10^4$的实验证实: $z^+=300$ 处流向湍动能随雷诺数的变化满足基于附着涡模型的预测; Hultmark 等(2013)在Superpipe的研究指出流向湍动能同样存在满足对数标度律的区域且湍动能的对数标度范围与平均速度的对数区一致, 这一结果被在HRTF开展的$Re_\tau=7.25\times 10^4$的边界层测量证实 (Vallikivi et al. 2015a); Vincenti 等 (2013) 在FPF进行的边界层实验(最高$Re_\tau=1.967\times 10^4)$以及Willert 等 (2017)利用CICLoPE开展的管道实验$(Re_\tau=4.0\times 10^4)$则证实湍动能的内区和外区峰值强度都随雷诺数增大而增强, 而这些现象在低雷诺数实验中没有发现.

HRNWT实验室研究的设施和测量仪器成本很高. 满足HRNWT实验测量的设施至少应同时满足两个条件: 一是流动形成的重叠区要足够大以保证该区域的平均速度服从对数律; 二是能清晰展示湍流脉动谱中的惯性子区(McKeon & Morrison 2007). 为此, 无论是通过压缩气体来提高气体密度从而降低其运动黏度, 还是通过增大边界层厚度来在实验室中实现较高雷诺数流动, 这两种方式的造价均在数百万美元以上, 如: HRNBLWT仅洞体造价就约400万澳元. 对于第一种方式, 由于压缩气体后的边界层流动的黏性尺度降低, 进而要求测量仪器的尺寸更小以减小对流场的干扰, 同时又要求有更高的时空分辨率来实现全尺度分辨. 以在Superpipe中$Re_\tau=7.0\times 10^4$的流动为例, 此时压缩气体的黏性尺度$v/u_\tau<1 \mu$m, 为了达到全分辨并避免"平均效应" (脉动信号沿热丝长度平均)和探头的"端部传导效应" (热量从热丝向支架连接臂传导引起的响应误差), 热丝的长度不应大于10倍的黏性尺度且热丝长细比应大于200 (Samie et al. 2018), 这样热丝长度和直径分别要小于10 $\mu$m和 0.05 $\mu$m. 目前标准热线探头(长120 $\mu$m、直径0.6 $\mu$m)分辨尺度大于12 $\mu$m是难以实现对黏性尺度$v/u_\tau<1 \mu$m流动的准确测量, 因此, 尽管Superpipe设备可以实现$Re_\tau\sim O(10^5)$的流动, 但由于测量精度的限制使得目前仍未给出在这一雷诺数条件下的实验结果. 对于第二种方式, 除了洞体增大带来的造价外, 高功率且运行稳定的气流驱动风扇等使得运行成本会很高. 也许正是这些原因, 目前实验室研究的流动$Re_\tau$还只是在$O(10^4)$量级, 进一步提高的难度较大.

解析湍流结构的DNS所达到的最高$Re_\tau$一直在$O(10^{13})$徘徊. 传统的雷诺平均方程方法(Reynolds averaged Navier-Stokes equations, RANS)能够以较小的计算量得到工程上所关心的湍流平均统计特性, 但不足之处是计算结果强烈依赖于用于雷诺平均运动方程封闭的湍流模型的选择和难以计及湍流脉动的影响. 大涡模拟(large eddy simulation, LES)或分离涡模拟(detached eddy simulation, DES)等方法能够接近或达到工程问题的雷诺数量级$O(10^5\sim 10^6)$, 但只能解析部分湍流结构且计算结果的好坏依赖于湍流模型和用于近壁流动简化计算的壁模型等. DNS能够解析全部的湍流结构而不引入任何模型和假设, 是目前最可靠和最精确的湍流模拟方法(Moin & Mahesh 1998). 最早采用DNS对湍流进行模拟的是美国学者Orszag 和 Patterson (1972), 他们模拟了$Re_ \lambda=35$的各向同性湍流. 受计算机能力的限制, 大约到20世纪80年代后期, 才有学者逐步开展了对槽道、管道和边界层流动的DNS研究. Kim 等(1987)Spalart (1988)先后采用DNS模拟了$Re_\tau=180$的槽道湍流和$Re_\tau$ 为$100\sim 550$的湍流边界层, 分析了平均速度、湍流强度、雷诺应力等湍流统计特征和近壁条带等. 随着计算能力的快速提升和并行计算方法的不断进步, 壁湍流DNS模拟的雷诺数记录不断被打破. 以槽道湍流为例, DNS模拟的雷诺数从20世纪80年代的$Re_\tau=1.8\times 10^2$ (Kim et al. 1987)到90年代的$Re_\tau=5.9\times 10^2$ (Moser et al. 1999), 再到本世纪初的$Re_\tau=2.003\times 10^3$ (Hoyas & Jiménez 2006), $Re_\tau=4.0\times 10^3$ (Bernardini et al. 2014)、$Re_\tau=4.2\times 10^3$ (Lozano-Durán & Jiménez 2014)、$Re_\tau\approx 5.2\times 10^3$ (Lee & Moser 2015)以及Yamamoto 和 Tsuji (2018) 的$Re_\tau\approx 8.0\times 10^3$ (其数据质量受到了质疑, 比如其内区湍动能峰值低于$Re_\tau\approx 5.2\times 10^3$的结果, 与目前内区湍动能峰值随雷诺数增加的基本认识相悖)和Hoyas 等(2018)在2018年美国物理学会流体力学分会的年会上报道的$Re_\tau=1.0\times 10^4$. 而针对湍流边界层和管道湍流的DNS模拟, 目前最高分别是$Re_\tau\approx 2.0\times 10^3$ (Sillero et al. 2013, 2014)和$Re_\tau\approx 3.0\times 10^3$ (Ahn et al. 2015). 清华大学许春晓课题组实现了$Re_\tau= 1.0\times 10^3$的槽道湍流DNS模拟(Deng et al. 2016). 综上可见, 这种不断提高所模拟的壁湍流雷诺数的努力一直在持续而且提高速度越来越快, 但即便如此, 目前DNS所能模拟的壁湍流$Re_\tau$最高也只是在$O(10^3)$量级, 比大多数工程实际的雷诺数还低$2\sim 3$个数量级. 这种局面可能还得持续一段时间, 因为DNS既需要足够大的计算区域以包含湍流中VLSMs, 也需要足够小的网格来捕捉最小尺度的湍流涡, 其网格数大约与$Re^{37/14}$成正比(Choi & Moin 2012). 因此, 显著的突破依赖于计算机速度的提升、数据容量的扩大以及并行算法效率的提高.

近中性大气表面层是HRNWT研究的天然实验平台. 实验室研究和数值模拟的困难使得针对HRNWT的研究不得不另辟蹊径, 将目光转向自然界. 大气表面层的厚度一般在$100\sim 200$ m, 即使是在较小风速情形(如2 m高度风速5 m/s), 其$Re_\tau$也可高达$O(10^6)$量级. 因此, "作为唯一真实的高雷诺数流体, 大气表面层被看作是此类实验的一个非常有用的基准" (Guala et al. 2011), 而且它代表了地球上可达到的最高雷诺数的流动条件, 且不会因为小尺度边界层对探头分辨率进行严格限制(Marusic et al. 2010c). 气象、海洋、地理包括风沙物理学界的研究者们对不同下垫面的大气表面层观测由来已久, 近期比较著名的例如美国的Kansas观测和Minnesota观测(Kaimal & Wyngaard 1989)、国内的黑河和青藏高原观测(胡隐樵等 1994, 徐祥德等 2001)以及中科院大气物理所关于城市边界层的观测(Zeng et al. 2010)等, 主要涉及对气压、温湿度、平均风速、降水等的常规观测和对大气湍流的湍流度、地表切应力、热通量、水汽通量等的近地层微气象观测, 以揭示局地气象及其对气候的影响并对大尺度天气预报模式提出湍流参数化方案等. 常规气象观测的测量频率通常较低, 很难满足对湍流信号分析的高分辨率要求. 微气象观测虽然频率较高, 但现有观测的下垫面和流动环境较为复杂, 多为草场、农田、河谷、戈壁、沙漠、城镇等, 其结果很难与规范平板湍流边界层的进行类比, 导致这些观测数据无法用于HRNWT研究. 地理学界对风沙运动系统的近地层野外观测和实验研究始于风沙物理学创始人Bagnold (1941), 随后许多学者围绕输沙通量及其影响因素和变化规律进行了观测和实验, 主要关注输沙率、土壤风蚀、沙粒起跳速度和临界启动风速等, 有关的介绍请见Zheng (2009). 除了侧重点的差异, 地理学界对大气表面层风场的观测主要是获得平均风速剖面, 其观测数据频率较低, 也还很难用于HRNWT研究. 由于实验观测发现: 非定常来流的输沙率与定常来流有明显差异(Jackson 1996, Rasmussen & Sørensen 1999), 近年来风沙物理学界的研究人员也逐渐开始关注湍流脉动对输沙率间歇特征和时空变化的影响等(Greeley et al. 1996, Stout & Zobeck 1997, Carneiro et al. 2015, Baas 2006, Martin & Kok 2018), 但尚未对近年来兴起的HRNWT研究进展予以应有的关注和采用. 美国犹他大学的表面层湍流及环境科学测试(Surface Layer Turbulence and Environmental Science Test, SLTEST)开启了对大气表面层HRNWT的观测. 该观测场地位于美国犹他州大盐湖湖床上, 由于每年周期性的干涸, 干涸后的湖床地表平坦且有坚硬的结皮. 由于该区域2 m处的风速一般低于8 m/s, 因此, 在该区域主要观测的是$Re_\tau=6.28\times 10^5\sim 3.8\times 10^6$的净风场. 观测装置包括沿风向来流方向的展向的一排高2.14 m、间距3 m的塔架和位于中心处的塔架高达27.5 m以及安装了 测量频率为20 Hz的超声风速仪, 可以开展对来流净风场单点梯度和展向多点的风速同步测量. 利用SLTEST, 研究者们采集到$Re_\tau=6.0\times 10^5\sim 3.0\times 10^6$的约70 h不同层结条件的平稳风场数据(Metzger & Klewicki 2001, Chauhan et al. 2013), 并证实所得到的大气表面层湍流统计量的标度律与经典零压力梯度平板湍流边界层的基本一致(Kunkel & Marusic 2006). 由此表明: 尽管大气流动受天气及环境因素的影响, 野外观测的风速和风向均不可控且可能会受到显著的热不稳定性的影响, 但通过严格的数据筛选, 近中性条件下高质量的表面层流动数据仍是可以提供规范湍流边界层研究所需的有效数据. 基于SLTEST的数据, 研究者们发现大气表面层中存在VLSMs (Hutchins & Marusic 2007a, Marusic & Hutchins 2008, Hutchins et al. 2012)且VLSMs的倾角不随雷诺数变化但受到层结稳定性的显著影响(Marusic & Heuer 2007, Chauhan et al. 2013)、湍动能内区峰值随雷诺数增大(Metzger & Klewicki 2001)且湍动能分布存在对数标度区(Marusic et al. 2013)、证实了外区VLSMs对内区小尺度运动有显著的调制作用(Mathis et al. 2009, Guala et al. 2011)等. 然而, 由于缺少流向观测塔架, SLTEST无法实现沿来流流向的测量, 而且由于测量周期短, 所获得的可直接用于零压力梯度规范湍流边界层研究的近中性层结的数据较少.

兰州大学西部灾害与环境力学教育部重点实验室的青土湖观测阵列(Qingtu Lake Observation Array, QLOA)提供了对大气表面层HRNWT的最全面的高质量观测. 该阵列位于中国甘肃的青土湖(E: 103$^\circ$ 40${'}$ 03${''}$, N: 39$^\circ$ 12${'}$ 27${''}$)湖床上, 由于大面积干涸, 裸露的湖床地表平坦、开阔、无植被, 加之还位于巴丹吉林沙漠和腾格里沙漠之间的中国沙尘暴多发的西北路径上, 因此, 相对于美国犹他州的SLTEST, 在QLOA可进行含颗粒的HRNWT研究. 自2012年以来, 本文作者团队在国家自然科学基金委重点项目和重大项目的资助下, 在青土湖建立起一个占地约11万平方米的观测阵列, 见图1, 阵列前方20 km (流向) $\times$ 10 km (展向)范围内的地表起伏低于1 m. QLOA观测阵列包括一个高32 m的主观测塔、沿来流上风向(即西北方向)和与其正交的展向分别 布置的11个和12个间距为5 m的观测辅塔、沿来流上风向(正西方向)布置的可用于辅助流向测 量的9个间距10 m到30 m的观测辅塔. 在主塔沿高度按对数等间距配置了采样频率分别为50 Hz的超声风速仪和1 Hz的PM10粉尘仪、 温湿度测量仪、能见度仪以及三维电场仪等测量设备. 所使用的电场仪由本文作者团队自行研制, 探头尺寸只有常规大气电场仪的十分之一, 探头间的相互影响可以忽略, 可在较小的区域安装多个探头实现三维电场测量, 而且探头采用振动感应电极, 可以有效降低沙粒冲击对电场测量的干扰. 另外, 在每个辅塔的5 m处也配置了超声风速仪. 这些观测仪器经由多台相同型号的数据采集仪进行数据采集, 数据采集仪之间通过全球定位系统(GPS)来达到实时同步, 由此实现对流向390 m、展向60 m、垂向30 m空间区域内大气表面层净风场和含沙(尘)风场 (包括沙尘暴情形)的沿流向、展向和垂向的空间全场的三个方向风速分量和沙尘浓度, 温湿度, 能见度, 电场强度等多物理量同步实时高频测量. 本文作者团队由QLOA共获得7400余小时的多物理场同步观测数据, 积累的数据总量超过4.7Tb. 在这些观测数据中, 高质量的平稳数据近600 h, 其中净风条件下近中性层结数据120 h, 非中性层结数据302 h; 含沙流动中, 近中性层结数据83 h, 非中性层结数据90 h. 净风和含沙(包括沙尘暴情形)流动的特征雷诺数分别高达$4.7\times 10^6$和$5.4\times 10^6$, 均为目前净风及含颗粒两相流动测量的最高雷诺数. 北京大学佘振苏团队仔细分析QLOA观测数据后指出: "观测数据具有极高的精度, 可以为研究高雷诺数壁湍流、大气表面层、风沙流、沙尘暴起源等科学问题提供极佳的数据支撑" (纪勇 2019); Journal of Fluid Mechanics期刊审稿人也认为: QLOA观测阵列的观测数据是"独一无二"的. 目前, QLOA已经被认为是利用大气表面层进行HRNWT研究的最著名的两个观测设施之一(Heisel et al. 2018), 而相比于美国犹他州的SLTEST, QLOA由于设有沿风向的流向辅塔进而可以实现对大气表面层的全场观测, 而且观测的物理要素更多、周期更长、数据量更大. 特别是由于观测到的不仅有不同风速的净风场, 而且还有包括含沙(尘)流场以及沙尘暴从发展阶段到平稳阶段以及衰退阶段全过程信息, 因此所观测的大气表面层流动情形更丰富, 获得的净风和含颗粒两相流动的雷诺数均是目前最高的, 而且所测数据最全, 精度也是最好的.

图1

图1   (a)青土湖观测列阵(QLOA)和(b)观测塔布置


本文作者团队基于QLOA数据已获得HRNWT的一些新的现象和规律. 主要包括: 不仅发现大气表面层净风场存在VLSMs, 而且其含沙流场和相应的沙粒浓度场也存在有VLSMs (顾海华和郑晓静 2019); VLSMs的尺度和倾角等形态特征(Liu et al. 2017a, 2017b)以及其能量沿高度及尺度的分布与中低雷诺数情形不同(Wang & Zheng 2016); 首次直接测量了大气表面层的VLSMs流向尺度, 并依此给出泰勒冻结假设在估计VLSMs尺度时的误差及适用性(Han et al. 2019b); 揭示出VLSMs对不同高度沙尘垂向输运的不同作用(Wang et al. 2017)和对不同尺度湍流结构的调制作用(Liu et al. 2019) 明确指出大气学界的"阵风"概念在很大程度上丢失了对能量及物质输运起主导作用的VLSMs的流动信号(Gu et al. 2019). 详细结果请见本文后续部分. QLOA的观测数据还提供给墨尔本大学Ivan Marusic、明尼苏达大学Lian Shen以及北京大学陶建军和佘振苏、清华大学许春晓教授、北京航空航天大学王晋军等学者的团队, 以共同推进HRNWT的研究.

3 统计特性

湍流统计特性对认识湍流的随机性和湍流能谱以及速度关联规律乃至建立湍流封闭模型至关重要. 在湍流研究的很长一段时间内, 对湍流统计特性的分析占有主导地位, 其中Reynolds的贡献具有奠基性. 他将流体流动的速度分成平均项与脉动项之和, 并基于Navier-Stokes方程(N-S方程), 推导得到了现在广为应用的湍动能输运方程, 即Reynolds方程(Reynolds 1894), 揭示了从平均流向湍流传递的关系. Reynolds方程的建立启发了边界层混合长度理论的建立(Prandtl 1925)并促进了对湍流统计量的定量研究, 而且还极大地推动了流体力学的应用研究以及热线测量、速度相关仪和谱仪等实验技术的发展. 湍流统计特性包括与平均值有关的一阶和与脉动有关的二阶统计量以及高阶矩、自相关和互相关函数以及频谱等, 其中受到普遍关注的是湍流一阶统计量(即平均速度剖面)和二阶统计量(即湍动能或湍流强度、雷诺切应力). 这些统计量在对N-S方程简化进而建立便于工程应用的湍流模式, 如: 混合长度模型、$k-\varepsilon$, $k-\omega$两方程模型等Reynolds平均模型时至关重要, 因而被广泛研究. 基于中低雷诺数壁湍流的实验和数值模拟以及相应的理论分析, 研究者们得到了一系列关于湍流统计量的许多重要的规律, 而且认为这些规律和常数不受雷诺数的影响. 然而, 自20世纪90年代后, HRNWT的研究结果却告诉人们情况可能并非如此.

被广泛用于壁湍流研究和应用的平均速度剖面标度的对数律受到挑战. 平均速度剖面标度律是建立和检验壁湍流边界层分层的重要依据, 而标度律的形式、适用范围及参数均有重要的科学和应用价值. 标度律的形式对认识湍流的基本规律以及湍流模式研究非常重要, 如: 大多数RANS和大涡模拟的近壁模型是基于对数律建立的; 标度律的适用范围直接影响标度律中参数的确定, 如: 对数律中的卡门常数$\kappa$的计算受到对数区范围划分的影响; 标度参数则对实际应用非常关键, 如: $\kappa$减小2%会引起现代飞行器总阻力预测降低1% (George 2007). 著名流体力学家Prandtl (1925), von Kármán (1930)Millikan (1938)分别推导出的重叠区平均速度服从对数标度律得到普遍认同并广为应用, 如: 在实验中被用来经常估算壁面摩擦、在数值模拟中被用来给出近壁模型等. 然而, 随着HRNWT研究的深入, 研究者们对平均速度剖面标度律的形式、其中的参数是否是常数以及满足对数标度律的范围给出了不同结果. 对于标度律的形式, Barenblatt 和 Prostokishin (1993)等根据量纲分析认为: 有限雷诺数下壁湍流是不完全相似的, 平均速度分布应该服从幂次律. 随后, George (1995)则从平均运动的RANS方程出发, 给出了槽道和圆管湍流满足的幂次律, 并进一步推广到有压力梯度的湍流边界层 (Castillo & George 2001). 这种与雷诺数有关的标度形式的差异意味着Prandtl (1925)提出并获得广泛认可和应用的平均速度对数标度律不再是与雷诺数无关且普适的. 幂次标度律可能更适用于较低雷诺数, 其分界线, 对于管道流和湍流边界层, 分别大约是$Re_\tau=9000$ (Zagarola & Smits 1998)或者5000 (McKeon et al. 2004)和$Re_\tau=4.0\times 10^4$ (Vallikivi et al. 2015a), 目前仍无定论; 对于标度律适用的范围, 早期的研究认为平均速度对数区的范围不随雷诺数变化, 但近年来的发现则是与雷诺数有关, 且比较一致的观点是: 雷诺数越高, 幂次律适用的范围越小且高度越低 (Willert et al. 2017), 而对数律适用范围大致为$3Re_\tau^{1/2}<z^+<0.15Re_\tau$ (Marusic et al. 2013). 当然, 这也可能与流动形式有关, 如: 对于槽道流和湍流边界层, 可能分别是$400<z^+<0.16Re_\tau$ (Lee & Moser 2015)和$400<z^+<0.15Re_\tau$ (Vallikivi et al. 2015a), 而对于管道流, 对数区上界大致为$0.2Re_\tau$ (Furuichi et al. 2018). 对数律受到挑战的另一个方面是其中被认为是不依赖于流动类型和雷诺数的$\kappa=0.40\sim 0.41$ (Schlichting & Gersten 2000)可能不再是普适常数. 近来大量的证据表明, $\kappa $随雷诺数变化, 且在雷诺数足够高后的收敛值与流动类型有关 (Nagib & Chauhan 2008), 如: 圆管中$\kappa=0.40\pm 0.02$ (Bailey et al. 2014)或者$\kappa=0.384$ (Furuichi et al. 2018); 边界层和槽道中$0.384<\kappa<0.389$ (Österlund et al. 2000, Chauhan et al. 2007, Monty 2005, Lee & Moser 2015); SLTEST给出的$\kappa=0.41\pm 0.02$ (Morris et al. 2007). 尽管有关平均速度剖面表征形式尚未明确定论, 但有一点是基本共识: 既然平均速度的标度形式和参数以及适用范围均与雷诺数有关, 这样就需要对更高雷诺数情况下的流动开展研究.

高雷诺数情况下的湍动能峰值大小、个数及其出现的位置也有新的发现. Reichardt (1933)的风洞 $(Re_{\rm D}$约为7500, 对应的$Re_\tau $约为350)测量结果清晰给出流向湍流脉动均方根随高度变化在内区出现一个峰值. 一般认为在中低雷诺数下峰值一般出现在内区$z^+=15$且可以用摩擦速度进行标度. 然而, 近年的实验和数值研究发现: 随着雷诺数的升高, 虽然内区峰值的位置几乎不随雷诺数变化, 但内区峰值大小$u^{2+}$ $(u^{2+}=\langle uu\rangle /u^{2}_\tau)$随雷诺数呈近似对数线性增长, 其增长规律与流动形式有关, 如在槽道流(Hoyas & Jiménez 2006)的增长就比在湍流边界层 (Marusic & Kunkel 2003)相对缓慢. 这些发现意味着在非常靠近壁面的内区湍流不能完全由内尺度来标度 (Marusic et al. 2010a), 于是, 有学者(De Graaff & Eaton 2000)认为如果采用由摩擦速度$u_\tau$和外流速度$U_{\rm e}$构成的混合速度$(u_\tau U_{\rm e})^{1/2}$来标度$u^{2+}$则可消除内区湍动能峰值的雷诺数相关性. 可是, 雷诺数进一步升高的新结果又发现: 流向湍动能$u^{2+}$在对数区中间位置逐渐抬起, 在形成平台后出现第二个峰值 (Klewicki 2010, Willert et al. 2017, Samie et al. 2018), 如图2 所示; $u^{2+}$内区峰值随雷诺数的升高和外区峰值的出现都与VLSMs有关(Marusic et al. 2010a); $u^{2+}$外区的峰值位置正比于$Re_\tau^{1/2}$而峰值大小随雷诺数近似对数线性增大直到$Re_\tau=20 000$时趋于稳 定(Vallikivi et al. 2015b). 在定性标度的基础上, 研究者们还希望给出湍动能的定量标度. 利用附着涡模型, Perry 等 (1986)推导出流向湍动能分布在外区满足对数律, 但直到本世纪初不断开展的高雷诺数实验(Marusic & Kunkel 2003, Hultmark et al. 2012)才对此提供了支撑. 有学者认为流向湍动能分布是平均速度亏损律的线性函数(Alfredsson et al. 2011), 但目前比较认同的是流向湍动能分布在$2.0\times 10^4<Re_\tau<6.0\times 10^5$内服从$u^{2+}=B_1-A_1\lg(z/\delta)$, 其中$A_{1}$, $B_{1}$为标度系数, 且适用范围与平均速度的对数区一致, 均为$3Re_\tau^{2+}<z^+<0.15Re_\tau$ (Marusic et al. 2013). 但也有学者质疑这一标度关系中的标度系数可能不是普适的, 不仅与流动形式有关(Vallikivi et al. 2015a, 2015b)也与实验数据的拟合有关. 对于湍流流场的垂向湍动能$w^{2+}$ $(w^{2+}=\langle ww\rangle /u^{2}_\tau)$和展向湍动能$v^{2+}$ $(v^{2+}=\langle vv\rangle /u^{2}_\tau)$, 由于实验测量的困难, 结果相对较少. 目前的主要发现和结论是对于$w^{2+}$和$v^{2+}$无论采用内尺度$u_\tau$还是混合尺度$(u_\tau U_{\rm e})^{1/2}$均不能消除其雷诺数相关性 (Bernardini et al. 2014). 关于$w^{2+}$, 虽然有学者指出当$Re_\tau$高于2000以后, $w^{2+}$的峰值逐渐趋于常数(DeGraaff & Eaton 2000), 然而利用SLTEST数据结合实验室结果分析发现随雷诺数的升高, $w^{2+}$峰值增大、峰值位置外移(Kunkel & Marusic 2006, Bernardini et al. 2014). 关于$v^{2+}$, 已有研究指出随雷诺数升高, $v^{2+}$以近似正比于雷诺数对数的速率增大, 并与$u^{2+}$类似, 在$z^+=15$出现峰值(Zhao & Smits 2006), 然后出现明显的随$z^+$的对数衰减区 (Hoyas & Jiménez 2006). Dixit 和 Ramesh (2018)利用Talluru 等 (2014) 的边界层数据分析后指出当$Re_\tau>7000$时, $w^{2+}$和$v^{2+}$也会满足对数标度. 对于雷诺应力$\langle uw\rangle^{+}$ $(\langle uw\rangle^{+}=\langle uw\rangle /u^{2}_\tau)$, 在边界层湍流情形的最大值稍大于1 (Wei et al. 2005), 在槽道湍流则只有当雷诺数趋于无穷时才趋于1 (Lee & Moser 2015). 在内尺度标度下, 其峰值位置正比于雷诺数的1/2次方, 即$z_{\rm m}^+=C(Re_\tau)^{1/2}$, 并且随雷诺数升高峰值变大, 其附近的平台范围变宽, 在峰值平台以内应为黏性尺度$v/u_\tau$, 在平台以外为外尺度$\delta$, 而在平台附近由内外尺度构成的混合尺度$(v\delta/u_\tau)^{1/2}$效果更好(Klewicki 2010). 综上所述, 湍动能分布及其标度规律也呈现出较为明显的高雷诺数效应.

图2

图2   流向速度预乘谱云图与流向湍动能沿高度分布的比较(Hutchins & Marusic 2007a)


QLOA的观测给出了目前雷诺数最高情形$Re_\tau\sim O(10^6)$下边界层湍流统计量结果. 本文作者团队基于对净风场和含沙尘风场的观测数据分析认为: 大气表面层平均速度剖面均基本满足对数标度关系; 采用通常的数据拟合方法得到的净风条件下$\kappa=0.407$, 含沙流动中$\kappa=0.427$, 见图3, 均高于中低雷诺数条件下湍流边界层给出的$\kappa=0.384$ (Nagib & Chauhan 2008) 渐进平稳值; 在证实了净风流动中湍动能存在对数标度的基础上得到含沙尘流动中的湍动能随高度的变化近似满足对数线性减小的规律, 只是相同外标度高度处的湍动能随着雷诺数的增加而增大, 由此揭示出已有湍动能对数标度关系中的系数$A_{1}$和$B_{1}$应当具有雷诺数效应, 如图4(a)所示; 发现对数区上部的垂向湍动能并未出现中低雷诺数边界层流动中显示的衰减区, 而是随着高度增加而增大, 且这一趋势随着雷诺数增加更为明显(Yang & Bo 2018); 在$Re_\tau\sim O(10^6)$的ASL的净风和含沙流动中的雷诺切应力分布, 见图4(b), 符合经典理论预测的雷诺切应力分布规律. 由QLOA数据得到的ASL边界层湍流统计量的新结果对现有HRNWT研究是一种更高雷诺数情形的非常难得的更新, 也是对壁湍流雷诺数效应的非常宝贵的依据.

图3

图3   槽道({\tiny{$\square{}$}})、管道($\vartriangle$)以及零压力梯度边界层($\circ{}$, $\bullet$, $\bullet$, ${ \blacksquare}$ )中卡门常数$\kappa$随雷诺数$Re_\tau$的变化. 空心符号结果取自Nagib & Chauhan (2008), 实心符号分别为SLTEST (Morris et al. 2007)和QLOA结果(顾海华, 郑晓静 2019), 其中净风及含沙流动的$\kappa$分别由QLOA中性层结条件下的18组和22组数据拟合得到


图4

图4   (a)流向湍动能随高度的变化, (b)雷诺切应力随内标度高度的变化. $(\bullet)$ 和(${ \blacksquare}$)分别为QLOA净风和含沙流动结果, $(\blacktriangle)$ 是SLTEST结果(Hutchins et al. 2012), ($\vartriangle\triangledown$) 是湍流边界层结果(DeGraaff & Eaton 2000), ($\circ{}$, ${\square{}}$) 是槽道结果(Schultz & Flack 2013), - - - 是理论公式结果(Chauhan 2007)


4 VLSMs的尺度与形态

VLSMs的发现丰富了人们对湍流及其拟序结构的认识. 自 Corrsin和Kistler (1954)在湍流尾流中发现流向速度脉动的间歇性、 Kline等(1967)通过采用氢气泡技术展示出平板湍流边界层近壁重复出现的高低速条带并将其称之为湍流的拟序结构后, 研究者们逐渐认识到湍流并非完全不规则的随机运动, 而且湍流拟序结构在湍流的脉动生成和演化、能量的输运和耗散中都起着重要作用(Cantwell 1981, Robinson 1991). 如: 马蹄涡对湍动能输运起着重要作用(Theodorsen 1955), 其引发的喷射现象在增加雷诺剪切应力的同时, 还将导致壁面摩擦阻力的增加(Offen & Kline 1975). 又如: 湍流拟序结构的产生还决定着高超声速飞行器的气动加热水平、航空发动机性能的改变、潜艇噪声的产生等. 因此, 20世纪50年代后, 研究者们对湍流拟序结构的形态和尺度以及起源和影响等一直给予高度关注. 在较低雷诺数流动情形, 一些湍流拟序结构陆续被发现, 如: 各向同性湍流中的涡管或涡片、钝体绕流尾迹中的涡街、热对流中的羽流、声波的波包等, 以及边界层中的条带结构、发卡涡(或马蹄涡)、类孤立波(李存标 2009). 针对边界层湍流, Smits等(2010)认为主要有以Kline 等(1967)的发现为代表的流向尺度约为$1000v/u_\tau$、展向间距约$100v/u_\tau$的近壁面条带和以Theodorsen (1952)的推测为代表的最小尺度为约$100v/u_\tau$的发卡涡或者马蹄涡, 以及以Kovasznay 等(1970)的发现为代表的最大流向尺度可达$(2\sim 3)\delta$的大尺度运动(large scale motions, LSMs)和以Kim & Adrian (1999)的发现为代表的流向尺度大于$ 3\delta$或甚至$(10\sim 15)\delta$的VLSMs. 可见, VLSMs是湍流拟序结构家族的最新成员, 而且由于拟序结构的尺度越大其主导湍流输运的作用越明显, 因此, VLSMs的发现引起研究者们的高度关注. 除此之外, VLSMs还一直被认为是在湍流特征雷诺数较高情形才出现的一种拟序结构, 如: 在管道流中发现VLSMs的$Re_\tau=1058\sim 3175$ (Kim & Adrian 1999)、在槽道流中发现VLSMs的$Re_\tau=3178$ (Monty et al. 2007), 而在湍流边界层实验中发现VLSMs的分别是$Re_\tau=1476\sim 2395$ 和$Re_\tau=1120\sim 19960$ (Balakumar & Adrian 2007, Hutchins & Marusic 2007a). 对于大气表面层, 尽管气象学家们在其野外观测中发现过近地层存在近壁涡(Drobinski et al. 2004), 但对VLSMs较为严格和精细的观测和分析是由SLTEST的一系列实验给出的. 基于SLTEST的数据, Guala 等 (2011)Hutchins 等 (2012)发现在$Re_\tau=5.0\times 10^5\sim 7.7\times 10^5$ 的大气表面层净风条件下存在VLSMs并分析了它的调制作用. 因此, VLSMs被认为是HRNWT的特征之一.

QLOA数据发现大气表面层含沙风场存在VLSMs而且PM10沙尘浓度场中也存在类似VLSMs的结构, 即沙尘超大尺度结构. 本文作者团队基于QLOA的实时高频观测数据, 通过对流向风速相关系数云图分析和谱分析, 不仅证实了在大气表面层净风流动中存在VLSMs (Wang & Zheng 2016, Liu et al. 2017a), 而且首次报道了沙尘暴情况下的$Re_\tau\approx 5.0\times 10^6$的含沙流场中也存在VLSMs (Wang et al. 2017, Zheng 2018). 沙尘流场脉动信号瞬时分布, 见图5(a), 清晰展示出的VLSMs与Hutchins和 Marusic (2007a)在SLTEST的观测结果非常相似, 其展向如蛇状蜿蜒摆动, 其流向尺度超过$3\delta $长达1.3 km. 在将PM10沙尘浓度的时间相关函数换算到空间上后, 发现在相应的浓度场中也存在与含沙流场VLSMs尺度相当的沙尘超大尺度结构(顾海华, 郑晓静 2019), 见图5(b), 并得到LES结果(Zhang et al. 2018)的证实. 通过对中性层结大气表面层悬移粉尘浓度分布的分析发现这种粉尘浓度场中的超大尺度结构是由含沙流场中VLSMs引起的上抛事件导致. 值得指出的是: 目前已有利用点测量速度脉动研究VLSMs尺度特征的工作与本文这里给出的大气表面层净风场和含沙场VLSMs以及沙尘浓度场中沙尘超大尺度结构的流向尺度均是采用了泰勒冻结假设将风速或浓度脉动时间序列换算为脉动的空间信号, 其中将湍流结构和PM10颗粒的对流速度用局地平均流体速度来代替. 与此同时, 本文作者团队利用QLOA的流向阵列, 首次直接测量得到了近中性大气表面层VLSMs平均流向尺度, 实测结果及其与利用泰勒冻结假设换算得到结果的比较见表1.

图5

图5   (a)基于QLOA数据得到的含沙尘流动$({Re}_\tau=5.4\times10^{6}$, 5 m高度处PM10平均浓度$C=1.09$ mgm$^{-3}$, 平均风速$U =13.1$ ms$^{-1})$的VLSMs以及(b) PM10浓度场中超大尺度结构 (粗黑虚线为相关系数${R}=0.05)$


表1   不同流动类型仿真策略对比

新窗口打开| 下载CSV


通过泰勒冻结假设对实测数据时空换算得到的大气表面层VLSMs的流向长度将被低估. 由于VLSMs的流向尺度较大, 尽管有一些在实验室得到的VLSMs流动图像是通过PIV空间拍摄得到的(Hutchins & Marusic 2007a, Wang et al. 2018), 但受限于目前相机CCD的尺寸及像素密度, 很难进行更高雷诺数条件下的更大尺度高分辨率的测量, 尤其在大气表面层中这种测量手段难以应用, 因此, VLSMs流向尺度的确定大多仍基于对风速脉动的时间序列信号借助泰勒冻结假设转换得到. 这种转换的前提一是湍流结构的迁移速度与所在位置的平均流速一致, 二是结构在迁移过程中不发生变形. 由于这些前提在实际的流动, 特别是边界层流动中很难完全满足, 因此泰勒冻结假设的适用性成为湍流研究的一个热点(Squire et al. 2017), 这方面研究进展的系统总结请见He 等 (2017). 本文作者团队基于QLOA的数据分析发现: 尽管由风速脉动时间序列经泰勒冻结假设换算得到的流向速度二阶结构函数与空间实测结果在惯性区没有显著差异, 但直接测量得到的VLSMs的流向尺度普遍大于由风速脉动时间序列利用泰勒冻结假设换算得到的尺度, 见表1, 最大相对误差可超过30% (Han et al. 2019b); 大气表面层对数区中拟序结构的迁移速度随迁移距离变化, 见图6(a), 并高于当地的平均速度, 在2.5 m~5 m范围内涡结构平均迁移速度比局地平均速度高约14%, 这将导致利用泰勒冻结假设换算流向速度时空互相关函数存在较大误差. 大气表面层的VLSMs沿流向迁移其自身尺度1.6倍的距离后已发生了完全的变形, 见图6(b). 因此, ASL中VLSMs流向尺度误差产生的原因主要是泰勒冻结假设使用的前提没有得到满足; 鉴于何国威等提出的椭圆模型(He & Zhang 2006)在考虑湍流结构变形后显著改进了泰勒冻结假设对时空互相关函数的估计, 针对ASL中的VLSMs, 提出将何的椭圆模型中的变形速度$V_{\rm t}$取为$3V_{\rm t}$, 修正后的椭圆模型得到的时空互相关函数与实测结果的吻合度明显提高, 见图7.

图6

图6   基于QLOA数据给出的(a)迁移速度随迁移距离的变化;(b)相干谱分析, 其中$U _{\rm m}$ 和$U$分别为QLOA实测迁移速度和局地平均速度, $\Delta x$为流向距离, $f$为频率(郑晓静 2017)


图7

图7   基于修正后的椭圆模型给出的时空互相关函数与实测结果的比较(郑晓静 2017)


雷诺数对VLSMs流向、展向和垂向尺度以及流向方向与壁面间所存在的倾角的影响一直不清楚. 与近壁条带和马蹄涡(或发卡涡)主要出现在近壁黏性和过渡区内不同, VLSMs是在整个边界层内都存在, 因此, 揭示其三维尺度随高度的变化规律是非常重要的. VLSMs的倾角决定着湍流输运进而包括热量、质量、动量和能量的输运路径(Marusic et al. 2010c, Jacob & Anderson 2016), 同时, 对大涡模拟亚格子尺度耗散中的正向传输和逆向传输(Carper & Porté-Agel 2004)以及大涡模拟壁函数(Marusic et al. 2001)有着至关重要的作用. 早期, Tritton (1967), Moin 和 Kim (1982)分别在$Re_\tau=800$和640的湍流边界层和槽道流中通过一维两点相关分析, 发现LSMs的展向尺度随高度一直增加, 但流向尺度在对数区以外的尾流区随高度减小, 而Krogstad 和 Antonia (1994)在$Re_\tau=1850$的湍流边界层结果则是流向尺度随高度以双曲正切的规律增加. 对于VLSMs, 不同学者对不同情况下给出的三维尺度及其随高度变化的规律也各不相同. 如: Tomkins 和 Adrian (2003)Hutchins 等(2005)的$Re_\tau=690\sim 2800$的湍流边界层结论是VLSMs的流向尺度在对数区以近似线性的规律增加, 展向尺度在整个边界层中均以线性的规律增加, Monty 等 (2007)利用在管道流$Re_\tau=1000\sim 4000$和槽道流$Re_\tau=3100$中的热线多点测量结果并结合已有湍流边界层结果, 给出了$Re_\tau \sim O(10^3\sim 10^4)$时不同类型的流动中VLSMs展向尺度在尾流区随高度分段线性增加的规律. 至于拟序结构的倾角, Kovasznay 等 (1970)在用热线风速仪测量$Re_\tau=1240$的零压力梯度湍流边界层风速时, 通过空间两点相关分析, 不仅发现了LSMs, 而且还发现这种结构沿流向与壁面间存在一个明显的倾斜角度. 这个发现很快得到Blackwelder 和 Kovasznay (1972) 以及 Falco (1977)的边界层流动实验的证实, 随后的研究就集中在这种结构倾角的大小方面. 仅就湍流边界层而言, 对于LSMs的倾角, 就有$Re_\tau=3413$时为$18^\circ$ (Brown & Thomas 1977), $500<Re_\theta<17500$时是$15^\circ\sim 20^\circ$ (Head & Bandyopadhyay 1981)、$Re_\tau=2227$时是$12.3^\circ$ (Tomkins 1997)等; 对于VLSMs的倾角, 有$Re_\tau=355$, 836, 2000时是$3^\circ\sim 35^\circ$ (Adrian et al. 2000), $Re_\tau=1.3\times 10^6$时是$18.7^\circ$ (Hommema & Adrian 2003), $Re_\tau=6.0\times 10^6$和$Re_\tau=5.0\times 10^5$时是$11^\circ$ (Morris et al. 2007, Guala et al. 2011), $Re_\tau=7.7\times 10^5$时是$25^\circ$ (Hutchins et al. 2012)等. 尽管结果各不相同, 非常分散, 但大致上, $Re_\tau\sim O(10^3)$的低雷诺数时的结构倾角范围约为$3^\circ\sim 35^\circ$, 而$Re_\tau\sim O(10^6)$的高雷诺数时则约为$11^\circ\sim 25^\circ$. Marusic 和 Heuer (2007)认为结构倾角不随雷诺数变化, 因为即使是在低雷诺数的风洞实验$(Re_\tau =1350)$和流动尺度完全分离高雷诺数的近中性大气表面层实验$(Re_\tau \approx 1.8\times 10^6)$, 得到结构倾角分别为13.8$^\circ$和14.4$^\circ$, 没有显著差别. 总之, VLSMs的三维尺度及其倾角是否受雷诺数影响? 是否存在相对普适的变化规律? 主导变化的关键因素是什么? 一直没有结论.

基于QLOA的更高雷诺数和更大雷诺数范围的研究深化了对VLSMs尺度和倾角的认识. 本文作者团队通过分析在更高的雷诺数条件下以及更广的雷诺数范围$Re_\tau \sim O(10^3\sim 10^6)$内不同高度处流向风速脉动的沿流向、展向及垂向方向的一维两点相关, 揭示出VLSMs三维尺度随雷诺数及高度变化的相对普适性的规律(Liu et al. 2017a), 具体表征形式见图8.

图8

图8   基于QLOA数据得到的VLSMs展向和垂向尺度随基准点高度变化的规律,净风结果引自Liu 等 (2017a)


净风场中VLSMs的流向尺度在对数区中随高度近似满足指数增加规律, 含沙流动中VLSMs流向尺度在对数区中下部略低于净风流场, 但随高度的变化规律与净风场一致, 见图8(a); VLSMs的展向和垂向尺度在整个边界层中随高度的变化定性一致, 即: 以分段线性的规律增加且在对数区中的增加明显快于尾流区, 见图8(b)和图8(c); 沙尘暴流场中沙尘超大尺度结构的流向和垂向尺度与净风场情形类似, 但在近壁面粉尘结构的流向尺度明显要大一些, 见图8(a)中的红色空心圆点. 这是因为在较高处的沙尘具有与流场含能大涡(尤其是大尺度拟序结构)较好的跟随性, 而在地表处, VLSMs对地表较小尺度结构的调制使得地表剪切作用加强, 促进了地表粉尘的释放, 进而导致近地表处的粉尘结构尺度比净风场VLSMs的尺度大; 含沙流场中的VLSMs和粉尘结构也存在倾角, 其随剪切风速增加而减小的规律与净风场情形类似, 但倾角明显要大一些, 这是由于含沙流动中大量颗粒的存在使流场中的速度梯度减小所致; VLSMs倾角随剪切风速的增加依近似线性的规律减小, 这对中低雷诺数情形的LSMs也是适用的, 见图9. Liu 等(2017a)还给出剪切风速主导倾角大小的机理解释, 即: 由于VLSMs和LSMs的结构倾角实际上是由多个沿流向排列发卡涡的涡头连线与壁面的夹角(Adrian et al. 2000), 而在同一个发卡涡包中, 较早形成的较大发卡涡的涡头距壁面更高, 比后来形成的较小发卡涡具有更快的迁移速度(Dennis 2015), 流向风速沿垂向的速度梯度使得VLSMs和LSMs被拉伸. 速度梯度的增加, 拉伸作用增强, 结构倾角减小. 剪切速度反映了边界层中的速度梯度的大小, 因此能够很好的表征结构倾角的变化规律. 本文作者团队通过对高雷诺数情形下VLSMs尺度及倾角的研究, 提炼出了具有普适性的规律, 使得已有研究给出的分散度较大的结果得到有效统一和合理解释.

图9

图9   VLSMs倾角随摩擦速度的变化, 其中的净风结果取自Liu 等(2017a)


5 VLSMs的起源和影响

由于VLSMs的尺度和所携带的能量占湍流总动能的比例较大, 因此, 它的起源和影响是认识和理解这一现象自身的关键, 也是HRNWT应用研究的重要基础. 关于VLSMs的起源, 一直以来存在"自下而上"和"自上而下"的两种截然不同的观点; 关于VLSMs的影响, 本文将主要集中在它对较小拟序结构的调制和对自然界沙尘的输运方面.

"自下而上" (Bottom-Up)机制是VLSMs起源的主流观点. Kim 和 Adrian (1999)在$Re_\tau=1984$管道流动中发现VLSMs后认为: VLSMs是由一系列近壁发卡涡包组成的, 是由近壁向对数区发展的, 即"Bottom-Up"机制. 这一观点随即得到$Re_\tau=2000$边界层流动实验的支持(Adrian et al. 2000), 进而被广泛的用于解释VLSMs的起源(Liu et al. 2001, Marusic 2001, Monty et al. 2007, Wu & Moin 2009). 不仅如此, 这一观点似乎还适用于中低雷诺数情形. 如: Lee 等 (2014) 通过分析$Re_\tau=930$的时间解析槽道湍流DNS数据, 发现外区流向尺度大于$3\delta$的超大尺度结构主要由尺度小于3$\delta$ 的大尺度结构合并产生; 又如: Deng 等(2018)用本征正交分解(proper orthogonal decomposition, POD)方法分析$Re_\tau=380\sim 2000$的零压力梯度湍流边界层流向-法向平面二维速度场, 给出了发卡涡自组织为中等尺度的发卡涡包的统计学证据, 也在时间-法向平面上观察到发卡涡包沿流向排列形成VLSMs的瞬时事件, 都进一步支持了中低雷诺数下的"Bottom-Up"机制. 与此同时, 与"Bottom-Up"机制相悖的现象也在HRNWT陆续被发现. 如: $Re_\tau=1.01\times 10^5$ 的管道实验(Morrison et al. 2004)发现外区的大涡对近壁湍动能的产生有重要贡献, $Re_\tau\sim 1.7\times 10^5$的管道实验(Zhao & Smits 2007)也证实了这一点. ASL观测结果也对"Bottom-Up"机制提出质疑. 先是发现VLSMs对内区的小尺度运动有显著的调制作用, 而且在完全粗糙的大气表面层中仍存在的VLSMs, 这说明VLSMs可能是剪切驱动的而不依赖于近壁发卡涡的组织过程(Hutchins & Marusic 2007a); SLTEST $Re_\tau=5.0\times 10^5$的数据分析发现: "Bottom-Up"机制仅在近壁区$z^+=O(10^3)$适用(Guala et al. 2011). 直观上也很难理解ASL涡结构能依循"Bottom-Up"机制从近壁的毫米量级增长到外区的千米量级(Hutchins et al. 2012).

"自上而下" (Top-Down)机制是对VLSMs起源的一种猜测. 相对"Bottom-Up"机制, 一些学者提出了在高雷诺数的流动中VLSMs在外区生成并向下运动的"Top-Down"机制(Hunt & Morrion 2000, Hunt & Carlotti 2001, Högström et al. 2002), 并得到部分较低雷诺数情况下的数值模拟结果间接支持. 如: $Re_\tau=1901$ 的DNS结果(Álamo et al. 2006)发现附着于壁面的涡包在有背景速度脉动的情况下存活的时间很短, 不可能发展到太高的壁面高度, 而的确发现了在更高处的涡包, 由此推断涡包是"自上而下"的; 又如: $Re_\tau=674$槽道DNS结果(Flores et al. 2007)显示光滑和粗糙两种壁面条件对附着涡包的条件平均流场也都有相似的低速动量区, 进而推断湍流边界层外区的涡包结构是在远离壁面处产生的. 然而, "Top-Down"机制也遭遇挑战. 如: $Re_\tau=1060$ 湍流边界层PIV实验数据(Ganapathisubramani et al. 2005)发现在$z/\delta=0.1$处存在流向尺度为$1.5\delta$的LSMs, 然而在$z/\delta=0.5$处其流向尺度仅为$0.6\delta$, 这是VLSMs "Top-Down"机制无法解释. $Re_\tau=7959$的管道实验(Guala et al. 2006)发现壁面以上直至边界层约一半的高度处$(z/R=0.5)$的VLSMs流向尺度随高度线性增加, 而超过$z/R=0.5$后则没有VLSMs. 类似的现象在$Re_\tau=1584$槽道流动和$Re_\tau=2395$边界层流动中也被发现 (Balakumar & Adrian 2007). 中低雷诺数壁湍流流动中的外区顶部不存在VLSMs, 也就谈不上"Top-Down"了. 另外, "Top-Down"机制显然不能够解释近壁条带结构的产生. 由此推测, VLSMs的起源机制与雷诺数有关, 而"Bottom-Up"机制可能更多地适用于雷诺数较低的流动 (Smits et al. 2011).

QLOA观测数据首次给出"Top-Down"机制的直接依据. 本文作者对QLOA观测数据的分析揭示了VLSMs能量随高度的变化以及能量在谱空间的分布与中低雷诺数流动情形有显著差异(Wang & Zheng 2016). ASL中VLSMs能量及能量分数, 即: VLSMs能量占总湍动能的比例, 均随着高度增加而增大, 而中低雷诺数的实验结果则显示虽然VLSMs的能量分数有着与大气表面层类似的随高度增加, 但其能量本身却随着高度的增加而减小. 这一定性差异也反映在由预乘谱中能量沿尺度分布的"翻转"中, 如图10(a)所示, 即: 中低雷诺数边界层流动的预乘谱(图10(b))中各湍流尺度的能量均随着高度增加而降低, 而ASL 流动高波数区能量随高度降低但低波数区随高度增大. 前者可以理解为能量在由近壁产生向上传递的过程中逐渐衰减, 即"Bottom-Up"过程, 而后者则可以理解为大尺度能量自上而下的衰减, 即"Top-Down"过程. 考虑到ASL流动的特殊性, 表面层以上的Ekman层依旧是湍流状态(理论上实验室边界层流动外自由流为层流), 依能量层间交换的观点, 表面层外能量的输入是 "Top-Down"机制的原因, 在能谱上即表现为低波数区能量随高度增加. 这一能谱"翻转"现象在本文作者团队的风洞边界层实验和半槽流动的数值模拟中均未发现, 这更进一步说明: 正是因为大气表面层上部存在自上而下的能量输入, 而风洞和数值模拟的上边界没有能量输入, 才引起流向湍动能能谱的"翻转". 因此, 边界层外存在能量输入可能是引起能谱"翻转"的主要因素. 考虑到大气表面层近壁依旧存在强烈的能量产生, 无法排除VLSMs存在自下而上的产生过程, 所以大气表面层中VLSMs可能存在两种机制的共同作用.

图10

图10   流向风速脉动的预乘谱, 引自Wang 和 Zheng (2016)


VLSMs对湍流能量、雷诺应力以及物质输运等的贡献显著大于其他拟序结构. 以流向湍动能为例, Kim 和 Adrian (1999) 在其发现VLSMs伊始, 通过分析$1058<Re_\tau<3175$的管道实验结果的能谱, 就注意到在外区存在与VLSMs对应的能谱峰值, 由此意味着VLSMs对流向湍动能贡献显著. 随后的研究结果不仅定性上证实了这一点, 而且给出了VLSMs流向湍动能在湍流总能量中所占的比例. 如: 对$1500\leq Re_\tau\leq 1.01\times 10^5$管道流动的数据分析结果(Morrison et al. 2004)显示VLSMs流向湍动能占比可达50%, 在$3815\leq Re_\tau\leq 7959$的管道流结果(Guala et al. 2006)认为最高可达65%且VLSMs对雷诺应力的贡献率也占到$50%\sim 60%$; 又如: 对$1476\leq Re_\tau\leq 2395$的湍流边界层和$531\leq Re_\tau\leq 1584$的槽道流中的实验结果(Balakumar & Adrian 2007)认为VLSMs对湍动能和雷诺应力的贡献分别约为$40% \sim 65%$和 $30%\sim 50%$, 并具有随雷诺数增加而增加的雷诺数效应. 近年来的数值模拟结果支持了这一观点, 如: 对$459\leq Re_\tau\leq 732$湍流边界层的DNS模拟(Lee & Sung 2011)发现VLSMs的雷诺切应力贡献率超过45%, 而对$Re_\tau \sim 4000$的湍流边界层的DES结果(Deck et al. 2014)指出VLSMs/LSMs显著影响壁面切应力, 其贡献了超过60%的湍流摩擦系数(即: 通过雷诺应力加权积分得到的湍流对平均剪切的贡献). VLSMs如此之高的湍动能占比对物质输运的影响显然也是十分关键的, 特别是在ASL, VLSMs对动量的向下输运起到主要作用 (Horiguchi et al. 2012), 这将显著影响水汽、二氧化碳的传输(Cooper et al. 2006, Serafimovich et al. 2011).

VLSMs对近壁小尺度湍流脉动幅值有很明显的影响. 这种影响, 也称之为对小尺度湍流脉动幅值的调制 (Hutchins & Marusic 2007b), 有助于对湍流形成与发展的理解(Mathis et al. 2009a), 成为近壁风速和壁面剪切应力预测模型的基础和前提(Marusic et al. 2010b), 并对流动控制有指导作用(Deng et al. 2016). 有关湍流脉动幅值调制的研究可分为现象揭示、定量化研究、应用与推广三个阶段. Brown和 Thomas (1977)在$Re_\tau=3413$的湍流边界层中通过对比流向风速的低频和高频脉动信号发现: 大幅值的高频(小尺度)脉动会出现在低频(大尺度)脉动的峰值附近, 这表明低频脉动与高频脉动的幅值是有关联的. 通过对边界层、混合层、射流等进行速度脉动的尺度分解(截断频率100 Hz), Bandyopadhyay 和 Hussain (1984)发现: 不同形式的剪切流动中低频脉动和高频脉动部分之间均有很强的相关性. Hutchins 和 Marusic (2007b)通过对湍流边界层中$(Re_\tau=7300 )$用热线测得的风速脉动信号的低波数脉动和高波数脉动进行对比, 系统描述了幅值调制现象, 即: 流向风速大尺度正的脉动使小尺度运动的幅值增加而负的脉动使小尺度运动的幅值减小. 在调制现象定性描述的基础上, 研究者们将原始信号以一定的截断波长(记为: $\lambda_{\rm c} )$分解为大尺度和小尺度脉动, 通过计算大尺度运动和小尺度运动的包络线之间的相关系数, 即脉动幅值调制系数的大小来定量表征调制作用的强弱$(R_{\rm AM})$. 利用这一定量化方法, Mathis 等 (2009a)通过对实验室湍流边界层$(Re_\tau=2800\sim 19000)$和大气表面层$(Re_\tau=6.5\times 10^5)$的测量数据分析, 最早给出近壁脉动幅值调制系数最高可达0.6并给出调制系数随高度和随雷诺数变化的规律, 具体是: 随高度的增加从近壁开始减小直至在近壁对数区中心处附近为零后在对数区中上部出现负值; 随雷诺数的增加在过渡区$(20<z^+<100)$以近似服从对数线性的规律增加. 调制系数为"负值", 也即在对数区中上层出现"反转", 意味着流向风速大尺度正的脉动使小尺度运动的幅值减小而负的脉动使小尺度运动的幅值增加. Mathis 等 (2009b)通过分析以截断尺度$\lambda_{\rm c}=\delta$得到的$Re_\tau\approx 3000$的边界层流动、管道流和槽道流中幅值调制系数发现: 三种流动类型中的幅值调制系数仅在外区有微小的差异, 而在内区完全一致, 也就是幅值调制系数与流动类型无关. 调制系数也会随着壁面粗糙度的增加而增加, 其影响随着壁面距离增大而减小 (Squire et al. 2016, Pathikonda & Christensen 2017). Schlatter 和 örlü (2010)通过对比$800<Re_\tau<5500$情况下的幅值调制系数$(\lambda_{\rm c}=\delta)$和流向风速的偏度系数, 发现两者具有很好的线性关系, 由此表明: 流向风速的偏度这一统计量在某种程度上是幅值调制作用的一种反映, 是可以与幅值调制系数一样用来量化VLSMs对较小尺度结构幅值调制的强弱程度的. 通过将偏度系数分解, 并逐一和幅值调制系数$(\lambda_{\rm c}^+=7000)$进行对比, Mathis 等(2011a) 建立了在$Re_\tau=2800\sim 19000$范围内利用偏度系数分析调制作用强弱的方法. 最近, Yao等(2018)在分析不同雷诺数$(Re_\tau=540$, 1000, 2000)槽道湍流中大尺度结构$(\lambda_{\rm c}>\delta)$对近壁湍流调制作用时又发现: 近壁区的极端回流事件与外区大尺度运动的调制密切相关. 当雷诺数升高时, VLSMs对小尺度湍流脉动的调制效应增强, 此时近壁区极端事件出现的概率也增大, 即垂向脉动速度概率密度分布的尾部上翘, 这意味着近壁垂向速度脉动概率密度分布尾部的上翘程度也可以用来反映调制作用的强弱, 只是尚未建立起二者间定量的对应关系. 除了对流向幅值的调制外, 研究发现VLSMs对较小尺度结构展向和垂向运动也有调制作用, 其对展向和垂向运动的幅值调制系数在$ Re_\tau=15000$的湍流边界层情形基本一致(Talluru et al. 2014)并高于压力脉动中大尺度对小尺度的调制作用(Tsuji et al. 2016). 调制作用的定量描述为近壁湍流信号的预测提供了有效途径. 将标定实验得到的近壁小尺度普适速度信号基于调制系数进行脉动幅值修正并考虑外区大尺度结构的线性叠加作用, 研究者们分别建立了近壁风速预测模型(Marusic et al. 2010b, Mathis et al. 2011a)和壁面剪切应力预测模型(Marusic et al. 2011, Inoue et al. 2012, Mathis et al. 2013), 并被推广到对法向和展向速度的预测(Yin et al. 2018). 这些模型利用对数区实测的大尺度风速脉动信号来预测近壁风速和壁面剪切应力脉动信号, 进而避免了因近壁难以直接测量无法获得近壁区风速和剪切应力的问题, 可用于验证及改进现有的LES数值模型.

本文作者团队给出了雷诺数$Re_\tau\sim (1.0\sim 5.0)\times 10^6$大气表面层VLSMs能量占比随高度的变化和对不同尺度的拟序结构的调制作用. 由于VLSMs在整个外区普遍存在, 因此, VLSMs的能量占比随高度的变化是非常重要的, 却一直缺乏研究. 基于QLOA的观测数据, 本文作者团队的分析结果指出: 大气表面层净风场和含沙风场的VLSMs对雷诺应力的贡献率在对数区上部也可达50%, 其能量占比在对数区可超过60%并且随高度的增大均服从近似对数线性规律, 这与尺度在$0.3\delta$到$3\delta$的LSMs能量占比在$30%\sim 50%$之间并随着高度增加略有减小有明显不同(Wang & Zheng 2016), 见图11. 由图11可见, VLSMs的能量占比并不是在大气表面层一直高于LSMs, 而是在对数区的某一高度以上其能量占比高于LSMs, 在这一高度之下, VLSMs的能量占比是低于LSMs 的. 对VLSMs能量占比研究的另一个十分有意义的工作是对 "阵风"本质的揭示. 世界气象组织将周期为1至数分钟的脉动风定义为"阵风" (World Meteorological Organization 1983), 而曾庆存及其合作者将风速脉动周期为$1\sim 10$ min脉动风定义为中性层结条件下的"阵风", 通过对城市边界层观测数据分析认为阵风具有相干性且对沙尘输运有重要作用(Zeng et al. 2010). 本文作者团队发现气象学界用"阵风"刻画的脉动风与HRNWT刻画的脉动风有较大差异: 前者包含了大量天气尺度的脉动, 而没有包含对湍动能有较大贡献的部分的VLSMs和LSM, 特别是在风速$U>20$ ms$^{-1}$时遗漏了大部分VLSMs和全部LSMs, 这一差异可以从归一化小波能谱中的VLSMs和LSMs以及阵风的尺度对比中清晰可见(Gu et al. 2019), 见图12. 由图12可见, "阵风"对湍动能的贡献明显小于VLSMs, 被"阵风"丢失的VLSMs和LSMs对湍动能贡献达到50%.

图11

图11   VLSMs和LSMs能量分数及雷诺应力贡献随高度的变化,数据取自Wang和Zheng (2016)


图12

图12   QLOA实测风速(2015年4月30日08:00 -- 09:00)脉动信号的小波谱分析, 其中红色箭头及蓝色剪头分别指示归一化能量$E_{uu}$的VLSMs和LSMs区域, 阴影部分代表阵风区域(引自Gu et al. 2019)


关于VLSMs的调制作用, 本文作者团队也给出了更为精细的分析并得到一些新结果(Liu et al. 2019). 一是并非所有尺度的湍流运动间都存在幅值调制作用. 基于QLOA数据的分析发现: VLSMs对小尺度湍流脉动$(\lambda_x<0.3\delta)$运动有着显著的幅值调制作用, 其作用的强弱与VLSMs的尺度密切相关; VLSMs对LSMs的调制作用非常弱, 几乎可以忽略; LSMs对小尺度运动的调制作用也可以忽略不计. 二是并非所有尺度大于$3\delta$的VLSMs对小尺度运动幅值的调制作用都相同. 通过改变大尺度部分及小尺度部分的截断尺度进行速度脉动的尺度分解并分析不同尺度间的调制作用发现: 流向尺度大于$28(z\delta)^{1/2}$的VLSMs对流向尺度小于$12z$的小尺度运动幅值的调制作用最强. 由于尺度为$28(z\delta)^{1/2}$为含能最高的湍流运动, 而尺度小于$12z$为能谱中惯性子区和耗散区的非含能结构, 因此, 不是尺度最大的VLSMs对尺度最小的小尺度运动的调制作用最强, 而是含能最高的VLSMs对小尺度运动才有最显著的脉动幅值调制作用, 而介于两者之间的湍流运动对小尺度运动的调制作用以及更大尺度运动的调制作用均可以忽略(见图13). 这一发现除了更为精细地刻画出调制作用最显著的VLSMs和受调制作用最明显的小尺度运动的尺度外, 还为VLSMs的尺度划分提供了一种新判据, 即: 将幅值调制作用最强的流向长度$ \lambda_x>28(z\delta)^{1/2}$结构的定义为VLSMs. 按照这样的判据, VLSMs将与高度有关: 在$z<0.012\delta$区域的$ \lambda_x<3\delta$的结构可能就被视为VLSMs, 而在$z>0.012\delta$区域的部分$ \lambda_x>3\delta$的结构可能还不能认为是VLSMs.

图13

图13   不同大尺度结构$(\lambda_{x}> a\delta)$对不同小尺度脉动$(\lambda_{x} < b \delta )$的幅值调制系数 ${R}_{\rm AM}$的云图及其与流向风速脉动预乘能谱的对比, 其中aB为尺度滤波的参数: (a) $z\approx 0.01 \delta$; (b) $z\approx 0.02 \delta$. 引自Liu et al. (2019)


VLSMs对沙尘输运的影响是VLSMs对物质输运影响的一个重要方面, 以往研究对此的关注严重缺乏. 本文作者团队对此进行了深入研究, 得到一系列新结果. 通过对青土湖地区沙尘暴的观测, 本文作者团队对比分析了沙尘暴流向风速和沙尘浓度的归一化能谱, 发现流向速度与沙尘浓度能谱具有几乎一致的谱峰尺度, 这意味着沙尘暴风场中的VLSMs主导了沙尘的流向输运(Zheng et al. 2013, Zheng et al. 2015). 然而, 利用垂向风速与PM10浓度脉动的相关分析及交叉谱分析则发现, 其相关性随着高度由负相关逐渐变为正相关, 且出现负相关的脉动尺度对应于VLSMs尺度, 由此说明VLSMs对沙尘垂向输运的作用随高度存在差异, 即在近壁风沙跃移层内(约2 m以下)起抑制作用, 而在跃移层以上起促进作用(Wang et al. 2017), 如图14所示. 这是由于流向上低速的VLSMs使得壁面剪切减弱, 不利于地表沙尘的释放, 但垂向上低速的VLSMs具有向上的脉动速度, 有助于高处沙尘的向上传输.

图14

图14   QLOA含沙风场不同高度处的垂向脉动速度与PM10浓度脉动的互谱(Wang et al. 2017)


6 颗粒与HRNWT相互的影响

湍流多相流的实验和数值模拟都要比单相流来的复杂得多, 是更加难以对付的挑战. 自然界和工程中的大部分流动几乎都是多相的, 其中湍流, 特别是剪切流中的壁湍流与颗粒运动构成的两相流尤为普遍, 例如: 河流和风沙流、污染物扩散和煤粉燃烧等, 深刻认识并定量揭示湍流与颗粒相互作用的机理和规律不仅有助于深化对湍流的认知, 而且对灾害预报和提高相关技术水平有指导作用. 就两相流动而言, 依据颗粒体积分数$\varPhi_{\rm v}$ (单位体积内颗粒体积占比)的大小, 经验性地把$\varPhi_{\rm v}<10^{ -3}$的两相流称为稀疏流动而把$\varPhi_{\rm v}>10^{ -3}$的称为稠密流动(Elghobashi 1994). 对于前者, 又可根据颗粒对湍流影响的强弱进一步简化为只需考虑湍流对颗粒影响的单向作用稀疏流动$(\varPhi_{\rm v}<10^{ -6})$和还需要考虑颗粒对湍流影响的双向耦合稀疏流动 $(10^{ -6}<\varPhi_{\rm v}<10^{ -3})$; 而对于后者, 则还需要考虑颗粒间相互作用影响, 称之为四向耦合稠密流动. 就湍流与颗粒相互作用研究而言, 一方面主要针对湍流对颗粒运动行为的影响, 另一方面则聚焦颗粒对湍流的统计量(如: 平均速度剖面、雷诺应力和湍流强度等)和湍流拟序结构的影响. 湍流流相本身的随机性和颗粒扩散相的随机分布, 使得湍流多相流的实验和数值模拟都要比单相流更加困难(Balachandar & Eaton 2010).

湍流施加于颗粒的作用力以及对颗粒运动影响的研究一直在不断深化. 两相流动中流相介质对固相介质的作用力是颗粒运动分析的基础和关键. 除了颗粒旋转引起的Magnus力、颗粒表面的气流速度差引起的 Saffman 力、流体存在压力梯度时产生的压力梯度力、反映颗粒运动历史效应的Basset力等, 具体定义和表征可见Zheng (2009), 流体驱动颗粒运动的拖曳力因是颗粒所受力中最大的而尤为重要, 一直得到普遍关注. 拖曳力的提出始于1851年Stokes的理论分析. 针对层流中的单个颗粒雷诺数$(Re_{\rm p}=|u_{\rm f}-u_{\rm p}|L/v$, $u_{\rm f}$为颗粒周围流体速度, $u_{\rm p}$为颗粒速度, $L$为颗粒特征尺度)远小于1的球形颗粒, Stokes认为流体作用于颗粒的拖曳力与颗粒直径、流体动力黏度及颗粒与流体的速度差呈正比. 这一特性反映在随后的拖曳力计算公式中, 即拖曳力系数与$Re_{\rm p}$有关. 然而, 一方面在湍流中, 拖曳力的平均值和脉动值都正比于湍流度(Bagchi & Balachandar 2003, Kim & Balachandar 2012, Homann et al. 2013), 这意味着壁湍流中颗粒的拖曳力系数可能与流动雷诺数有关; 另一方面 颗粒的形状(Chhabra et al. 1999)和数量也会对拖曳力系数有影响, 而且还随颗粒体积分数的增加而增大(Kaye & Boardman 1962, Helland et al. 2005). 湍流的流动形式和两相流的颗粒浓度还会影响颗粒的相对运动速度. 以壁湍流为例, 目前主要的结论大致有: 小颗粒低浓度中的颗粒相平均速度一般小于流体平均速度(Kaftori et al. 1995, Taniere et al. 1997), 但在近壁区会有例外(Righetti & Romano 2004) 颗粒与湍流直接的动量交换会随颗粒尺寸增大而增强导致大颗粒的平均速度大于流体平均速度(Wang & Levy 2006); 颗粒的流向速度脉动通常比流体的脉动强(Zhou et al. 1994, Zhou et al. 2001)但法向脉动速度小于流体的脉动(Kulick et al 1994, Wang et al. 1996). 颗粒在近壁面还会发生优先富集的"涡泳"现象(Caporaloni et al. 1975), 这主要与湍流拟序结构的上抛和下扫运动密切相关(McLaughlin 1989, Kaftori et al. 1995, Marchioli & Soldati 2002), 且这种颗粒优先富集区域的颗粒不是均匀分布, 较多的颗粒聚集在低速、高涡量的流动结构中形成带状分布(Ninto & Garcia 1996, Pan & Banerjee 1996).

颗粒对壁湍流统计特性的影响主要反映在对壁湍流的平均速度以及其整体湍流强度、雷诺应力的调制方面. 颗粒对湍流平均速度剖面影响主要与颗粒惯性, 颗粒与颗粒、颗粒与壁面的碰撞有关, 目前的基本共识是颗粒对流体平均速度的影响是使其在近壁面区域增大而在离壁面区域减小. 对于湍流的脉动, Owen (1969)在水平管道实验中最早指出: 当颗粒的弛豫时间(颗粒对流动的响应时间)小于湍流含能涡特征时间时, 颗粒相对流体运动的滞后导致对湍流强度有抑制作用. 更为定量的分析结果给出的是: 大颗粒 (3.4 mm) 和小颗粒 (0.2 mm)会分别增强和抑制整个管道区域流体的速度脉动, 而中等颗粒 (0.5 mm, 1 mm)则在管道中心区域起增强而在近壁区域起抑制流体速度脉动的作用(Tsuji & Morikawa 1982, Tsuji et al. 1984), 这在$Re_\tau<1000$的槽道流中也基本如此(Rashidi et al. l990, Rogers & Eaton 1991, Kussin & Sommerfeld 2002). 除了颗粒的大小, 两相流中颗粒相与流体的体积分数或质量分数也是一个重要参数. 较大的颗粒对流相湍流强度增强的作用会随着颗粒体积分数的增大变得更加显著(Liljegren 1990, Varaksin 2000, Zhang et al. 2008), 而较小的颗粒(20 $\mu$m)对湍流强度的影响则对质量分数更加敏感(Li et al. 2001). 颗粒对湍流脉动的影响还与颗粒的Stokes数、颗粒雷诺数、颗粒湍流尺度比、颗粒体积分数、颗粒动量数等无量纲参数有关. 对于颗粒$St$数, 有学者认为大于60会增强湍流强度, 反之减弱(Luo et al. 2005, Elgobashi et al. 2006, Tanaka & Eaton 2010), 但也有研究发现可能还与颗粒湍流尺度比以及密度比有关而不仅仅与颗粒$St$数有关 (Lucci et al. 2011); 对于颗粒雷诺数, 有研究认为大于400会增加湍流强度 (Geiss et al. 2004), 但Mandø (2009)的实验却发现$Re_{\rm p}=600$的颗粒也会减小湍流强度; Tanaka 和 Eaton (2008)提出用颗粒动量数$(Pa_{\rm St})$作为判据, 发现$Pa_{\rm St}$在$10^3\sim 10^5$范围内的颗粒削弱湍流, 之外则增强湍流强度. 另外, 颗粒的形状也对湍流强度具有不同的影响, 林建忠等(2002)的槽道两相柱状颗粒数值模拟发现, 相比于球形颗粒柱状颗粒对湍流强度的抑制作用更强, 且抑制程度随颗粒的长径比增加而增大. 颗粒对湍流强度影响在不同壁面位置也会不同, 如针对$Re_{\tau}\approx 650$槽道 (Righetti & Romano 2004)和$Re_{\tau}\approx 100$边界层(Li et al. 2016)的DNS两相流模拟指出: 流向和垂向湍流强度在外区受到抑制而在内区得到加强. 然而, $Re_{\tau}\approx 1000$管道两相流动实验(Ljus et al. 2002)则发现颗粒对湍流强度的影响还会因外区的不同位置而不同: 颗粒在外区的对数区削弱而在管道的中心区增强湍流强度. 更为重要的是, 有研究指出颗粒对湍流脉动的影响与流动雷诺数有关, 如Hadinoto et al. (2005)的管道实验研究发现颗粒粒径为200 $\mu$m的颗粒增强湍流强度, 且增强作用随雷诺数增大而增强. 综上可见, 对于颗粒对湍流统计特性的影响不仅是多因素的, 而且其表征也是多参数的, 远未形成共识.

颗粒对壁湍流结构影响的研究相对较少且主要集中在对近壁条带和准流向涡的能量、数量以及尺度的影响方面. 除不断发现两相流中颗粒的存在会使得准流向涡的能量减弱进而导致近壁条带结构强度变弱 (林建忠 1998, Portela & Oliemans 2003)、会增大近壁准流向涡的尺寸而减少流向涡的数量(Dritselis & Vlachos 2008) 并缩小了近壁条带间距(Luo et al. 2017) 等外, 研究者们还发现颗粒对壁湍流结构影响的程度也是有差异的. 一是粒径差异. 如: $Re_{\tau}\approx 150\sim 410$ 槽道两相流中颗粒的大小会增强或抑制壁面低速流体上抛运动 (Rashidi et al. 1990); 又如: $Re_{\tau}\approx 100$对于发展边界层, 小颗粒$(St=10)$增强而大颗粒$(St=50)$削弱近壁条带结构的强度(Li et al. 2016). 二是尺度差异. 如: 圆管两相流实验发现颗粒使得湍流低频大尺度结构能量减弱而高频小尺度结构能量增强(Tsuji & Morikawa 1982, Sato & Hishida 1996). 三是区位差异. 如: $Re_{\tau}\approx 10^3$管道两相流实验发现颗粒在管道中心区域会增强但在近壁区会削弱大尺度湍流脉动(Ljus et al. 2002). 四是流动差异. 如: 颗粒会增大竖直槽道流向涡的尺寸 (Dritselis & Vlachos 2008) 减小水平槽道流向涡的尺寸(Li et al. 2012). 五是流动雷诺数差异. 如: 颗粒减小近壁流向涡的尺寸, 而且这一影响随着雷诺数增大而增强(Richter & Sullivan 2014). 导致颗粒对壁湍流结构影响程度的差异可能还有其他因素, 比如颗粒尺度比、体积分数等, 但颗粒$St$数是根本影响因素之一. 这是因为颗粒对近壁准流向涡的影响与颗粒对流体的作用力的脉动和流向速度脉动乘积${u'}_1^+{f'}_1^+$相关, 小$St$数颗粒会作为能量的"源"而存在, 跟随流场运动, 且${u'}_1^+{f'}_1^+>0$区域与近壁低速条带区域重叠, 使得低速条带不稳定从而增加准流向涡的数量; 随着$St$数的增大, 颗粒会作为能量的"汇"而存在, 且${u'}_1^+{f'}_1^+<0$的区域与低速条带区域一致, 这会使得低速条带结构变得更加稳定从而减少流向涡结构的数量(Lee & Lee 2015). 这里需要指出两点: 一是关于颗粒对外区湍流结构影响的报道很少, 目前仅能见到Tay 等 (2015) $Re_{\tau}\approx 600$ 的水平水槽两相流的实验, 其结果显示颗粒使得外区大尺度结构的尺度和倾角均增大, 这与颗粒减小内区近壁条带的长度及高度但不影响结构倾角有着定性上的不同(Li et al. 2012); 二是目前已有关于颗粒对湍流拟序结构影响的研究主要集中在雷诺数较低$(Re_{\tau}<10^3)$的两相流, 极少有关于高雷诺数情形颗粒与壁湍流相互作用, 特别是颗粒对VLSMs影响的研究.

HRNWT与颗粒的相互作用直接影响颗粒的启动和壁面脉动应力以及颗粒垂向通量的预测. 由于壁面可解大涡模拟(wall-resolved LES, WRLES)求解内区流动进而可以给出更接近DNS结果的湍流场, 本文作者团队采用WRLES对含颗粒的两相流进行较为精确的数值模拟, 其$Re_{\tau}\approx 4000$ (王萍等 2019a), 这应该是目前颗粒两相流WRLES的最高纪录. 数值结果发现: 采用现有各类用于高雷诺数壁湍流的壁模型得到的颗粒两相流wall-modeled LES (WMLES)结果与WRLES在预测颗粒通量时存在可达100%以上的差异. 这是因为已有LES壁模型均是依据单相流动分析建立的, 颗粒与近壁湍流的作用没有被考虑. 通过对壁面应力模型的修正(Yang et al. 2015), 即在积分壁模型(integral wall model, IWM)中引入颗粒体力项, 所得的WMLES结果与WRLES结果的误差降至20%以下. 在此基础上的计算搜索得到在$Re_{\tau}\sim O(10^4)$半槽流动中颗粒流体起动的临界值仅为传统的颗粒流体起动风速$u_{*t}$的70% (Zheng et al. 2020), 且随边界层厚度的增加而减小, 这主要是由于湍流壁面应力脉动所致. 由此揭示了基于传统$u_{*t}$得到的输沙率预测结果与野外观测存在较大的误差(Rasmussen & Sorensen 1999), 即当平均风速小于$u_{*t}$时所测输沙率不为零的原因. 为了便于地学界和工程界使用, 基于QLOA数据, 分别提出了净风和含沙风场的风速表征模型(Han et al. 2019a, 王萍等 2019b). 该模型由平均速度、VLSMs和受VLSMs调制的小尺度三部分组成, 所涉及的系数仅与摩擦风速、动力学粗糙度等常规参数有关. 通过这一模型可以仅由在任一高度, 如5 m处, 的实测风速时间序列预测出其他任意高度处的风速脉动, 所预测的风速时间序列其统计性质和谱结构等均与实测结果有较好的一致性, 见图15(a). 这样在进行实际风场的计算模拟时, 就可以不用进行类似QLOA的大规模测量和WRLES或WMLES也能得到计及高雷诺数效应自然界风场. 依此风速表征模型结合颗粒点力模型, 计算得到2.5 m, 8.5 m和21 m处的不同粒径颗粒的垂向通量随摩擦速度的变化规律, 与QLOA测得的粒径小于10 $\mu$m的沙尘通量结果基本吻合, 见图15(b). 需要指出的是: 现有沙尘暴预报模式的下边界条件之一是2 m以上甚至20 m以上的沙尘垂向通量, 而基于RANS的已有风沙流预测模型所给出的2 m以上的沙尘垂向通量为零, 这显然与实际情况不符合, 其主要原因就是基于雷诺平均的RANS方程无法计及ASL中的HRNWT和VLSMs的影响. 另外, 基于风场表征模型得到颗粒垂向输运通量的计算可在普通计算机上进行, 计算时间比大涡模拟所需时间大大缩短, 同样模拟条件节省计算时长超过90%.

图15

图15   由风速预测模型得到的(a)风速功率谱和(b)计算得到PM10浓度与实测结果的比较(王萍等 2019)


HRNWT与颗粒的相互作用直接影响颗粒两相流中的近壁条带和VLSMs. 基于所建立的计及颗粒两相流中颗粒-颗粒床面碰撞击溅过程的半槽WMLES, Wang 等(2019)模拟了600 m (长) $\times$ 12 m (高) $\times$ 75 m (宽)区域内的$Re_{\tau}=3.2\sim 5.6\times 10^5$时的颗粒(粒径$0.2\sim 0.5$ mm)两相流, 发现: 颗粒两相流中的颗粒在近壁流向、展向的分布不是均匀的, 这也许是诸多风沙流输沙率测量结果间存在差异的一个重要原因, 提示实际风沙流输沙率单点测量的不准确性以及现有颗粒两相流和风沙流的二维模拟(即展向均匀假设)的局限性. 她们的模拟结果还发现: 颗粒相在近壁沿流向会出现一股股尺度大于30 m蜿蜒起伏的"蛇"状结构, 见图16(a), 与风沙流的"streamer"结构(Baas & Sherman 2005)非常类似, 但与QLOA含沙风场中沙尘浓度场的超大尺度结构相比, 其流向尺度偏小而倾角偏大. 将模拟得到的近壁颗粒条带结构的特征尺度与野外"sand streamers"观测结果相比, 发现其吻合尺度远优于Dupont 等 (2013)采用ARPS (the advanced regional prediction system)模拟结果. 其主要原因是可能归结于采用ARPS没能模拟出VLSMs, 而Wang 等 (2019)可以模拟出, 见图16(b). 由于经条件平均后的近壁颗粒条带结构出现在VLSMs近壁面尾迹中, 由此说明这种颗粒条带结构是VLSMs在近壁面的"足迹", 由此不仅揭示出VLSMs对两相流中颗粒运动的影响, 也指出了现有模拟软件的不足; 考虑到颗粒与颗粒床面碰撞的不同壁面过程, 即冲击床面的颗粒只反弹和既反弹又溅起其他颗粒, 后者是自然界风沙流与其他颗粒两相流的根本差异, Wang 等(2019)的模拟结果还发现: 近壁颗粒条带结构在颗粒只反弹而无溅起的两相流中很难长时间维持, 由此反映出壁面过程对颗粒两相流的影响. 为了进一步说明颗粒与颗粒床面的击溅过程对近壁VLSMs的影响, 本文作者团队在$Re_{\tau}\sim O(10^3)$风洞中, 利用大视域平面PIV测量了相同来流风速下分别由平均粒径190 $\mu$m, 粒径范围约$70\sim 350$ $\mu$m满足对数正态分布的沙粒铺成的可侵蚀床面吹起和与由风洞上方投下的相同沙粒所形成的湍流边界层风沙两相流的风速和颗粒速度, 在进行空间相关分析后发现(Zheng 2018): 上述二种情况的流场均出现了VLSMs; 对于前者(即起沙情形), 边界层内不同高度处VLSMs尺度均降低, 对数区VLSMs尺度减小尤为明显, 在对数区底部VLSMs甚至被完全破坏; 对于后者(即投沙情形), 虽然对数区VLSMs尺度也显著减小, 这与起沙情形一致, 但对数区以上区域VLSMs的尺度则明显增大. 这种差异的原因主要是两相流中是否存在颗粒与壁面的作用过程(即粒壁作用), 起沙情形下空中运动的所有沙粒和投沙情形下对数区内运动的绝大多数沙粒均与壁面反复碰撞反弹从而进行持续的跃移运动, 这一粒壁作用对流体而言带来了额外的能量耗散, 使得VLSMs难以维持发生衰减甚至破碎, 如图17所示; 而投沙情形下对数区以上区域内大多数颗粒未发生粒壁作用, 这些沙粒从边界层外更高速的流体获得了能量, 相比边界层内的流体具有更高的速度, 使得在沙粒作用下VLSMs被拉伸增大. 由此表明: 粒壁作用过程直接影响着颗粒两相流中颗粒对VLSMs影响的程度, 这种颗粒对湍流结构影响的现象和机制一直没有得到关注.

图16

图16   (a)沙尘浓度瞬时分布云图, 红色实线标注高浓度区域, (b)流向瞬时脉动风速云图, 红色实线标注高速区, 蓝色实线标注低速VLSMs


图17

图17   对数区底部$z=0.036\delta$ $(z^{+}=72)$处, (a)净风、(b)颗粒地表侵蚀释放两相流、(c)颗粒边界层外释放两相流中湍流结构的对比, 其中白色实线是相关系数$R_{uu}=0.2$的等值线, 在净风流动中其尺度超过$3\delta$为VLSMs


7 结论与展望

HRNWT是在自然界和工业应用领域广泛存在的流动, 对其规律和机制的研究既是认识湍流的基础科学问题又具有重要的应用价值. 然而, 受制于研究手段的不足, 这一领域的研究难度较大, 对HRNWT的认知仍然非常匮乏. 本文在对目前HRNWT研究现状回顾总结基础上, 较为系统地介绍了作者团队在这一研究领域的工作, 尤其是利用QLOA野外观测给出的$Re_{\tau}\sim O(10^{6})$ ASL的净风流动中湍流统计特性及其雷诺数效应、ASL净风场和含沙场中VLSMs的存在和形态特征以及起源机制 和影响规律、平均粒径的沙尘颗粒对卡门常数、湍动能、VLSMs的尺度、倾角及能量的影响、沙尘浓度场中 的超大尺度结构和颗粒两相流中的近壁条带结构、以及颗粒近壁过程对VLSMs的影响等. 这些理论研究对于ASL中风沙流的预测, 特别是已有基于定常风场和稳态跃移假设的风沙流输沙模型(Anderson & Haff 1988, McEwan & Willetts 1991)的改进, 以及风沙物理学研究都具有指导意义.

虽然已有研究已经表明HRNWT出现了许多与低雷诺数情形有所不同的新的流动现象和规律, 但是根据本文作者的理解, 仍然存在需要进一步深入研究的如下关键问题.

(1) 关于HRNWT统计量和VLSMs特征的雷诺数效应. 从目前已有研究结果来看, 在不同流动形式下的壁湍流统计量, 如平均速度、湍动能、雷诺切应力的分布形式及其标度关系等, 均不同程度地显示出与低雷诺数流动情形的差异; 不同形式流动中出现的VLSMs的特征尺度似乎并不都是随雷诺数的增加而增大, 如QLOA给出的ASL中VLSMs的湍动能随雷诺数增加而增大, 而其尺度在$Re_{\tau}\sim O(10^{6})$与$Re_{\tau}>2000$的中低雷诺数实验室边界层实验结果差别并不大. 因此, 一方面需要准确揭示HRNWT的各阶统计量和VLSMs的相关特征量随雷诺数增加的变化规律, 另一方面也需要给出高低雷诺数流动出现差异以及统计量和VLSMs的相关特征量不再随雷诺数变化的临界雷诺数.

(2) 关于HRNWT中VLSMs的生成演化机制和影响. 目前关于VLSMs的起源究竟仍然是低雷诺数情况的"Botom-Up"机制还是与低雷诺数情况不一样的"Top-Down"机制, 还是一直没有明确结论. 虽然利用ASL的数据分析可以推论得到VLSMs起源的"Top-Down"机制, 并可以肯定VLSMs起源绝非由壁面自下而上的自组织这一种机制, 但即使是"Top-Down"机制其决定因素也仍不清楚. 至于VLSMs在湍流输运及湍流调制方面的作用规律及机制, 尽管本文作者团队已有一些较高雷诺数情形计及VLSMs影响的颗粒两相壁湍流模拟结果和对PM10以下沙尘输运的测量结果, 但还是比较初步, 对热量、水汽、$CO_{2}$等其他标量的输运作用及其规律的研究也比较缺乏.

(3) 关于颗粒对HRNWT以及对VLSMs的影响. 这方面的研究是Science列出的需要解决的125个科学问题之43"能否发展关于湍流动力学和颗粒材料运动学的普适理论"的一个重要组成部分; 需要特别关注颗粒相对HRNWT和VLSMs影响的关键参数、核心机制和主要规律以及影响的雷诺数效应. 由于难度较大, 这方面的研究目前基本处于尚未开展的状态.

(4) 关于HRNWT的测试. 尽管QLOA的观测极大地推动了HRNWT的研究, 但目前的数据分析还主要针对近中性层结条件的平稳流动. 非平稳流动是一个雷诺数不断变化的过程, 对这一流动现象的准确分析将有助于深化对HRNWT雷诺数效应的认知. 而对非中性层结流动的研究则对传热与HRNWT尤其是VLSMs相互影响以及对湍流输运的规律的揭示和应用十分必要. 对此, 需要更多的实验观测外, 还需要发展可靠的数据处理和分析的方法. 另外, 还需要发展可控条件不同流动形式的单相和两相壁湍流研究的装置和可以实现精准测量的仪器, 这方面我国与国外的差距明显.

(5)关于HRNWT的数值模拟. 在流动中添加颗粒物后, 尤其是不引入颗粒模型的全分辨精确模拟, 对颗粒运动的计算追踪会额外带来巨大的计算量, 因此, 在不断提高DNS在对单相流和颗粒两相流模拟的雷诺数的同时, 需要发展跨尺度的高精度数值算法, 包括进一步提高并行算法的效率, 以及行之有效的计及近壁湍流脉动效应和多颗粒影响的简化模型.

(6)关于HRNWT研究成果的应用. 自然界和工程界的流动大多是HRNWT和多相流, 而在实际应用中对这类流动还主要沿用低雷诺数情况的简化处理. 一个典型的例子就是对自然界风沙运动的预测. 需要建立适用于HRNWT情形的理论框架及预测方法, 这不仅针对风沙运动问题, 还包括HRNWT情形的飞行器和船舶舰艇的湍流噪声、壁面阻力预测及控制策略, 飞行器发动机燃烧及化工催化中多相混合效果及效率的提升方案等.

(责任编委: 许春晓)

致谢

国家自然科学基金资助项目 (11490553).

参考文献

胡隐樵, 高由禧, 王介, 季国良, 沈志宝, 程麟生, 陈家宜, 李守谦 . 1994.

黑河实验 (HEIFE)的一些研究成果

高原气象, 13:225-236

URL     [本文引用: 1]

本文回顾了黑河实验(HEIFE)的执行情况,总结了到目前为止的主要研究成果。得到了干旱地区陆面过程地表参数、地表能量平衡特征和湍流通量参数化的结果。同时得到了干旱地区陆面过程的物理机制同湿润地区有质的差异这一认识:同湿润地区不同,干旱地区地表蒸散和植被过程并不是地表能量平衡中的控制因子,因为感热并不输送到自由大气仅是大气边界层的强迫因子;而潜热输送给自由大气并且潜热的释放是大气环流的主要能源,但干旱地区很少或没有水汽和潜热输送到自由大气,所以干旱地区自由大气不能从地面得到热能;地表反射率大和向上长波辐射强致使净辐射亏损大;基于以上原因,干旱地区可能是一个水汽输送和热能的"汇"。这些结果使我们能更好地理解干旱地区的陆面过程。黑河实验的结果以系统的观测事实证实了绿洲的冷岛效应,临近绿洲沙漠的逆湿现象,即夏季白天常常是绿洲有感热向下输送,临近绿洲的沙漠有水汽向下输送。这些结果揭示了绿洲和沙漠环境相互作用的较完整图像。

( Hu Y Q, Gao Y X, Wang J M , et al. 1994.

Some achievements in scientific research during HEIFE

Plateau Meteorology, 13:225-236).

URL     [本文引用: 1]

本文回顾了黑河实验(HEIFE)的执行情况,总结了到目前为止的主要研究成果。得到了干旱地区陆面过程地表参数、地表能量平衡特征和湍流通量参数化的结果。同时得到了干旱地区陆面过程的物理机制同湿润地区有质的差异这一认识:同湿润地区不同,干旱地区地表蒸散和植被过程并不是地表能量平衡中的控制因子,因为感热并不输送到自由大气仅是大气边界层的强迫因子;而潜热输送给自由大气并且潜热的释放是大气环流的主要能源,但干旱地区很少或没有水汽和潜热输送到自由大气,所以干旱地区自由大气不能从地面得到热能;地表反射率大和向上长波辐射强致使净辐射亏损大;基于以上原因,干旱地区可能是一个水汽输送和热能的"汇"。这些结果使我们能更好地理解干旱地区的陆面过程。黑河实验的结果以系统的观测事实证实了绿洲的冷岛效应,临近绿洲沙漠的逆湿现象,即夏季白天常常是绿洲有感热向下输送,临近绿洲的沙漠有水汽向下输送。这些结果揭示了绿洲和沙漠环境相互作用的较完整图像。

纪勇 . 2019.

基于结构系综理论的大气表面层研究

博士学位论文. 北京大学湍流与复杂系统国家重点实验室, 北京大学.

[本文引用: 1]

李存标, 吴介之, 白夜 . 2009.

壁流动中的转捩

力学进展, 39:480-507

URL     [本文引用: 1]

对壁流动的不同实验结果做了对比,这些实验结果来自于流动显示、热膜测量以及PIV测量, 对比的同时,还总结了与此相应的理论方面的进展.这些进展是在对所选大约120篇文献中内容归纳提炼的基础上给出的.尽管实验中所使用的初始扰动条件不同,但所发现的流动结构几乎是完全一样的. 在对壁流动转捩的认识方面,认为下列所观察到的流动结构是最基本最重要的:在边界层和管流中被称为类孤子相干结构(SCS)的三维非线性涡包、$\Lambda$涡、二次涡环和涡环链.近期的实验中发现了这些结构形成和转捩的动力学过程,具体包括以下内容: (1) $\Lambda$ 涡和二次涡环间持续的相互作用过程.该过程决定了涡环链的产生方式, 总是从壁面区域周期性地形成,并进入到边界层的外部区域; (2)高频涡的生成,这是理解转捩和湍流边界层(以及其他流动)发展的关键问题之一.尽管已经提出了一些解释,但是二次涡环的实验发现将对此提供一个特别清晰的解释.(3)在所有湍流猝发中SCS所起的关键作用.这一点被看做是低雷诺数湍流边界层中湍流产生的关键机制.与猝发直接相关联的是低速条带. 基于SCS的动力学过程, 针对壁流动情况,可以比以前更清晰地解释低速条带的形成机制及其与流动结构的关系.在实验中所观察到的SCS和二次涡环,不仅能使我们重温壁面流动转捩中的经典故事, 同时还开辟一条新的途径,可以基于此建立壁面流动转捩可能具有的普适性的动力学过程.

( Li C B, Wu J Z, Bai Y . 2009.

Transition in wall-bouned flows

Advances in Mechanics, 39:480-507).

URL     [本文引用: 1]

对壁流动的不同实验结果做了对比,这些实验结果来自于流动显示、热膜测量以及PIV测量, 对比的同时,还总结了与此相应的理论方面的进展.这些进展是在对所选大约120篇文献中内容归纳提炼的基础上给出的.尽管实验中所使用的初始扰动条件不同,但所发现的流动结构几乎是完全一样的. 在对壁流动转捩的认识方面,认为下列所观察到的流动结构是最基本最重要的:在边界层和管流中被称为类孤子相干结构(SCS)的三维非线性涡包、$\Lambda$涡、二次涡环和涡环链.近期的实验中发现了这些结构形成和转捩的动力学过程,具体包括以下内容: (1) $\Lambda$ 涡和二次涡环间持续的相互作用过程.该过程决定了涡环链的产生方式, 总是从壁面区域周期性地形成,并进入到边界层的外部区域; (2)高频涡的生成,这是理解转捩和湍流边界层(以及其他流动)发展的关键问题之一.尽管已经提出了一些解释,但是二次涡环的实验发现将对此提供一个特别清晰的解释.(3)在所有湍流猝发中SCS所起的关键作用.这一点被看做是低雷诺数湍流边界层中湍流产生的关键机制.与猝发直接相关联的是低速条带. 基于SCS的动力学过程, 针对壁流动情况,可以比以前更清晰地解释低速条带的形成机制及其与流动结构的关系.在实验中所观察到的SCS和二次涡环,不仅能使我们重温壁面流动转捩中的经典故事, 同时还开辟一条新的途径,可以基于此建立壁面流动转捩可能具有的普适性的动力学过程.

林建忠, 朱泽飞, 沈利平 . 1998.

气固两相边界层中固粒与拟序结构相互作用的研究

上海力学, 4:310-317

URL     [本文引用: 1]

( Lin J Z, Zhu Z F, Shen L P . 1998.

Research on the interaction between the solid particles and the coherent structure in turbulent boundary layer

Shanghai Mechanics, 4:310-317).

URL     [本文引用: 1]

林建忠, 林江, 石兴 . 2002.

两相流中柱状固粒对流体湍动特性影响的研究

应用数学和力学, 23:483-488

URL     [本文引用: 1]

对含柱状固粒的两相流场,建立了包含柱状固粒对流场影响的流体脉动速度方程,在求解脉动速度方程的基础上,经平均得到流体的湍流强度和雷诺应力.将该方法用于槽流湍流场的求解,并与单相流实验结果进行了比较.计算中变化柱状固粒的参数,给出了固粒的体积分数、长径比、松驰时间对流场湍动特性的影响,说明粒子对流场的湍动特性起着抑制作用,其抑制的程度与粒子的体积分数、长径比成正比,与粒子的松弛时间成反比.

( Lin J Z, Lin J, Shi X . 2002.

Research on the effect of cylinder particles on the turbulent properties in particulate flows

Applied Mathematics and Mechanics, 23:483-488).

URL     [本文引用: 1]

对含柱状固粒的两相流场,建立了包含柱状固粒对流场影响的流体脉动速度方程,在求解脉动速度方程的基础上,经平均得到流体的湍流强度和雷诺应力.将该方法用于槽流湍流场的求解,并与单相流实验结果进行了比较.计算中变化柱状固粒的参数,给出了固粒的体积分数、长径比、松驰时间对流场湍动特性的影响,说明粒子对流场的湍动特性起着抑制作用,其抑制的程度与粒子的体积分数、长径比成正比,与粒子的松弛时间成反比.

许春晓 . 2015.

壁湍流相干结构和减阻控制机理

力学进展, 45:201504

DOI      URL     [本文引用: 1]

剪切湍流中相干结构的发现是上世纪湍流研究的重大进展之一,这些大尺度的相干运动在湍流的动力学过程中起重要作用,也为湍流的控制指出了新的方向.壁湍流高摩擦阻力的产生与近壁区流动结构密切相关,基于近壁区湍流动力学过程的减阻控制方案可以有效降低湍流的摩擦阻力,但是随着雷诺数的升高, 这些控制方案的有效性逐渐降低.近年来研究发现, 在高雷诺数情况下外区存在大尺度的相干运动,这种大尺度运动对近壁区湍流和壁面摩擦阻力的产生有重要影响,为高雷诺数湍流减阻控制策略的设计提出了新的挑战.该文将对壁湍流相干结构的研究历史加以简单的回顾,重点介绍近壁区相干结构及其控制机理、近年来高雷诺数外区大尺度运动的研究进展,在此基础上提出高雷诺数减阻控制研究的关键科学问题.

( Xu C X . 2015.

Coherent structures and drag-reduction mechanism in wall turbulence

Advances in Mechanics, 45:201504).

DOI      URL     [本文引用: 1]

剪切湍流中相干结构的发现是上世纪湍流研究的重大进展之一,这些大尺度的相干运动在湍流的动力学过程中起重要作用,也为湍流的控制指出了新的方向.壁湍流高摩擦阻力的产生与近壁区流动结构密切相关,基于近壁区湍流动力学过程的减阻控制方案可以有效降低湍流的摩擦阻力,但是随着雷诺数的升高, 这些控制方案的有效性逐渐降低.近年来研究发现, 在高雷诺数情况下外区存在大尺度的相干运动,这种大尺度运动对近壁区湍流和壁面摩擦阻力的产生有重要影响,为高雷诺数湍流减阻控制策略的设计提出了新的挑战.该文将对壁湍流相干结构的研究历史加以简单的回顾,重点介绍近壁区相干结构及其控制机理、近年来高雷诺数外区大尺度运动的研究进展,在此基础上提出高雷诺数减阻控制研究的关键科学问题.

徐祥德, 周明煜, 陈家宜, 卞林根, 张光智, 刘辉志, 李诗明, 张宏升, 赵冀俊, 索朗多吉, 王继志 . 2001.

青藏高原地-气过程动力、热力结构综合物理图象

中国科学: 地球科学, 31:428-440.

URL     [本文引用: 1]

郑晓静 . 2017.

大气表面湍流超大/大尺度结构

中国力学大会2017, 2017年8月13-16日, 北京.

[本文引用: 2]

顾海华, 郑晓静 . 2019.

大气表面层粉尘(PM10)输运的超大尺度结构

中国力学大会2019, 2019年 8月25-28日, 杭州.

[本文引用: 3]

王萍, 靳婷, 郑晓静 . 2019a.

壁湍流多相流大涡模拟壁函数改进

中国力学大会2019, 2019年 8月25-28日, 杭州.

[本文引用: 2]

王萍, 柳丽, 郑晓静 . 2019b.

基于沙尘暴期间大气表面层风场的粉尘输运数值模拟

中国力学大会2019, 2019年8月25-28日, 杭州.

[本文引用: 1]

周力行, 廖昌明, 陈涛 . 1994.

强旋气-粒两相湍流的统一二阶矩封闭模型

工程热物理学报. 15:327-330

URL     [本文引用: 2]

强旋气-粒两相流中两相的湍流均为强烈各向异性。本文由双流体模型出发,运用单相湍流的统计矩模拟方法建立了两相湍流的统一二阶矩封闭(USM)模型。这一模型是本文作者之一几年前提出的k-ε-kp模型的发展,在此前已初具雏型,本文中进一步加以完善。USM模型正确反映了两相湍流的各向异性,并且更深刻地揭示了颗粒与流体的湍流相互作用的机理,是今后研究强旋气-粒湍流两相流动值得探讨的重要方向之一。

( Zhou L, Liao C, Chen T . 1994.

A unified second-order moment two-phase turbulence model for strongly swirling gas-particle flows

Journal of Engineering Thermophysics, 15:327-330).

URL     [本文引用: 2]

强旋气-粒两相流中两相的湍流均为强烈各向异性。本文由双流体模型出发,运用单相湍流的统计矩模拟方法建立了两相湍流的统一二阶矩封闭(USM)模型。这一模型是本文作者之一几年前提出的k-ε-kp模型的发展,在此前已初具雏型,本文中进一步加以完善。USM模型正确反映了两相湍流的各向异性,并且更深刻地揭示了颗粒与流体的湍流相互作用的机理,是今后研究强旋气-粒湍流两相流动值得探讨的重要方向之一。

Adrian R J, Meinhart C D, Tomkins C D. 2000.

Vortex organization in the outer region of the turbulent boundary layer

Journal of Fluid Mechanics, 422:1-54.

DOI      URL     [本文引用: 3]

Ahn J, Lee J H, Lee J, Kang J H, Sung H J. 2015.

Direct numerical simulation of a 30R long turbulent pipe flow at $Re_\tau=3008$

Physics of Fluids, 27:065110.

DOI      URL     [本文引用: 1]

álamo J C, Jimenez J, Zandonade P, Moser R D. 2006.

Self-similar vortex clusters in the turbulent logarithmic region

Journal of Fluid Mechanics, 561:329-358.

DOI      URL     [本文引用: 1]

Alfredsson P H, Segalini A, örlü R. 2011.

A new scaling for the streamwise turbulence intensity in wall-bounded turbulent flows and what it tells us about the "outer" peak

Physics of Fluids, 23:041702.

DOI      URL     [本文引用: 1]

Anderson R S, Haff P K. 1988.

Simulation of Eolian Saltation

Science, 241:820-823.

DOI      URL     PMID      [本文引用: 1]

Saltation is important in the transport of sand-sized granular material by wind and in the ejection of dust from the bed both on Earth and on Mars. The evolution of the saltating population and all its characteristic profiles is calculated from inception by pure aerodynamic entrainment through to steady state. Results of numerical simulations of single-grain impacts into granular beds are condensed into analytic expressions for the number and speeds of grains rebounding or rejected (splashed) from the bed. A model is combined with (i) this numerical representation, (ii) an expression for the aerodynamic entrainment rate, and (iii) the modification of the wind velocity profile by saltating grains. Calculated steady state mass fluxes are within the range of mass fluxes measured in wind tunnel experiments; mass flux is nonlinearly dependent on the shear velocity. Aerodynamically entrained grains in the system are primarily seeding agents; at steady state, aerodynamic entrainment is rare. The time for the entire system to reach steady state is roughly 1 second, or several long-trajectory hop times.

Baas A C, Sherman D J. 2005.

Formation and behavior of aeolian streamers

Journal of Geophysical Research-Earth Surface, 110:F03011.

[本文引用: 1]

Baas A C. 2006.

Wavelet power spectra of Aeolian sand transport by boundary layer turbulence

Geophysical Research Letters, 33:L05403.

[本文引用: 1]

Bagchi P, Balachandar S. 2003.

Effect of turbulence on the drag and lift of a particle

Physics of Fluids, 15:3496-3513.

DOI      URL     [本文引用: 1]

Bagnold R A. 1941.

The physics of blown sand and desert dunes

Nature, 18:167-187.

[本文引用: 1]

Bailey S C C, Vallikivi M, Hultmark M, Smits A J. 2014.

Estimating the value of von Karman's constant in turbulent pipe flow

Journal of Fluid Mechanics, 749:79-98.

DOI      URL     [本文引用: 1]

Five separate data sets on the mean velocity distributions in the Princeton University/ONR Superpipe are used to establish the best estimate for the value of von Karman's constant for the flow in a fully developed, hydraulically smooth pipe. The profiles were taken using Pitot tubes, conventional hot wires and nanoscale thermal anemometry probes. The value of the constant was found to vary significantly due to measurement uncertainties in the mean velocity, friction velocity and the wall distance, and the number of data points included in the analysis. The best estimate for the von Karman constant in turbulent pipe flow is found to be 0.40 +/- 0.02. A more precise estimate will require improved instrumentation.

Balachandar S, Eaton J K. 2010.

Turbulent dispersed multiphase flow

Annual Review of Fluid Mechanics, 42:111-133.

DOI      URL     [本文引用: 1]

Balakumar B J, Adrian R J. 2007.

Large- and very-large-scale motions in channel and boundary-layer flows

Philos Trans A Math Phys Eng Sci, 365:665-681.

DOI      URL     PMID      [本文引用: 3]

Large-scale motions (LSMs; having wavelengths up to 2-3 pipe radii) and very-LSMs (having wavelengths more than 3 pipe radii) have been shown to carry more than half of the kinetic energy and Reynolds shear stress in a fully developed pipe flow. Studies using essentially the same methods of measurement and analysis have been extended to channel and zero-pressure-gradient boundary-layer flows to determine whether large structures appear in these canonical wall flows and how their properties compare with that of the pipe flow. The very large scales, especially those of the boundary layer, are shorter than the corresponding scales in the pipe flow, but otherwise share a common behaviour, suggesting that they arise from similar mechanism(s) aside from the modifying influences of the outer geometries. Spectra of the net force due to the Reynolds shear stress in the channel and boundary layer flows are similar to those in the pipe flow. They show that the very-large-scale and main turbulent motions act to decelerate the flow in the region above the maximum of the Reynolds shear stress.

Bandyopadhyay P R, Hussain P A K M F. 1984.

The coupling between scales in shear flows

Physics of Fluids, 27:2221-2228.

[本文引用: 1]

Barenblatt G I, Prostokishin V M. 1993.

Scaling laws for fully developed turbulent shear flows. Part 1. Basic hypotheses and analysis

Journal of Fluid Mechanics, 248:513-520.

DOI      URL     [本文引用: 1]

Bernardini M, Pirozzoli S, Orlandi P. 2014.

Velocity statistics in turbulent channel flow up to

Journal of Fluid Mechanics, 742:171-191.

DOI      URL     [本文引用: 3]

The high-Reynolds-number behaviour of the canonical incompressible turbulent channel flow is investigated through large-scale direct numerical simulation (DNS). A Reynolds number is achieved (Re-tau =h/delta(nu) approximate to 4000, where h is the channel half-height, and delta(nu) is the viscous length scale) at which theory predicts the onset of phenomena typical of the asymptotic Reynolds number regime, namely a sensible layer with logarithmic variation of the mean velocity profile, and Kolmogorov scaling of the velocity spectra. Although higher Reynolds numbers can be achieved in experiments, the main advantage of the present DNS study is access to the full three-dimensional flow field. Consistent with refined overlap arguments (Afzal & Yajnik, J. Fluid Mech. vol. 61, 1973, pp. 23-31; Jimenez & Moser, Phil. Trans. R. Soc. Lond. A, vol. 365, 2007, pp. 715-732), our results suggest that the mean velocity profile never achieves a truly logarithmic profile, and the logarithmic diagnostic function instead exhibits a linear variation in the outer layer whose slope decreases with the Reynolds number. The extrapolated value of the von Karman constant is k approximate to 0.41. A near logarithmic layer is observed in the spanwise velocity variance, as predicted by Townsend's attached eddy hypothesis, whereas the streamwise variance seems to exhibit a shoulder, perhaps being still affected by low-Reynolds-number effects. Comparison with previous DNS data at lower Reynolds number suggests enhancement of the imprinting effect of outer-layer eddies onto the near-wall region. This mechanisms is associated with excess turbulence kinetic energy production in the outer layer, and it reflects in flow visualizations and in the streamwise velocity spectra, which exhibit sharp peaks in the outer layer. Associated with the outer energy production site, we find evidence of a Kolmogorov-like inertial range, limited to the spanwise spectral density of u, whereas power laws with different exponents are found for the other spectra. Finally, arguments are given to explain the 'odd' scaling of the streamwise velocity variances, based on the analysis of the kinetic energy production term.

Blackwelder R F, Kovasznay L S G. 1972.

Time scales and correlations in a turbulent boundary layer

Physics of Fluids, 15:1545-1554.

DOI      URL     [本文引用: 1]

Brown G L, Thomas A S W. 1977.

Large structure in a turbulent boundary layer

Physics of Fluids, 20:S243-S252.

[本文引用: 2]

Cantwell B J. 1981.

Organized motion in turbulent-flow

Annual Review of Fluid Mechanics, 13:457-515.

DOI      URL     [本文引用: 1]

Caporaloni M, Tampieri F, Trombetti F. 1975.

Transfer of particles in nonisotropic air turbulence

Journal of the Atmospheric Sciences, 32:565-568.

DOI      URL     [本文引用: 1]

Carneiro M V, Rasmussen K R, Herrmann H J. 2015.

Bursts in discontinuous Aeolian saltation

Scientific Reports, 5:11109.

DOI      URL     PMID      [本文引用: 1]

Close to the onset of Aeolian particle transport through saltation we find in wind tunnel experiments a regime of discontinuous flux characterized by bursts of activity. Scaling laws are observed in the time delay between each burst and in the measurements of the wind fluctuations at the fluid threshold Shields number thetac. The time delay between each burst decreases on average with the increase of the Shields number until sand flux becomes continuous. A numerical model for saltation including the wind-entrainment from the turbulent fluctuations can reproduce these observations and gives insight about their origin. We present here also for the first time measurements showing that with feeding it becomes possible to sustain discontinuous flux even below the fluid threshold.

Carper M A, Porte-Agel F. 2004.

The role of coherent structures in subfilter-scale dissipation of turbulence measured in the atmospheric surface layer

Journal of Turbulence, 5:040.

[本文引用: 1]

Castillo L, George W K. 2001.

Similarity analysis for turbulent boundary layer with pressure gradient: Outer flow

AIAA Journal, 39:41-47.

DOI      URL     [本文引用: 1]

Chauhan K A. 2007.

Study of canonical wall-bounded turbulent flows. [PhD Thesis]

Illinois: Illinois Institute of Technology.

DOI      URL     PMID      [本文引用: 1]

A major challenge for translating cell-based therapies is understanding the dynamics of cells and cell populations in complex in vivo environments. Intravital microscopy has shown great promise for directly visualizing cell behavior in vivo. However, current methods are limited to relatively short imaging times (hours), by ways to track cell and cell population dynamics over extended time-lapse periods (days to weeks to months), and by relatively few imaging contrast mechanisms that persist over extended investigations. We present technology to visualize and quantify complex, multifaceted dynamic changes in natural deformable skin over long time periods using novel multimodal imaging and a non-rigid image registration method. These are demonstrated in green fluorescent protein (GFP) bone marrow (BM) transplanted mice to study dynamic skin regeneration. This technology provides a novel perspective for studying dynamic biological processes and will enable future studies of stem, immune, and tumor cell biology in vivo.

Chauhan K A, Nagib H, Monkewitz P. 2007.

On the composite logarithmic profile in zero pressure gradient turbulent boundary layers

//45th AIAA Aerospace Sciences Meeting and Exhibit.

[本文引用: 1]

Chauhan K A, Hutchins N, Monty J, Marusic I. 2013.

Structure inclination angles in the convective atmospheric surface layer

Boundary-Layer Meteorology, 147:41-50.

DOI      URL     [本文引用: 1]

Two-point correlations of the fluctuating streamwise velocity are examined in the atmospheric surface layer over the salt flats of Utah's western desert, and corresponding structure inclination angles are obtained for neutral, stable and unstable conditions. The neutral surface-layer results supplement evidence for the invariance of the inclination angle given in Marusic and Heuer (Phys Rev Lett 99:114504, 2007). In an extension of those results it is found that the inclination angle changes drastically under different stability conditions in the surface layer, varying systematically with the Monin-Obukhov stability parameter in the unstable regime. The variation is parametrized and subsequently can be used to improve existing near-wall models in the large-eddy simulation of the atmospheric surface layer.

Chhabra R P, Agarwal L, Sinha N Kl. 1999.

Drag on non-spherical particles: An evaluation of available methods

Powder Technology, 101:288-295.

[本文引用: 1]

Choi H, Moin P. 2012.

Grid-point requirement for large eddy simulation: Chapman's estimation revisited

Physics of Fluids, 24:011702.

[本文引用: 1]

Cooper D I, Leclerc M Y, Archuleta J, Coulter R, Eichinger W E, Kao C Y J, Nappo C J. 2006.

Mass exchange in the stable boundary layer by coherent structures

Agricultural & Forest Meteorology, 136:114-131.

[本文引用: 1]

Corrsin S, Kistler A L. 1954.

Free-stream boundaries of turbulent flows

N.A.C.A. Tech. Note no. 3133.(See also N.A.C.A. Tech. Rep. no.1244, 1955).

[本文引用: 1]

De Graaff D B, Eaton J K. 2000.

Reynolds-number scaling of the flat-plate turbulent boundary layer

Journal of Fluid Mechanics, 422:319-346.

[本文引用: 3]

Deck S, Renard N, Laraufie R, Weiss P-é. 2014.

Large-scale contribution to mean wall shear stress in high-Reynolds-number flat-plate boundary layers up to $Re_\theta=13650$

Journal of Fluid Mechanics, 743:202-248.

[本文引用: 1]

Deng B, Huang W, Xu C. 2016.

Origin of effectiveness degradation in active drag reduction control of turbulent channel flow at $Re_\tau=1000$

Journal of Turbulence, 17:758-786.

[本文引用: 2]

Deng S, Pan C, Wang J J, He G S. 2018.

On the spatial organization of hairpin packets in a turbulent boundary layer at low-to-moderate Reynolds number

Journal of Fluid Mechanics, 844:635-668.

[本文引用: 1]

Dennis D J C. 2015.

Coherent structures in wall-bounded turbulence

Anais Da Academia Brasileira De Ciências, 87:513-537.

[本文引用: 1]

Dixit S A, Ramesh O N. 2018.

Streamwise self-similarity and log scaling in turbulent boundary layers

Journal of Fluid Mechanics, 851:R1.

[本文引用: 1]

Dritselis C D, Vlachos N S. 2008.

Numerical study of educed coherent structures in the near-wall region of a particle-laden channel flow

Physics of Fluids, 20:055103.

[本文引用: 2]

Drobinski P, Carlotti P, Newsom R K, Banta R M, Foster R C, Redelsperger J L. 2004.

The structure of the near neutral atmospheric surface layer

Journal of the Atmospheric Sciences, 61:699-714.

[本文引用: 1]

Dupont S, Bergametti G, Marticorena B, Simoens S. 2013.

Modeling saltation intermittency

Journal of Geophysical Research-Atmospheres, 118:7109-7128.

[本文引用: 1]

Elghobashi S. 1994.

On predicting particle-laden turbulent flows

Applied Scientific Research, 52:309-329.

[本文引用: 1]

Elgobashi S, Balachandar S, Prosperetti A. 2006.

An updated classification map of particle-laden turbulent flows

IUTAM Symposium on Computational Approaches to Multiphase Flow, 81:3-10.

[本文引用: 1]

Falco R E. 1977.

Coherent motions in the outer region of turbulent boundary layers

The Physics of Fluids, 20:S124-S132.

Fernholz H H, Krause E, Nockemann M, Schober M. 1995.

Comparative measurements in the canonical boundary-layer at $Re _{\delta 2}\leq 6\times 10^{4}$ on the wall of the German-Dutch Wind Tunnel

Physics of Fluids, 7:1275-1281.

[本文引用: 1]

Flores O, Jiménez J, álamo J C D. 2007.

Vorticity organization in the outer layer of turbulent channels with disturbed walls

Journal of Fluid Mechanics, 591:145-154.

[本文引用: 1]

Frisch U. 1995. Turbulence. Cambridge: Cambridge University Press.

[本文引用: 1]

Furuichi N, Terao Y, Wada Y, Tsuji Y. 2018.

Further experiments for mean velocity profile of pipe flow at high Reynolds number

Physics of Fluids, 30:7.

[本文引用: 3]

Ganapathisubramani B, Longmire E K, Marusic I, Pothos S. 2005.

Dual-plane PIV technique to determine the complete velocity gradient tensor in a turbulent boundary layer

Experiments In Fluids, 39:222-231.

DOI      URL     [本文引用: 1]

Geiss S, Dreizler A, Stojanovic Z. 2004.

Investigation of turbulence modification in a non-reactive two phase flow

Experiments in Fluids, 36:344-354.

DOI      URL     [本文引用: 1]

In a two-phase flow the influence of a dispersed phase on the turbulence properties of a continuous phase, known as turbulence modification, is investigated. An experimental approach is discussed that is suitable for studying the decay of grid-generated turbulence in a vertically orientated wind tunnel with a cross-section large enough to avoid influences from walls. Phase Doppler anemometry is used to characterize both single and two-phase flow by measuring mean axial and radial velocity components, velocity fluctuations, turbulent kinetic energy, and integral time scales. By direct comparison of results from single- and two-phase flows, the feedback of the dispersed phase on the continuous phase can be isolated. The data is used to deduce a source term for particle-induced turbulence production appropriate for a numerical simulation of the flow, based on the Reynolds-averaged Navier-Stokes equations. Although of special importance for a detailed understanding of turbulent two-phase combustion, additional complexity introduced by evaporation and chemical reactions is avoided by using glass beads as dispersed phase.

George W K. 1995.

Some new ideas for similarity of turbulent shear flows

Turbulence, Heat and Mass Transfer, 1:13-24.

[本文引用: 1]

George W K. 2007.

Is there a universal log law for turbulent wall-bounded flows

Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 365:789-806.

DOI      URL     [本文引用: 1]

Greeley R, Blumberg D G, Williams S H. 1996.

Field measurements of the flux and speed of wind-blown sand

Sedimentology, 43:41-52.

DOI      URL     [本文引用: 1]

Gu H, Wang G, Zhu W, Zheng X J. 2019.

Gusty wind disturbances and large-scale turbulent structures in the neutral atmospheric surface layer

Science China Physics, Mechanics & Astronomy, 62:114711.

DOI      URL     [本文引用: 3]

Guala M, Hommema S E, Adrian R J. 2006.

Large-scale and very-large-scale motions in turbulent pipe flow

Journal of Fluid Mechanics, 554:521-542.

DOI      URL     [本文引用: 2]

Guala M, Metzger M, Mckeon B J. 2011.

Interactions within the turbulent boundary layer at high Reynolds number

Journal of Fluid Mechanics, 666:573-604.

DOI      URL     [本文引用: 5]

Simultaneous streamwise velocity measurements across the vertical direction obtained in the atmospheric surface layer (Re-tau similar or equal to 5 x 105) under near thermally neutral conditions are used to outline and quantify interactions between the scales of turbulence, from the very-large-scale motions to the dissipative scales. Results from conditioned spectra, joint probability density functions and conditional averages show that the signature of very-large-scale oscillations can be found across the whole wall region and that these scales interact with the near-wall turbulence from the energy-containing eddies to the dissipative scales, most strongly in a layer close to the wall, z(+) less than or similar to 10(3). The scale separation achievable in the atmospheric surface layer appears to be a key difference from the low-Reynolds-number picture, in which structures attached to the wall are known to extend through the full wall-normal extent of the boundary layer. A phenomenological picture of very-large-scale motions coexisting and interacting with structures from the hairpin paradigm is provided here for the high-Reynolds-number case. In particular, it is inferred that the hairpin-packet conceptual model may not be exhaustively representative of the whole wall region, but only of a near-wall layer of z(+) = O(10(3)), where scale interactions are mostly confined.

Hadinoto K, Jones E N, Yurteri C, Curtis J S. 2005.

Reynolds number dependence of gas-phase turbulence in gas-particle flows

International Journal of Multiphase Flow, 31:416-434.

DOI      URL     [本文引用: 1]

Hagen G. 1839.

Ueber die Bewegung des Wassers in engen cylindrischen Röhren

Annalen der Physik, 122:423-442.

DOI      URL     [本文引用: 1]

Han G, Liu L, Bo T, Zheng X. 2019 a.

A predictive model for the streamwise velocity in the near-neutral atmospheric surface layer

Journal of Geophysical Research-Atmospheres, 124:238-251.

DOI      URL     [本文引用: 1]

Han G, Wang G, Zheng X. 2019 b.

The applicability of Taylor's hypothesis for estimating the mean streamwise length scales of large-scale structures in the near-neutral atmospheric surface layer

Boundary-Layer Meteorol, 172:215-237.

DOI      URL     [本文引用: 2]

He G, Jin G, Yang Y. 2017.

Space-time correlations and dynamic coupling in turbulent flows

Annual Review of Fluid Mechanics, 49:51-70.

DOI      URL     [本文引用: 1]

He G W, Zhang J B. 2006.

Elliptic model for space-time correlations in turbulent shear flows

Phys Rev E Stat Nonlin Soft Matter Phys, 73:055303.

URL     PMID      [本文引用: 1]

Head M R, Bandyopadhyay P. 1981.

New aspects of turbulent boundary-layer structure

Journal of Fluid Mechanics, 107:297-338.

DOI      URL     [本文引用: 1]

Heisel M, Dasari T, Liu Y, Hong J R, Coletti F, Guala M. 2018.

The spatial structure of the logarithmic region in very-high-Reynolds-number rough wall turbulent boundary layers

Journal of Fluid Mechanics, 857:704-747.

DOI      URL     [本文引用: 1]

Helland E, Occelli R, Tadrist L. 2005.

Numerical study of cluster and particle rebound effects in a circulating fluidised bed

Chemical Engineering Science, 60:27-40.

DOI      URL     [本文引用: 1]

Abstract

Gas–particle flows in a vertical two-dimensional configuration appropriate for circulating fluidised bed applications were investigated numerically. In the computational study presented herein the motion of particles was calculated based on a Lagrangian approach and particles were assumed to interact through binary, instantaneous, non-frontal, inelastic collisions including friction. The model for the interstitial gas phase is based on the Navier–Stokes equations for two-phase flows. The numerical study of cluster structures has been validated with experimental results from literature in a previous investigation. Numerical experiments were performed in order to study the effects of different cluster and particle rebound characteristics on the gas–particle flow behaviour.

Firstly, we investigated the hard sphere collision model and its effect on gas–particle flow behaviour. The coefficient of restitution in an impact depends not only on the material properties of the colliding objects, but also on their relative impact velocity. We compared the effect of a variable restitution coefficient, dependent on the relative impact velocity, with the classical approach, which supposes the coefficient of restitution to be constant and independent of the relative impact velocity.

Secondly, we studied the effects of different cluster properties on the gas–particle flow behaviour. Opposing clustering effects have been observed for different particle concentrations: within a range of low concentrations, groups of particles fall faster than individual particles due to cluster formation, and within a well-defined higher concentration range, return flow predominates and hindered settling characterises the suspension. We propose herein a drag law, which takes into account both opposing effects and have compared the resulting flow behaviour with that predicted by a classical drag law, which takes into account only the hindered settling effect.

Högström U, Hunt J C R, Smedman A S. 2002.

Theory and measurements for turbulence spectra and variances in the atmospheric neutral surface layer

Boundary-Layer Meteorology, 103:101-124.

DOI      URL     [本文引用: 1]

Predictions from a new theory for high Reynolds number turbulent boundary layers during near-neutral conditions are shown to agree well with measurements of atmospheric surface-layer variances and spectra. The theory suggests surface-layer turbulence is determined by detached eddies that largely originate in the shearing motion immediately above the surface layer; as they descend into this layer, they are strongly distorted by the local shear and impinge onto the surface. Because the origin of these eddies is non-local, they are similar to those described in previous studies as `inactive' turbulence. However, they are, in fact, dynamically highly active, supplying the major mechanism for the momentum transport, including upward bursting on the time scale of the larger eddies. The vertical velocity results show that the variance and the low frequency parts of spectra increase with height in the surface layer, while in the self similar (k1-1) range the streamwise low frequency components are approximately constant with height. These large-scale longitudinal eddies extend to a length Lambdas, which is equal to the boundary-layer height near the surface andincreases linearly to a maximum of about three times the boundary-layer height at roughly 15 m and decreases in the upper parts of the surface layer. This lower part of the surface layer, the eddy surface layer, is the region in which the eddies impinging from layers above are strongly distorted. This new result for the atmospheric boundary layer has practical application for calculating fluctuating wind loads on structures and lateral dispersion of pollution from local sources.

Homann H, Jérémie B, Rainer G. 2013.

Effect of turbulent fluctuations on the drag and lift forces on a towed sphere and its boundary layer

Journal of Fluid Mechanics, 721:155-179.

DOI      URL     [本文引用: 1]

Hommema S E, Adrian R J. 2003.

Packet structure of surface eddies in the atmospheric boundary layer

Boundary-Layer Meteorology, 106:147-170.

DOI      URL     [本文引用: 1]

A smoke visualization experiment has beenperformed in the first 3,m ofneutral and unstable atmospheric boundary layersat very large Reynolds number(ReTHgr > 106). Under neutral atmosphericconditions mean wind profiles agreewell with those in the canonical flatplate zero-pressure-gradient turbulentboundary layer. The experiment was designedto minimize the temperaturedifference between the passive marker (smoke)and the air to ensure that anyobserved structures were due to vortical, ratherthan buoyant, motions. Imagesacquired in the streamwise–wall-normal planeusing a planar laser light-sheetare strikingly similar to those observed inlaboratory experiments at low to moderate Reynolds numbers. They reveal large-scaleramp-like structures withdownstream inclination of 3°–35°.This inclination isinterpreted as the hairpin packet growthangle following the hairpin vortexpacket model ofAdrian, Meinhart, and Tomkins.The distribution of this characteristicangle agrees with the results of experiments at far lower Reynolds numbers,suggesting a similarity in structures among low, moderate, and high Reynoldsnumber boundary layers at vastly different scales. These results indicate thatthe hairpin vortex packet model extends over a large range of scales. Theeffect of vertical heat transport in an unstable atmosphere on wall structuresis investigated in terms of the hairpin vortex packet model.

Horiguchi M, Hayashi T, Adachi A, Onogi S. 2012.

Large-scale turbulence structures and their contributions to the momentum flux and turbulence in the near-neutral atmospheric boundary layer observed from a 213 m tall meteorological tower

Boundary-Layer Meteorology, 144:179-198.

DOI      URL     [本文引用: 1]

Large-scale turbulence structures in the near-neutral atmospheric boundary layer (ABL) are investigated on the basis of observations made from the 213-m tall meteorological tower at Tsukuba, Japan. Vertical profiles of wind speed and turbulent fluxes in the ABL were obtained with sonic anemometer-thermometers at six levels of the tower. From the archived data, 31 near-neutral cases are selected for the analysis of turbulence structures. For the typical case, event detection by the integral wavelet transform with a large time scale (180 s) from the streamwise velocity component (u) at the highest level (200 m) reveals a descending high-speed structure with a time scale of approximately 100 s (a spatial scale of 1 km at the 200-m height). By applying the wavelet transform to the u velocity component at each level, the intermittent appearance of large-scale high-speed structures extending also in the vertical is detected. These structures usually make a large contribution to the downward momentum transfer and induce the enhancement of turbulent kinetic energy. This behaviour is like that of "active" turbulent motions. From the analysis of the two-point space-time correlation of wavelet coefficients for the u velocity component, the vertical extent and the downward influence of large-scale structures are examined. Large fluctuations in the large-scale range (wavelet variance at the selected time scale) at the 200-m level tend to induce the large correlation between the higher and lower levels.

Hoyas S, Jiménez J. 2006.

Scaling of the velocity fluctuations in turbulent channels up to $Re_\tau=2003$

Physics of Fluids, 18:011702.

DOI      URL     [本文引用: 3]

Hoyas S, Oberlack M, Kraheberger S, Alcantara-Avila F. 2018.

Turbulent channel flow at $Re_\tau=10000$

// APS Division of Fluid Dynamics Meeting. Nov. 18-20, Atlanta, USA.

[本文引用: 1]

Hultmark M, Vallikivi M, Bailey S C C, Smits A J. 2012.

Turbulent pipe flow at extreme Reynolds numbers

Physical Review Letters, 108:094501.

DOI      URL     PMID      [本文引用: 1]

Both the inherent intractability and complex beauty of turbulence reside in its large range of physical and temporal scales. This range of scales is captured by the Reynolds number, which in nature and in many engineering applications can be as large as 10(5)-10(6). Here, we report turbulence measurements over an unprecedented range of Reynolds numbers using a unique combination of a high-pressure air facility and a new nanoscale anemometry probe. The results reveal previously unknown universal scaling behavior for the turbulent velocity fluctuations, which is remarkably similar to the well-known scaling behavior of the mean velocity distribution.

Hultmark M, Vallikivi M, Bailey S C C, Smits A J. 2013.

Logarithmic scaling of turbulence in smooth- and rough-wall pipe flow

Journal of Fluid Mechanics, 728:376-395.

DOI      URL     [本文引用: 1]

Measurements of the streamwise component of the turbulent fluctuations in fully developed smooth and rough pipe flow are presented over an unprecedented Reynolds number range. For Reynolds numbers Re-tau > 20 000, the streamwise Reynolds stress closely follows the scaling of the mean velocity profile, independent of the roughness, and over the same spatial extent. This observation extends the findings of a logarithmic law in the turbulence fluctuations as reported by Hultmark, Vallikivi & Smits (Phys. Rev. Lett., vol. 108, 2012) to include rough flows. The onset of the logarithmic region is found at a location where the wall distance is equal to similar to 100 times the Kolmogorov length scale, which then marks sufficient scale separation for inertial scaling. Furthermore, in the logarithmic region the square root of the fourth-order moment also displays logarithmic behaviour, in accordance with the observation that the underlying probability density function is close to Gaussian in this region.

Hunt J C R, Carlotti P. 2001.

Statistical structure at the wall of the high Reynolds number turbulent boundary layer

Flow Turbulence & Combustion, 66:453-475.

[本文引用: 1]

Hunt J C R, Morrison J F. 2000.

Eddy structure in turbulent boundary layers

European Journal of Mechanics, 19:673-694.

DOI      URL     [本文引用: 1]

Hutchins N, Hambleton W T, Marusic I. 2005.

Inclined cross-stream stereo particle image velocimetry measurements in turbulent boundary layers

Journal of Fluid Mechanics, 541:21-54.

DOI      URL     [本文引用: 1]

Hutchins N, Marusic I. 2007 a.

Evidence of very long meandering features in the logarithmic region of turbulent boundary layers

Journal of Fluid Mechanics, 579:1-28.

DOI      URL     [本文引用: 7]

Hutchins N, Marusic I. 2007 b.

Large-scale influences in near-wall turbulence

Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 365:647-664.

DOI      URL     [本文引用: 2]

Hutchins N, Marusic I, Monty J, Klewicki J. 2012.

Towards Reconciling the Large-Scale Structure of Turbulent Boundary Layers in the Atmosphere and Laboratory

Boundary-Layer Meteorology, 145:273-306.

DOI      URL     [本文引用: 5]

A collaborative experimental effort employing the minimally perturbed atmospheric surface-layer flow over the salt playa of western Utah has enabled us to map coherence in turbulent boundary layers at very high Reynolds numbers, . It is found that the large-scale coherence noted in the logarithmic region of laboratory-scale boundary layers are also present in the very high Reynolds number atmospheric surface layer (ASL). In the ASL these features tend to scale on outer variables (approaching the kilometre scale in the streamwise direction for the present study). The mean statistics and two-point correlation map show that the surface layer under neutrally buoyant conditions behaves similarly to the canonical boundary layer. Linear stochastic estimation of the three-dimensional correlation map indicates that the low momentum fluid in the streamwise direction is accompanied by counter-rotating roll modes across the span of the flow. Instantaneous flow fields confirm the inferences made from the linear stochastic estimations. It is further shown that vortical structures aligned in the streamwise direction are present in the surface layer, and bear attributes that resemble the hairpin vortex features found in laboratory flows. Ramp-like high shear zones that contribute significantly to the Reynolds shear-stress are also present in the ASL in a form nearly identical to that found in laboratory flows. Overall, the present findings serve to draw useful connections between the vast number of observations made in the laboratory and in the atmosphere.

Hutchins N, Monty J P, Ganapathisubramani B, Ng H C H, Marusic I. 2011.

Three-dimensional conditional structure of a high-Reynolds-number turbulent boundary layer

Journal of Fluid Mechanics, 673:255-285.

DOI      URL     [本文引用: 1]

An array of surface hot-film shear-stress sensors together with a traversing hot-wire probe is used to identify the conditional structure associated with a large-scale skin-friction event in a high-Reynolds-number turbulent boundary layer. It is found that the large-scale skin-friction events convect at a velocity that is much faster than the local mean in the near-wall region (the convection velocity for large-scale skin-friction fluctuations is found to be close to the local mean at the midpoint of the logarithmic region). Instantaneous shear-stress data indicate the presence of large-scale structures at the wall that are comparable in scale and arrangement to the superstructure events that have been previously observed to populate the logarithmic regions of turbulent boundary layers. Conditional averages of streamwise velocity computed based on a low skin-friction footprint at the wall offer a wider three-dimensional view of the average superstructure event. These events consist of highly elongated forward-leaning low-speed structures, flanked on either side by high-speed events of similar general form. An analysis of small-scale energy associated with these large-scale events reveals that the small-scale velocity fluctuations are attenuated near the wall and upstream of a low skin-friction event, while downstream and above the low skin-friction event, the fluctuations are significantly amplified. In general, it is observed that the attenuation and amplification of the small-scale energy seems to approximately align with large-scale regions of streamwise acceleration and deceleration, respectively. Further conditional averaging based on streamwise skin-friction gradients confirms this observation. A conditioning scheme to detect the presence of meandering large-scale structures is also proposed. The large-scale meandering events are shown to be a possible source of the strong streamwise velocity gradients, and as such play a significant role in modulating the small-scale motions.

Inoue M, Mathis R, Marusic I, Pullin D I. 2012.

Inner-layer intensities for the flat-plate turbulent boundary layer combining a predictive wall-model with large-eddy simulations

Physics of Fluids, 24:075102.

DOI      URL     [本文引用: 1]

Jackson D W T. 1996.

Potential inertial effects in aeolian sand transport: Preliminary results

Sedimentary Geology, 106:193-201.

DOI      URL     [本文引用: 1]

Jackson N L, Sherman D J, Hesp P A, Klein A H F, Ballasteros Jr F, Nordstrom K F. 2006.

Small-scale spatial variations in aeolian sediment transport on a fine-sand beach

Journal of Coastal Research, 39:379-383.

Jacob C, Anderson W. 2016.

Conditionally averaged large-scale motions in the neutral atmospheric boundary layer: Insights for aeolian processes

Boundary-Layer Meteorology, 162:1-21.

DOI      URL     [本文引用: 1]

Kaftori D, Hetsroni G, Banerjee S. 1995.

Particle behavior in the turbulent boundary layer. II. Velocity and distribution profiles

Physics of Fluids, 7:1107-1121.

DOI      URL     [本文引用: 2]

Kaimal J C, Wyngaard J C. 1989.

The Kansas and Minnesota experiments

Boundary-Layer Meteorology, 50:31-47.

DOI      URL     [本文引用: 1]

Kaye B H, Boardman R P. 1962.

Cluster formation in dilute suspensions

// Proceedings of Symposium on the Interaction between Fluids and Particles, Institution of Chemical Engineers, London.

[本文引用: 1]

Kim J, Balachandar S. 2012.

Mean and fluctuating components of drag and lift forces on an isolated finite-sized particle in turbulence

Theoretical and Computational Fluid Dynamics, 26:185-204.

DOI      URL     [本文引用: 1]

We perform fully resolved direct numerical simulations of an isolated particle subjected to free-stream turbulence in order to investigate the effect of turbulence on the drag and lift forces at the level of a single particle, following Bagchi and Balachandar's work (Bagchi and Balachandar in Phys Fluids 15:3496-3513, 2003). The particle Reynolds numbers based on the mean relative particle velocity and the particle diameter are Re = 100, 250 and 350, which covers three different regimes of wake evolution in a uniform flow: steady axisymmetric wake, steady planar symmetric wake, and unsteady planar symmetric vortex shedding. At each particle Reynolds number, the turbulent intensity is 5-10% of the mean relative particle velocity, and the corresponding diameter of the particle is comparable to or larger than the Kolmogorov scale. The simulation results show that standard drag values determined from uniform flow simulations can accurately predict the drag force if the turbulence intensity is sufficiently weak (5% or less compared to the mean relative velocity). However, it is shown that for finite-sized particles, flow non-uniformity, which is usually neglected in the case of the small particles, can play an important role in determining the forces as the relative turbulence intensity becomes large. The influence of flow non-uniformity on drag force could be qualitatively similar to the Faxen correction. In addition, finite-sized particles at sufficient Reynolds number are inherently subjected to stochastic forces arising from their self-induced vortex shedding in addition to lift force arising from the local ambient flow properties (vorticity and strain rate). The effect of rotational and strain rate of the ambient turbulence seen by the particle on the lift force is explored based on the conditional averaging using the generalized representation of the quasi-steady force proposed by Bagchi and Balachandar (J Fluid Mech 481:105-148, 2003). From the present study, it is shown that at Re = 100, the lift force is mainly influenced by the surrounding turbulence, but at Re = 250 and 350, the lift force is affected by the wake structure as well as the surrounding turbulence. Thus, for a finite-sized particle of sufficient Reynolds number supporting self-induced vortex shedding, the lift force will not be completely correlated with the ambient flow. Therefore, it appears that in order to reliably predict the motion of a finite-sized particle in turbulence, it is important to incorporate both a deterministic component and a stochastic component in the force model. The best deterministic contribution is given by the conditional average. The influence of ambient turbulence at the scale of the particle, which are not accounted for in the deterministic contribution, can be considered in stochastic manner. In the modeling of lift force, additional stochastic contribution arising from self-induced vortex shedding must also be included.

Kim J, Moin P, Moser R. 1987.

Turbulence statistics in fully developed channel flow at low Reynolds number

Journal of Fluid Mechanics, 177:133-166.

DOI      URL     [本文引用: 2]

Kim K C, Adrian R J. 1999.

Very large-scale motion in the outer layer

Physics of Fluids, 11:417-422.

DOI      URL     [本文引用: 5]

Klewicki J C. 2010.

Reynolds number dependence, scaling, and dynamics of turbulent boundary layers

Journal of Fluids Engineering-Transactions of the Asme, 132:094001.

DOI      URL     [本文引用: 2]

Klewicki J C, Fife P, Wei T. 2009.

On the logarithmic mean profile

Journal of Fluid Mechanics, 638:73-93.

DOI      URL     [本文引用: 1]

Kline S J, Reynolds W C, Schraub F A, Runstadler P W. 1967.

The structure of turbulent boundary layers

Journal of Fluid Mechanics, 30:741-773.

DOI      URL     [本文引用: 2]

Kovasznay L S G, Kibens V, Blackwelder R F. 1970.

Large-scale motion in the intermittent region of a turbulent boundary layer

Journal of Fluid Mechanics, 41:283-325.

DOI      URL     [本文引用: 2]

Krogstad P Å, Antonia R A. 1994.

Structure of turbulent boundary layers on smooth and rough walls

Journal of Fluid Mechanics, 277:1-21.

DOI      URL     [本文引用: 1]

Kulick J D, Fessler J R, Eaton J K. 1994.

Particle response and turbulence modification in fully developed channel flow

Journal of Fluid Mechanics, 277:109-134.

DOI      URL     [本文引用: 1]

Kunkel G J, Marusic I. 2006.

Study of the near-wall-turbulent region of the high-Reynolds-number boundary layer using an atmospheric flow

Journal of Fluid Mechanics, 548:375-402.

DOI      URL     [本文引用: 2]

Kussin J, Sommerfeld M. 2002.

Experimental studies on particle behaviour and turbulence modification in horizontal channel flow with different wall roughness

Experiments in Fluids, 33:143-159.

DOI      URL     [本文引用: 1]

Lee J, Sung H. 2011.

Very-large-scale motions in a turbulent boundary layer

Journal of Fluid Mechanics, 673:80-120.

DOI      URL     [本文引用: 1]

Direct numerical simulation of a turbulent boundary layer was performed to investigate the spatially coherent structures associated with very-large-scale motions (VLSMs). The Reynolds number was varied in the range Re(theta) = 570-2560. The main simulation was conducted by using a computational box greater than 50 delta(o) in the streamwise domain, where delta(o) is the boundary layer thickness at the inlet, and inflow data was obtained from a separate inflow simulation based on Lund's method. Inspection of the three-dimensional instantaneous fields showed that groups of hairpin vortices are coherently arranged in the streamwise direction and that these groups create significantly elongated low-and high-momentum regions with large amounts of Reynolds shear stress. Adjacent packet-type structures combine to form the VLSMs; this formation process is attributed to continuous stretching of the hairpins coupled with lifting-up and backward curling of the vortices. The growth of the spanwise scale of the hairpin packets occurs continuously, so it increases rapidly to double that of the original width of the packets. We employed the modified feature extraction algorithm developed by Ganapathisubramani, Longmire & Marusic (J. Fluid Mech., vol. 478, 2003, p. 35) to identify the properties of the VLSMs of hairpin vortices. In the log layer, patches with the length greater than 3 delta-4 delta account for more than 40% of all the patches and these VLSMs contribute approximately 45% of the total Reynolds shear stress included in all the patches. The VLSMs have a statistical streamwise coherence of the order of similar to 6 delta; the spatial organization and coherence decrease away from the wall, but the spanwise width increases monotonically with the wall-normal distance. Finally, the application of linear stochastic estimation demonstrated the presence of packet organization in the form of a train of packets in the log layer.

Lee M, Moser R D. 2015.

Direct numerical simulation of turbulent channel flow up to $Re_\tau\sim 5200$

Journal of Fluid Mechanics, 774:395-415.

DOI      URL     [本文引用: 4]

Lee M, Ulerich R, Malaya N, Moser R D. 2014.

Experiences from leadership computing in simulations of turbulent fluid flows

Computing in Science & Engineering, 16:24-31.

[本文引用: 1]

Lee J, Lee C. 2015.

Modification of particle-laden near-wall turbulence: Effect of Stokes number

Physics of Fluids, 27:023303.

DOI      URL     [本文引用: 1]

Li D, Wei A, Luo K. 2016.

Direct numerical simulation of a particle-laden flow in a flat plate boundary layer

International Journal of Multiphase Flow, 79:124-143.

DOI      URL     [本文引用: 2]

Li Y, Mclaughlin J B, Kontomaris K. 2001.

Numerical simulation of particle-laden turbulent channel flow

Physics of Fluids, 13:2957-2967.

DOI      URL     [本文引用: 1]

Li J, Wang H, Liu Z, Chen S, Zheng C. 2012.

An experimental study on turbulence modification in the near-wall boundary layer of a dilute gas-particle channel flow

Experiments in Fluids, 53:1385-1403.

DOI      URL     [本文引用: 2]

Turbulence modifications of a dilute gas-particle flow are experimentally investigated in the lower boundary layer of a horizontal channel by means of a simultaneous two-phase PIV measurement technique. The measurements are conducted in the near-wall region with y(+) < 250 at Re-tau (based on the wall friction velocity u(tau) and half channel height h) = 430. High spatial resolution and small interrogation window are used to minimize the PIV measurement uncertainty due to the velocity gradient near the wall. Polythene beads with the diameter of 60 mu m (d(p)(+) = 1.71, normalized by the fluid kinematic viscosity nu and u(tau)) are used as dispersed phase, and three low mass loading ratios (Phi(m)) ranging from 10(-4) to 10(-3) are tested. It is found that the addition of the particles noticeably modifies the mean velocity and turbulent intensities of the gas-phase, as well as the turbulence coherent structures, even at Phi(m) = 0.025%. Particle inertia changes the viscous sublayer of the gas turbulence with a smaller thickness and a larger streamwise velocity gradient, which increases the peak value of the streamwise fluctuation velocity (u(rms)(+)) of the gas-phase with its location shifting to the wall. Particle sedimentation increases the roughness of the bottom wall, which significantly increases the wall-normal fluctuation velocity (nu(+)(rms)) and Reynolds shear stress (-< u'nu'>(+)) of the gas-phase in the inner region of the boundary layer (y(+) < 10). Under effect of particle-wall collision, the Q2 events (ejections) of the gas-phase are slightly increased by particles, while the Q4 events (sweeps) are obviously decreased. The spatial scale of the coherent structures near the wall shrinks remarkably with the presence of the particles, which may be attributed to the intensified crossing-trajectory effects due to particle saltation near the bottom wall. Meanwhile, the nu(+)(rms) and -< u'nu'>(+) of the gas-phase are significantly reduced in the outer region of the boundary layer (y (+)> 20).

Liljegren L M, Vlachos N S. 1990.

Laser velocimetry measurements in a horizontal gas-solid pipe flow

Experiments in Fluids, 9:205-212.

DOI      URL     [本文引用: 1]

Liu H, Bo T, Liang Y. 2017 a.

The variation of large-scale structure inclination angles in high Reynolds number atmospheric surface layers

Physics of Fluids, 29:035104.

DOI      URL     [本文引用: 6]

Liu H, Wang G, Zheng X. 2017 b.

Spatial length scales of large-scale structures in atmospheric surface layers

Physical Review Fluids, 2:064606.

DOI      URL     [本文引用: 1]

Liu H, Wang G, Zheng X. 2019.

Amplitude modulation between multi-scale turbulent motions in high-Reynolds-number atmospheric surface layers

Journal of Fluid Mechanics, 861:585-607.

DOI      URL     [本文引用: 3]

Liu Z, Adrian R J, Hanratty T J. 2001.

Large-scale modes of turbulent channel flow: Transport and structure

Journal of Fluid Mechanics, 448:53-80.

DOI      URL     [本文引用: 1]

Ljus C, Johansson B, Almstedt A E. 2002.

Turbulence modification by particles in a horizontal pipe flow

International Journal of Multiphase Flow, 28:1075-1090.

DOI      URL     [本文引用: 2]

Abstract

Measurements were made of turbulence intensities and turbulent energy spectra in a fully developed, turbulent air–particle pipe flow. The influence of the particles on the turbulence was studied. Measurements were made with spherical particles and particles with a large aspect ratio (pulp fibres). There is a significant change in turbulence intensity at higher particle concentrations with loading ratios of m=0.1 and 0.03. The measurements show that the turbulence intensity increases close to the centre of the pipe while the turbulence intensity decreases close to the pipe wall for the spherical particles. These results are in agreement with earlier measurements found in the literature. For the fibres, the turbulence intensity decreases over the whole pipe cross-section. Fibre flocs, however, give variations in the mean velocity that result in the production of turbulence in the lower part of the channel.

Lozano-Durán A, Jiménez J. 2014.

Effect of the computational domain on direct simulations of turbulent channels up to $Re_\tau= 4200$

Physics of Fluids, 26:011702.

DOI      URL     [本文引用: 1]

Lucci F, Ferrante A, Elghobashi S. 2011.

Is Stokes number an appropriate indicator for turbulence modulation by particles of taylor-length-scale size?

Physics of Fluids, 23:025101.

DOI      URL     [本文引用: 1]

Luo K, Hu C, Wu F. 2017.

Direct numerical simulation of turbulent boundary layer with fully resolved particles at low volume fraction

Physics of Fluids, 29:053301.

DOI      URL     PMID      [本文引用: 1]

In the present work, a direct numerical simulation (DNS) of dilute particulate flow in a turbulent boundary layer has been conducted, containing thousands of finite-sized solid rigid particles. The particle surfaces are resolved with the multi-direct forcing immersed-boundary method. This is, to the best of the authors' knowledge, the first DNS study of a turbulent boundary layer laden with finite-sized particles. The particles have a diameter of approximately 11.3 wall units, a density of 3.3 times that of the fluid, and a solid volume fraction of 1/1000. The simulation shows that the onset and the completion of the transition processes are shifted earlier with the inclusion of the solid phase and that the resulting streamwise mean velocity of the boundary layer in the particle-laden case is almost consistent with the results of the single-phase case. At the same time, relatively stronger particle movements are observed in the near-wall regions, due to the driving of the counterrotating streamwise vortexes. As a result, increased levels of dissipation occur on the particle surfaces, and the root mean square of the fluctuating velocities of the fluid in the near-wall regions is decreased. Under the present parameters, including the particle Stokes number St(+) = 24 and the particle Reynolds number Rep = 33 based on the maximum instantaneous fluid-solid velocity lag, no vortex shedding behind the particle is observed. Lastly, a trajectory analysis of the particles shows the influence of turbophoresis on particle wall-normal concentration, and the particles that originated between y(+) = 60 and 2/3 of the boundary-layer thickness are the most influenced.

Luo K, Fan J, Cen K. 2005.

Modulations on turbulent characteristics by dispersed particles in gas-solid jets

Proceedings of the Royal Society A: Mathematical, Physical and Engineering Science, 461:3279-3295.

[本文引用: 1]

Mandø M. 2009.

Turbulence modulation by non-spherical particles

[PhD Thesis]. Aalborg: Department of Energy Technology, Aalborg University.

Marchioli C, Soldati A. 2002.

Mechanisms for particle transfer and segregation in a turbulent boundary layer

Journal of Fluid Mechanics, 468:283-315.

DOI      URL     [本文引用: 1]

Martin R L, Kok J F. 2018.

Distinct thresholds for the initiation and cessation of aeolian saltation from field measurements

Journal of Geophysical Research: Earth Surface, 123:1546-1565.

[本文引用: 1]

Marusic I, Heuer W D C. 2007.

Reynolds number invariance of the structure inclination angle in wall turbulence

Physical Review Letters, 99:114504.

DOI      URL     PMID      [本文引用: 2]

Cross correlations of the fluctuating wall-shear stress and the streamwise velocity in the logarithmic region of turbulent boundary layers are reported over 3 orders of magnitude change in Reynolds number. These results are obtained using hot-film and hot-wire anemometry in a wind tunnel facility, and sonic anemometers and a purpose-built wall-shear stress sensor in the near-neutral atmospheric surface layer on the salt flats of Utah's western desert. The direct measurement of fluctuating wall-shear stress in the atmospheric surface layer has not been available before. Structure inclination angles are inferred from the cross correlation results and are found to be invariant over the large range of Reynolds number. The findings justify the prior use of low Reynolds number experiments for obtaining structure angles for near-wall models in the large-eddy simulation of atmospheric surface layer flows.

Marusic I, Hutchins N. 2008.

Study of the log-layer structure in wall turbulence over a very large range of Reynolds number

Flow Turbulence & Combustion, 81:115-130.

[本文引用: 1]

Marusic I, Kunkel G J. 2003.

Streamwise turbulence intensity formulation for flat-plate boundary layers

Physics of Fluids, 15:2461-2464.

DOI      URL     [本文引用: 2]

Marusic I, Kunkel G J, Porte-Agel F. 2001.

Experimental study of wall boundary conditions for large-eddy simulation

Journal of Fluid Mechanics, 446:309-320.

DOI      URL     [本文引用: 2]

Marusic I, Mathis R, Hutchins N. 2010 a.

High Reynolds number effects in wall turbulence

International Journal of Heat and Fluid Flow, 31:418-428.

DOI      URL     [本文引用: 3]

Abstract

A review of recent advances in the study of high Reynolds number turbulent boundary layers is given. The emergent regime of very large-scale structures in the logarithmic region and their subsequent influence on the near-wall cycle challenges many of the previously held assumptions regarding scaling of turbulent boundary layers at high Reynolds numbers. Experimental results are presented to illustrate the superimposition of large-scale energy onto the near-wall cycle, together with an interaction well described by an amplitude modulation effect. Both phenomena are shown to increase in magnitude (as compared to viscous-scaled events) as Reynolds number increases. These observations lead to a possible model for a statistically representative near-wall velocity signal (giving accurate energy spectra) based on a given filtered velocity signal from the log region of a high Reynolds number turbulent flow.

Marusic I, Mathis R, Hutchins N. 2010 b.

Predictive model for wall-bounded turbulent flow

Science, 329:193-196.

DOI      URL     PMID      [本文引用: 2]

The behavior of turbulent fluid motion, particularly in the thin chaotic fluid layers immediately adjacent to solid boundaries, can be difficult to understand or predict. These layers account for up to 50% of the aerodynamic drag on modern airliners and occupy the first 100 meters or so of the atmosphere, thus governing wider meteorological phenomena. The physics of these layers is such that the most important processes occur very close to the solid boundary--the region where accurate measurements and simulations are most challenging. We propose a mathematical model to predict the near-wall turbulence given only large-scale information from the outer boundary layer region. This predictive capability may enable new strategies for the control of turbulence and may provide a basis for improved engineering and weather prediction simulations.

Marusic I, Mathis R, Hutchins N. 2011.

A wall-shear stress predictive model

Journal of Physics: Conference Series, 318:012003.

[本文引用: 1]

Marusic I, McKeon B J, Monkewitz P A, Nagib H M, Smits A J, Sreenivasan K R. 2010 c.

Wall-bounded turbulent flows at high Reynolds numbers: Recent advances and key issues

Physics of Fluids, 22:065102.

DOI      URL     [本文引用: 3]

Marusic I, Monty J P, Hultmark M, Smits A J. 2013.

On the logarithmic region in wall turbulence

Journal of Fluid Mechanics, 716:R3.

DOI      URL     [本文引用: 5]

Mathis R, Hutchins N, Marusic I. 2009 a.

Large-scale amplitude modulation of the small-scale structures in turbulent boundary layers

Journal of Fluid Mechanics, 628:311-337.

DOI      URL     [本文引用: 4]

Mathis R, Hutchins N, Marusic I. 2011 a.

A predictive inner-outer model for streamwise turbulence statistics in wall-bounded flows

Journal of Fluid Mechanics, 681:537-566.

DOI      URL     [本文引用: 2]

A model is proposed with which the statistics of the fluctuating streamwise velocity in the inner region of wall-bounded turbulent flows are predicted from a measured large-scale velocity signature from an outer position in the logarithmic region of the flow. Results, including spectra and all moments up to sixth order, are shown and compared to experimental data for zero-pressure-gradient flows over a large range of Reynolds numbers. The model uses universal time-series and constants that were empirically determined from zero-pressure-gradient boundary layer data. In order to test the applicability of these for other flows, the model is also applied to channel, pipe and adverse-pressure-gradient flows. The results support the concept of a universal inner region that is modified through a modulation and superposition of the large-scale outer motions, which are specific to the geometry or imposed streamwise pressure gradient acting on the flow.

Mathis R, Marusic I, Chernyshenko S I, Hutchins N. 2013.

Estimating wall-shear-stress fluctuations given an outer region input

Journal of Fluid Mechanics, 715:163-180.

DOI      URL    

Mathis R, Marusic I, Hutchins N, Sreenivasan K R. 2011 b.

The relationship between the velocity skewness and the amplitude modulation of the small scale by the large scale in turbulent boundary layers

Physics of Fluids, 23:121702.

DOI      URL    

Mathis R, Monty J P, Hutchins N, Marusic I. 2009 b.

Comparison of large-scale amplitude modulation in turbulent boundary layers, pipes, and channel flows

Physics of Fluids, 21:111703.

DOI      URL     [本文引用: 1]

McEwan I K, Willetts B B. 1991.

Numerical model of the saltation cloud. Aeolian Grain Transport 1

Acta Mechanica (Suppl), 53-66.

[本文引用: 1]

Mckeon B J, Li J, Jiang W, Morrison J F, Smits A J. 2004.

Further observations on the mean velocity distribution in fully developed pipe flow

Journal of Fluid Mechanics, 501:135-147.

DOI      URL     [本文引用: 1]

Mckeon B J, Morrison J F. 2007.

Asymptotic scaling in turbulent pipe flow

Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 365:771-787.

DOI      URL     [本文引用: 2]

McLaughlin J B. 1989.

Aerosol particle deposition in numerically simulated channel flow

Physics of Fluids A: Fluid Dynamics, 1:1211-1224.

DOI      URL     [本文引用: 1]

Metzger M M, Klewicki J C. 2001.

A comparative study of near-wall turbulence in high and low Reynolds number boundary layers

Physics of Fluids, 13:692-701.

DOI      URL     [本文引用: 2]

Millikan C B. 1938.

A critical discussion of turbulent flows in channels and circular tubes

//Proceedings of the Fifth International Congress of Applied Mechanics, 386-392.

[本文引用: 1]

Moin P, Kim J. 1982.

Numerical investigation of turbulent channel flow

Journal of Fluid Mechanics, 118:1280-1284.

[本文引用: 1]

Moin P, Mahesh K. 1998.

Direct numerical simulation: A tool in turbulence research

Annual Review of Fluid Mechanics, 30:539-578.

DOI      URL     [本文引用: 1]

Monty J P, Stewart J A, Williams R C, Chong M S. 2007.

Large-scale features in turbulent pipe and channel flows

Journal of Fluid Mechanics, 589:147-156.

DOI      URL     [本文引用: 3]

Monty J P. 2005.

Developments in smooth wall turbulent duct flows. [PhD Thesis]

Melbourne: Department of Mechanical and Manufacturing Engineering, University of Melbourne.

[本文引用: 1]

Morris S C, Stolpa S R, Slaboch P E, Klewicki J C. 2007.

Near-surface particle image velocimetry measurements in a transitionally rough-wall atmospheric boundary layer

Journal of Fluid Mechanics, 580:319-338.

DOI      URL     [本文引用: 3]

Morrison J F, Mckeon B J, Jiang W, Smits A J. 2004.

Scaling of the streamwise velocity component in turbulent pipe flow

Journal of Fluid Mechanics, 508:99-131.

DOI      URL     [本文引用: 2]

Moser R D, Kim J, Mansour N N. 1999.

Direct numerical simulation of turbulent channel flow up to $Re_\tau=590$

Physics of Fluids, 11:943-945.

DOI      URL     [本文引用: 1]

Nagib H M, Chauhan K A. 2008.

Variations of von Kármán coefficient in canonical flows

Physics of Fluids, 20:101518.

DOI      URL     [本文引用: 3]

Nagib H M, Chauhan K A, Monkewitz P A. 2007.

Approach to an asymptotic state for zero pressure gradient turbulent boundary layers

Philosophical Transactions of the Royal Society A-Mathematical Physical And Engineering Sciences, 365:755-770.

DOI      URL     [本文引用: 2]

Nickels T B, Marusic I, Hafez S, Hutchins N, Chong M S. 2007.

Some predictions of the attached eddy model for a high Reynolds number boundary layer

Philosophical Transactions of the Royal Society A-Mathematical Physical And Engineering Sciences, 365:807-822.

DOI      URL     [本文引用: 1]

Ninto Y, Garcia M H. 1996.

Experiments on particle—turbulence interactions in the near-wall region of an open channel flow: Implications for sediment transport

Journal of Fluid Mechanics, 326:285-319.

DOI      URL     [本文引用: 1]

Oertel H. 2005.

Prandtl's Essntials of Fluid Mechanics. 2nd edn

Springer.

[本文引用: 1]

Offen G, Kline S. 1975.

A proposed model of the bursting process in turbulent boundary layers

Journal of Fluid Mechanics, 70:209-228.

DOI      URL     [本文引用: 1]

Orszag S A, Patterson G S. 1972.

Numerical simulation of turbulence

//M. Rosenblatt bt C. Van Atta eds.Statistical Models of' Ttirhulciice., New York: Springer.

[本文引用: 1]

österlund J M, Johansson A V, Nagib H M, Hites M H. 2000.

A note on the overlap region in turbulent boundary layers

Physics of Fluids, 12:1-4.

DOI      URL     [本文引用: 2]

Owen P R. 1969.

Pneumatic transport

Journal of Fluid Mechanics, 39:407-432.

DOI      URL     [本文引用: 1]

Pan Y, Banerjee S. 1996.

Numerical simulation of particle interactions with wall turbulence

Physics of Fluids, 8:2733-2755.

DOI      URL     [本文引用: 1]

Panton R L. 2001.

Overview of the self-sustaining mechanisms of wall turbulence

Progress in Aerospace Sciences, 37:341-383.

DOI      URL     [本文引用: 1]

Pathikonda G, Christensen K T. 2017.

Inner-outer interactions in a turbulent boundary layer overlying complex roughness

physical Review Fluids, 2:044603.

DOI      URL     [本文引用: 1]

Perry A E, Henbest S, Chong M S. 1986.

A theoretical and experimental study of wall turbulence

Journal of Fluid Mechanics, 165:163-199.

DOI      URL     [本文引用: 1]

Porté Agel F. 2004.

The role of coherent structures in subfilter-scale dissipation of turbulence measured in the atmospheric surface layer

Journal of Turbulence, 5:040.

Portela L M, Oliemans R V A. 2003.

Eulerian-Lagrangian DNS/LES of particle-turbulence interactions in wall-bounded flows

International Journal for Numerical Methods in Fluids, 43:1045-1065.

DOI      URL     [本文引用: 1]

Prandtl L. 1910.

Eine Beziehung zwischen Warmeaustausch and Stromungswiderstand der Flussigkeiten

Phys. Z., 11:1072-1078.

[本文引用: 1]

Prandtl L. 1925.

Bericht über Untersuchungen zur ausgebildeten Turbulenz

ZAMM-Journal of Applied Mathematics and Mechanics/Zeitschrift für Angewandte Mathematik und Mechanik, 5:136-139.

[本文引用: 3]

Rashidi Mr, Hetsroni G, Banerjee S. 1990.

Particle-turbulence interaction in a boundary layer

International Journal of Multiphase Flow, 16:935-949.

DOI      URL     [本文引用: 2]

Rasmussen K R, Sørensen M. 1999.

Aeolian mass transport near the saltation threshold

Earth Surface Processes and Landforms, 24:413-422.

[本文引用: 2]

Reichardt H. 1933.

Die Quadratischen Mittelwerte der Lanesschwankungen in der Turbulenten Kanalstromung

Zeitschrift für Angewandte Mathematik und Mechanik (ZAMM), 3:177-180.

Reynolds O. 1894.

On the dynamical theory of incompressible viscous fluids and the determination of the criterion

Philosophical Transactions of the Royal Society of London A, 186:123-164.

[本文引用: 1]

Richter D H, Sullivan P P. 2014.

Modification of near-wall coherent structures by inertial particles

Physics of Fluids, 26:103304.

[本文引用: 1]

Righetti M, Giovanni P R. 2004.

Particle-fluid interactions in a plane near-wall turbulent flow

Journal of Fluid Mechanics, 505:93-121.

[本文引用: 2]

Robinson S K. 1991.

Coherent motions in the turbulent boundary layer

Annual Review of Fluid Mechanics, 23:601-639.

[本文引用: 1]

Rogers C B, Eaton J K. 1991.

The effect of small particles on fluid turbulence in a flat-plate, turbulent boundary layer in air

Physics of Fluids A: Fluid Dynamics, 3:928-937.

[本文引用: 1]

Samie M, Marusic I, Hutchins N, Fu M K, Fan Y, Hultmark M, Smits A J. 2018.

Fully resolved measurements of turbulent boundary layer flows up to $Re_\tau=20000$

Journal of Fluid Mechanics, 851, 391-415.

[本文引用: 2]

Sato Y, Koichi H. 1996.

Transport process of turbulence energy in particle-laden turbulent flow

International Journal of Heat and Fluid Flow, 17:202-210.

[本文引用: 1]

Schlatter P, örlü R. 2010.

Quantifying the interaction between large and small scales in wall-bounded turbulent flows: A note of caution

Physicsof Fluids, 22:051704.

[本文引用: 1]

Schlichting H, Gersten K, Krause E, Oertel H J, Mayes C. 2000.

Boundary Layer Theory

8th edn. Springer.

[本文引用: 1]

Schultz M P, Flack K A. 2013.

Reynolds-number scaling of turbulent channel flow

Physics of Fluids, 25:011702.

[本文引用: 1]

Serafimovich A, Thomas C, Foken T. 2011.

Vertical and horizontal transport of energy and matter by coherent motions in a tall spruce canopy

Boundary-Layer Meteorology, 140:429-451.

[本文引用: 1]

Sillero J A, Jimenez J, Moser R D. 2013.

One-point statistics for turbulent wall-bounded flows at Reynolds numbers up to $\delta^{+}$ approximate to 2000

Physics of Fluids, 25:105102.

[本文引用: 1]

Sillero J A, Jimenez J, Moser R D. 2014.

Two-point statistics for turbulent boundary layers and channels at Reynolds numbers up to $\delta^{+}$ approximate to 2000

Physics of Fluids, 26:105109.

[本文引用: 1]

Smits A J, Mckeon B J, Marusic I. 2011.

High-Reynolds number wall turbulence

Annual Review of Fluid Mechanics, 43:353-375.

[本文引用: 2]

Spalart P R. 1988.

Vortex methods for separated flows

// VKI, Computational Fluid Dynamics.

[本文引用: 1]

Squire D T, Baars W J, Hutchins N, Marusic I. 2016.

Inner-outer interactions in rough-wall turbulence

Journal of Turbulence, 17:1159-1172.

[本文引用: 1]

Squire D T, Hutchins N, Morrill-Winter C, Schultz M P, Klewicki J C, Marusic I. 2017.

Applicability of Taylor's hypothesis in rough and smooth-wall boundary layers

Journal of Fluid Mechanics, 812:398-417.

[本文引用: 1]

Stout J E, Zobeck T M. 1997.

Intermittent saltation

Sedimentology, 44:959-970.

DOI      URL     [本文引用: 1]

Talluru K M, Baidya R, Hutchins N, Marusic I. 2014.

Amplitude modulation of all three velocity components in turbulent boundary layers

Journal of Fluid Mechanics, 746:R1.

[本文引用: 2]

Tanaka T, Eaton, J.K.. 2008.

Classification of turbulence modification by dispersed spheres using a novel di-mensionless number

Physical Review Letters, 101:114502.

DOI      URL     PMID      [本文引用: 1]

A novel dimensionless parameter, the particle moment number Pa, was derived using dimensional analysis of the particle-laden Navier-Stokes equations, in order to understand the underlying physics of turbulence modification by particles. A set of 80 previous experimental measurements where the turbulent kinetic energy was modified by particles was examined, and all results could clearly be divided into three groups in Re-Pa mappings. The turbulence attenuation region was observed between the augmentation regions with two critical particle momentum numbers.

Tanaka T, Eaton J K. 2010.

Sub-Kolmogorov resolution partical image velocimetry measurements of particle laden forced turbulence

Journal of Fluid Mechanics, 643:177-206.

[本文引用: 1]

Taniere A, Oesterle B, Monnier J C. 1997.

On the behaviour of solid particles in a horizontal boundary layer with turbulence and saltation effects

Experiments in Fluids, 23:463-471.

[本文引用: 1]

Tay Godwin F K, Kuhn D C S, Tachie M F. 2015.

Effects of sedimenting particles on the turbulence structure in a horizontal channel flow

Physics of Fluids, 27:025106.

[本文引用: 1]

Theodorsen T. 1952.

Mechanism of turbulence

// Proceedings of the Second Midwestern Conference on Fluid Mechanics. Ohio State University, USA.

[本文引用: 1]

Theodorsen T. 1955

The structure of turbulence

// 50 Jahre Grenzschichtforschung. H. Gortler & W. Tollmien Vieweg and Sohn. pp52-56.

[本文引用: 1]

Tomkins C D. 1997.

A partucle image velocimetry study of ciherent structures in a turbulent boundary layer

[PhD Thesis]. Urbana-Champaign: University of Illions.

[本文引用: 1]

Tomkins C D, Adrian R J. 2003.

Spanwise structure and scale growth in turbulent boundary layers

Journal of Fluid Mechanics, 490:37-74.

[本文引用: 1]

Tritton D J. 1967.

Some new correlation measurements in a turbulent boundary layer

Journal of Fluid Mechanics, 28:439-462.

[本文引用: 1]

Tsuji Y, Marusic I, Johansson A V. 2016.

Amplitude modulation of pressure in turbulent boundary layer

International Journal of Heat And Fluid Flow, 61:2-11.

[本文引用: 1]

Tsuji Y, Morikawa Y. 1982.

LDV measurements of an air-solid two-phase flow in a horizontal pipe

Journal of Fluid Mechanics, 120:385-409.

[本文引用: 2]

Tsuji Y, Morikawa Y, Shiomi H. 1984.

LDV measurements of an air-solid two-phase flow in a vertical pipe

Journal of Fluid Mechanics, 139:417-434.

[本文引用: 1]

Vallikivi M, Hultmark M, Smits A J. 2015 a.

Turbulent boundary layer statistics at very high Reynolds number

Journal of Fluid Mechanics, 779:371-389.

[本文引用: 4]

Vallikivi M, Ganapathisubramani B, Smits A J. 2015 b.

Spectral scaling in boundary layers and pipes at very high Reynolds numbers

Journal of Fluid Mechanics, 771:303-326.

[本文引用: 2]

Varaksin A Y, Polezhaev Y V, Polyakov A F. 2000.

Effect of particle concentration on fluctuating velocity of the disperse phase for turbulent pipe flow

International Journal of Heat and Fluid Flow, 21:562-567.

[本文引用: 1]

Vincenti P, Klewicki J, Morrill-Winter C, White C M, Wosnik M. 2013.

Streamwise velocity statistics in turbulent boundary layers that spatially develop to high Reynolds number

Experiments in Fluids, 54:1629.

[本文引用: 2]

Von Kármán T. 1930.

Mechanische änlichkeit und turbulenz

Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse, 1930: 58-76.

[本文引用: 1]

Wang G, Zheng X. 2016.

Very large scale motions in the atmospheric surface layer: A field investigation

Journal of Fluid Mechanics. 802:464-489.

[本文引用: 6]

Wang G, Zheng X, Tao J. 2017.

Very large scale motions and PM10 concentration in a high-Re boundary layer

Physics of Fluids, 29:061701.

[本文引用: 4]

Wang P, Feng S, Zheng X, Sung H. 2019.

The scale characteristics and formation mechanism of aeolian sand streamers based on large eddy simulation,

Journal of Geophysical Research Atmosphere, 124:11372-11388.

[本文引用: 3]

Wang J, Levy E K. 2006.

Particle behavior in the turbulent boundary layer of a dilute gas-particle flow past a flat plate

Experimental Thermal and Fluid Science, 30:473-483.

[本文引用: 1]

Wang Q, Squires K D. 1996.

Large eddy simulation of particle-laden turbulent channel flow

Physics of Fluid, 8:1207-1223.

[本文引用: 1]

Wang W, Pan C, Wang J. 2018.

Quasi-bivariate variational mode decomposition as a tool of scale analysis in wall-bounded turbulence

Experiments in Fluids, 59:1-18.

[本文引用: 1]

Wei T, Fife P, Klewicki J, Mcmurtry P. 2005.

Properties of the mean momentum balance in turbulent boundary layer, pipe and channel flows

Journal of Fluid Mechanics, 522:303-327.

[本文引用: 1]

Willert C, Eisfelder M, Stanislas M, Klinner J, Talamelli A. 2017.

Near-wall statistics of a turbulent pipe flow at shear Reynolds numbers up to 40000

Journal of Fluid Mechanics, 826:R5.

[本文引用: 3]

WMO. 1983.

Guide to Meteorological Instruments and Methods of Observation

5th edn. Geneva: The World Meteorological Organization.

[本文引用: 1]

Wu X, Moin P. 2009.

Forest of hairpins in a low-Reynolds-number zero-pressure-gradient flat-plate boundary layer

Physics of Fluids, 21:091106.

[本文引用: 1]

Yamamoto Y, Tsuji Y. 2018.

Numerical evidence of logarithmic regions in channel flow at $Re_\tau= 8000$

Physical Review Fluids, 3:012602.

[本文引用: 1]

Yang H, Bo T. 2018.

Scaling of wall-normal turbulence intensity and vertical eddy Structures in the atmospheric surface layer

Boundary-Layer Meteorology, 166:199-216.

[本文引用: 1]

Yang X, Sadique J, Mittal R, Meneveau C. 2015.

Integral wall model for large eddy simulations of wall-bounded turbulent flows

Physics of Fluids, 27:025112.

[本文引用: 1]

Yao Y, Huang W, Xu C. 2018.

Amplitude modulation and extreme events inturbulent channel flow

Acta Mechanica Sinica, 34:1-9.

[本文引用: 1]

Yin G, Huang W, Xu C. 2018.

Prediction of near-wall turbulence using minimal flow unit

Journal of Fluid Mechanics, 841:654-673.

[本文引用: 1]

Zagarola M V, Smits A J. 1998.

Mean-flow scaling of turbulent pipe flow

Journal of Fluid Mechanics, 373:33-79.

DOI      URL     [本文引用: 1]

Zeng Q, Cheng X, Hu F, Peng Z. 2010.

Gustiness and coherent structure of strong winds and their role in dust emission and entrainment

Advances in Atmospheric Sciences, 27:1-13.

DOI      URL     [本文引用: 2]

After the passage of a cold front, spring in northernChina, the outbreak of strong wind is often accompanied by dustemissions. Through analyses of data in the atmospheric boundarylayer during a typical case, it is revealed there are rather regulargust wave packets superimposed on the basic strong wind flow. Thegust-wind wave packets have a period equal to around 3--6 mins andpossess coherent structure. As the vertical transport of momentum isdecomposed into separate parts by (a) basic flow, (b) gust-wind,and (c) turbulence, they are all in a downwards direction at thelower levels of the atmospheric boundary layer during strong windperiods. However, (a) is the largest, while (b) and (c) arecomparable. All these are very different from the case of normalweather. Besides, the friction velocity at the ground surface isalso much larger than that of normal weather and should be correctedby taking the contributions of the basic flow and gust-wind intoaccount.The strong basic flow with descending motion is veryfavorable for soil erosion and sand/dust emissions, but suppressesthe entrainment of dust particles by keeping them within the bottomlevels of the atmospheric boundary layer. Owing to the coherentstructure of gust-wind, dust particles can effectively overcome thesystematic descending air motion and penetrate into the middle andupper levels of the atmospheric boundary layer, and then propagatefurther and diffuse into the troposphere where ascending air motionprevails.

Zhang W, Wang Y, Sang J L. 2008.

Simultaneous PIV and PTV measurements of wind and sand particle velocities

Experiments in Fluids, 45:241-256.

DOI      URL     [本文引用: 1]

Wind-blown sand is a typical example of two-phase particle-laden flows. Owing to lack of simultaneous measured data of the wind and wind-blown sand, interactions between them have not yet been fully understood. In this study, natural sand of 100–125 μm taken from Taklimakan Desert was tested at the freestream wind speed of 8.3 m/s in an atmospheric boundary layer wind tunnel. The captured flow images containing both saltating sand and small wind tracer particles, were separated by using a digital phase mask technique. The 2-D PIV (particle imaging velocimetry) and PTV (particle tracking velocimetry) techniques were employed to extract simultaneously the wind velocity field and the velocity field of dispersed sand particles, respectively. Comparison of the mean streamwise wind velocity profile and the turbulence statistics with and without sand transportation reveal a significant influence of sand movement on the wind field, especially in the dense saltating sand layer (y/δ < 0.1). The ensemble-averaged streamwise velocity profile of sand particles was also evaluated to investigate the velocity lag between the sand and the wind. This study would be helpful in improving the understanding of interactions between the wind and the wind-blown sand.

Zhang Y, Hu R, Zheng X. 2018.

Large-scale coherent structures of suspended dust concentration in the neutral atmospheric surface layer: A large-eddy simulation study

Physics of Fluids, 30:046601.

DOI      URL     [本文引用: 1]

Zhao R, Smits A J. 2006.

Binormal cooling errors in crossed hot-wire measurements

Experiments in Fluids, 40:212-217.

DOI      URL     [本文引用: 1]

In crossed hot-wire measurements, cooling of the filament in the direction normal to the hot-wire measurement plane will cause additional heat transfer, called binormal cooling. It is shown that binormal cooling can lead to significant errors in measuring the mean velocity and the turbulence levels in regions where the turbulence level is high, such as the near-wall region of pipes and boundary layers. In a near-wall measurement of a pipe flow, binormal cooling effects lead to errors up to 1.5% in mean velocity and 4% in turbulent shear stress.

Zhao R, Smits A J. 2007.

Scaling of the wall-normal turbulence component in high-Reynolds-number pipe flow

Journal of Fluid Mechanics, 576:457-473.

DOI      URL     [本文引用: 1]

Zheng X. 2009.

Mechanics of Wind-blown Sand Movements

Springer Science & Business Media.

[本文引用: 2]

Zheng X. 2018.

Modulations of sand particles on VLSMs in sand-laden flows

JFM Symposia: From Fundamentals to Applied Fluid Mechanics. November 9-11, Beijing.

[本文引用: 1]

Zheng X, Jin T, Wang P. 2020.

The influence of surface stress fluctuation on saltation sand transport around threshold

Journal of Geophysical Research: Earth Surface, 125: e2019JF005246.

DOI      URL     PMID      [本文引用: 1]

We investigated the frictional strength recovery (healing) and subsequent reactivation and slip-weakening behavior of simulated fault gouges derived from key stratigraphic units in the seismogenic Groningen gas field (N. E. Netherlands). Direct-shear, slide-hold-slide (SHS) experiments were performed at in situ conditions of 100 degrees C, 40 MPa effective normal stress and 10-15 MPa pore fluid pressure (synthetic formation brine). Sheared gouges were allowed to heal for periods up to 100 days before subsequent reshearing. The initial coefficient of (steady) sliding friction mu was highest in the Basal Zechstein caprock (mu = 0.65 +/- 0.02) and Slochteren sandstone reservoir (mu = 0.61 +/- 0.02) gouges, and the lowest in the Ten Boer claystone at the reservoir top (mu = 0.38 +/- 0.01) and in the Carboniferous shale substrate (mu approximately 0.45). Healing and subsequent reactivation led to a marked increase (mu) in (static) friction coefficient of up to ~0.16 in Basal Zechstein and ~0.07 in Slochteren sandstone gouges for the longest hold periods investigated, followed by a sharp strength drop (up to ~25%) and slip-weakening trajectory. By contrast, the Ten Boer and Carboniferous gouges showed virtually no healing or strength drop. Healing rates in the Basal Zechstein and Slochteren sandstone gouges were significantly affected by the stiffness of different machines used, in line with the Ruina slip law, and with a microphysical model for gouge healing. Our results point to marked stratigraphic variation in healed frictional strength and healing rate of faults in the Groningen system, and high seismogenic potential of healed faults cutting the reservoir and Basal Zechstein caprock units, upon reactivation.

Zheng X, Wang G, Bo T, Zhu W. 2015.

Field observations on the turbulent features of the near-surface flow fields and dust transport during dust storms

Procedia IUTAM, 17:13-19.

DOI      URL     [本文引用: 1]

Zheng X, Zhang J, Wang G, Liu H, Zhu W. 2013.

Investigation on very large scale motions (VLSMs) and their influence in a dust storm

Science China (Physics, Mechanics & Astronomy), 56:306-314.

DOI      URL     PMID      [本文引用: 1]

Beta cell mass and function are decreased to varying degrees in diabetes. Islet cell replacement or regenerative therapy may offer great therapeutic promise to people with diabetes. In addition to primary pancreatic beta cells, recent studies on regeneration of functional insulin producing cells (IPCs) revealed that several alternative cell sources, including embryonic stem cells, induced pluripotent stem cells and adult stem cells, can generate IPCs by differentiation, reprogramming, and trans-differentiation. In this review, we discuss stem cells as a potential alternative cell source for the treatment of diabetes.

/

Baidu
map