Wind Field Modeling over Hilly Terrain: A Review of Methods, Challenges, Limitations, and Future Directions
Abstract
1. Introduction
- To critically and comprehensively review the key technical advancements, inherent advantages, and core limitations of the four main paradigms in hilly terrain wind field modeling since 2015: field measurements, wind tunnel experiments, CFD simulations, and data-driven methods.
- To systematically evaluate the applicability, validation challenges, and existing knowledge gaps of these modeling methods in the three major engineering applications: wind resource development, wind-resistant design of infrastructure, and pollutant dispersion.
- To identify and analyze the deep-seated, systemic challenges currently facing the field (e.g., the contradiction between high-fidelity models and sparse validation data, the inherent complexity of multi-physics coupling), and to propose a clear and actionable roadmap for future research directions.
Review | Detail | Year | Key Findings |
---|---|---|---|
Bradley et al. [10] | Remote sensing winds | 2015 | Under the recirculating detached flow, the linear model may be inaccurate. |
Serafin et al. [8] | Airflow exchange | 2018 | Studying diurnal boundary layer variation aids analysis of exchange processes over mountains. |
Giovannini et al. [9] | Pollutant dispersion | 2020 | Pollutant dispersion modeling is more complicated over complex terrain. |
Finnigan et al. [11] | Gravity-driven flow | 2020 | The study of gravity-driven flow on hillsides and valley slopes faces challenges. |
Wani et al. [2] | Wind tunnel test | 2021 | More experimental studies on the cliff are needed. |
Farina et al. [12] | Heat-driven wind | 2023 | Future efforts should focus on field studies over near-ideal slopes. |
2. Review Method
- The document must be a peer-reviewed journal article or a significant conference paper.
- The publication must be in English.
- The research must directly focus on the modeling or validation of wind fields over hilly/complex terrain, or their application in areas like wind energy, infrastructure, and pollutant dispersion.
- Studies with low relevance to the main topic.
- Abstracts, short communications, or non-peer-reviewed literature.
- Duplicate publications (for conference papers later extended into journal articles, the more comprehensive journal version was prioritized).
- Shift in Research Objects: “Complex terrain” is the most frequent and central node, far surpassing idealized keywords like “2-D hill.” This indicates a decisive shift in research focus since 2015 from idealized models to more realistic and practical complex terrains.
- Evolution of Research Methods: Methodologically, “numerical simulation,” “CFD,” and “large-eddy simulation” form a prominent cluster. Their node sizes and link densities demonstrate that numerical methods, represented by CFD, have become the dominant research tools. The frequent appearance of “large-eddy simulation,” in particular, aligns with our assessment of HPC advancements driving high-fidelity simulations post-2015. Meanwhile, “wind tunnel” and “field measurement,” as classical physical methods, remain important nodes closely connected to numerical simulation, highlighting their indispensable role in providing validation data.
- Driving Forces of Research Needs: On the application level, keywords such as “wind energy,” “atmospheric boundary layer,” “turbulence,” and “infrastructure design” (e.g., “bridge site”) also hold significant positions. This illustrates that two major engineering demands—wind resource development and infrastructure safety—are the core drivers for research on wind fields over hilly terrain.
3. Methods for Wind Field Modeling over Hilly Terrain
3.1. Field Measurement
Author (Year) | Location | Mountain Feature | Parameter | Detail | Device |
---|---|---|---|---|---|
Russell et al. [31] (2016) | Idaho, the United States | A complex terrain | WS, WD, AT | The forest canopy | Sonic anemometer |
Fenerci et al. [32] (2017) | Hordaland, Norway | Hardangerfjord | WS | Dynamic response, wind characteristics | Sonic anemometer, accelerometer |
Chaurasiya et al. [33] (2018) | Karyatal, India | A gentle slope | WS, WD | wind resource assessment | SODAR, LIDAR |
Huang et al. [34] (2019) | western China | A “Y”-shaped valley | WS, WD, AT, RH, AP | thunderstorm wind; thermally developed wind | Ultrasonic anemometers |
Zhang et al. [35] (2020) | Southwest of China | A typical canyon | WS, WD, AT | Wind resource assessment | Meteorological mast, SODAR; automatic meteorological station |
Radünz et al. [36] (2021) | Morrinhos, Brazil | A mixture of hills, ridges and plateaus | WS, WD, AT | Atmospheric stability | Sonic anemometer |
Radünz et al. [37] (2022) | Northeastern Brazil | A plateau | WS, WD, AT | Wind farm design | Meteorological mast |
Jiang et al. [38] (2023) | southwest of China | A U-shaped canyon | WS, WD, AT, P | Mixed wind climate | Sonic anemometer, automatic meteorological station, meteorological mast |
Adler et al. [39] (2021) | Inn Valley, Austria | Alpine valley | AT, WS, RH, TI | Atmospheric phenomenon | Doppler, microwave radiometer, unmanned aerial vehicle, ground flux tower array |
Coimbra et al. [40] (2025) | Perdigão, Portugal | Double parallel ridges | WS, TI | Wind field characteristics | Dual Doppler scanning Lidar, acoustic anemometer |
Desnijder et al. [41] (2024) | Krummendeich, Germany | Wind farms in complex terrains | WS, TI | Wind farm design | High-frequency sensor, unmanned aerial vehicle swarm |
Instrument | Advantages | Limitations | Cost | Typical Accuracy/Uncertainty | Spatial/Temporal Resolution | Ideal Application Environment |
---|---|---|---|---|---|---|
Ultrasonic Anemometer | Provides high-precision 3D wind velocity and high-frequency turbulence data. | Limited to single-point measurements, leading to poor spatial representativeness; high installation and maintenance costs. | Moderate instrument cost, but high costs for measurement towers and long-term maintenance. | Wind speed uncertainty < 0.1 m/s; accuracy of approx. ±1% of reading. | Spatial: Point measurement (~cm) Temporal: >10 Hz | Long-term, fixed-point, high-resolution observation of boundary layer structure and turbulence at a specific location. |
Doppler Lidar | High spatial coverage and relatively good portability. | Performance degrades in extreme weather; the instrument is expensive and has a near-ground measurement blind zone. | Very high acquisition and maintenance costs. | Wind speed accuracy: ±0.1~0.2 m/s; affected by volume averaging effects. | Spatial: Gate length 10–50 m Temporal: ~0.1–1 Hz (scanning mode) | Capturing meso-scale wind field structures, such as the evolution of canyon and mountain-valley winds. |
Sodar | Effectively measures low-level wind profiles; relatively sensitive to temperature fluctuations. | Data quality is susceptible to background noise and ground clutter; vertical detection range and stability are generally inferior to Lidar. | Significantly lower cost compared to Lidar. | Wind speed accuracy: ±0.2~0.5 m/s. | Spatial: Vertical resolution 10–30 m Temporal: On the order of minutes | Studies of the urban boundary layer, low-level atmospheric stability, and the nocturnal stable boundary layer structure. |
UAV (Unmanned Aerial Vehicle) | High mobility and portability; enables flexible measurements to rapidly acquire high-spatial-resolution data. | Limited by flight endurance and weather conditions; rotor wash can disturb the local flow field, requiring data correction. | Relatively low hardware acquisition and operational costs. | Sensor-dependent; post-correction wind speed uncertainty >0.5 m/s. | Spatial: Variable (~m) Temporal: Dependent on flight speed | A powerful complement to traditional methods, for probing fine-scale structures (e.g., ridge flow, wake regions) and measuring in inaccessible areas. |
3.2. Wind Tunnel Experiment
Author (Year) | Mountain Feature | Reduced Scale | Parameter | Detail | Device | Re |
---|---|---|---|---|---|---|
Mattuella et al. [47] (2016) | Actual complex terrain | 1:1000 | WS, TI | Wind power utilization | Hot-wire anemometer | - |
Li et al. [48] (2017) | Actual canyon terrain | 1:1000 | WS, WD | Bridge site | Cobra probe | - |
Lystad et al. [53] (2018) | Actual complex terrain | 1:2000 | WS, WD, TI | Long-span bridge; nonuniform wind field | Hot-wire anemometer | ≈7.1 × 105 |
Tian et al. [54] (2018) | 2D hill; 3D hill | 1:320 | WS, TI | Wind power utilization | Cobra probe | <7000 |
Kamada et al. [55] (2019) | 2D hill | 1:200 | WS, TI | ABL | PIV | ≈2.5 × 105 |
Shen et al. [56] (2021) | Actual mountain pass | 1:1000 | WS, WD | Long-span bridge; nonuniform wind field | Cobra probe | - |
Zhu et al. [57] (2022) | 2D slope | 1:20 | WS, TI | Railway infrastructure | Cobra probe | ≈1.0 × 105 |
Raffaele et al. [58] (2023) | 2D hill | - | WS, TI | Wind–sand coupling | PIV | =7.4 × 104 |
Wu et al. [59] (2025) | 2D hill | 1:1000 | WS, TI | Pollutant dispersion | LDV | - |
Instrument | Key Measured Parameters | Accuracy | Resolution | Advantages | Limitations | Cost and Feasibility | Ideal Applications |
---|---|---|---|---|---|---|---|
Hot-wire Anemometer (HWA) | Velocity magnitude, high-frequency turbulent fluctuations | ~1–2% | Spatial: Very high (~1 mm) Temporal: Very high (>100 kHz) | Excellent for measuring turbulence spectra and Reynolds stresses. Relatively low cost | Intrusive and fragile probe. Requires frequent calibration; sensitive to temperature. | Medium cost, high maintenance. Probes are fragile and need frequent calibration; time-consuming for large-scale flow mapping. | Characterizing fine-scale turbulence in attached flows over hills. Quantifying Turbulence Intensity (TI) for wind energy. |
Cobra Probe | 3D mean velocity, turbulence intensity | ~2–5% | Spatial: Moderate (~5–10 mm) Temporal: Low (~500 Hz) | Robust; provides 3D velocity vectors. Requires less frequent calibration | Intrusive; limited frequency response fails to capture fine turbulence. Accuracy degrades in high turbulence or reverse flow. | Higher cost, but durable. Simple operation, but low-frequency response is unsuitable for detailed turbulence studies. | Routine mapping of mean wind fields in complex terrain. Studies requiring robust instruments (e.g., bridge siting). |
Laser Doppler Velocimetry (LDV) | 1, 2, or 3 components of instantaneous velocity | ~1% | Spatial: High (measurement volume ~0.1 mm) Temporal: High (>10 kHz) | Non-intrusive and highly accurate. Can measure reverse flow; no calibration needed | Requires tracer particles and expensive equipment. Time-consuming for flow field mapping (single-point). | High cost. A key advantage is its non-intrusive measurement, but it has low efficiency. | Precise, non-intrusive measurements at critical points (e.g., separation bubbles). CFD validation; pollutant dispersion studies. |
Particle Image Velocimetry (PIV) | Instantaneous 2D or 3D velocity fields (“snapshots” of the flow) | ~2–5% | Spatial: Good (dependent on camera resolution) Temporal: Low (typical systems <20 Hz) | Non-intrusive; provides full-field data. Excellent for visualizing complex flow structures | Requires optical access, particle seeding, and expensive equipment. Lower temporal resolution. | Highest cost. Provides instantaneous full-field data, but post-processing is complex. | Investigating large-scale turbulence, separation, and recirculation zones over hills |
Surface Pressure Taps/Pressure Scanner | Mean and fluctuating surface pressure; pressure coefficient (C_p) | ~1% | Spatial: High (dependent on tap density) Temporal: Variable (high-speed scanners can reach up to 500 kHz) | Directly measures wind loads on surfaces | Cannot provide flow field data away from the surface. Difficult to install on models with complex curvature. | The sensor itself is not expensive, but model fabrication is costly. Directly measures load data but offers no flow field information. | Wind load studies on structures in hilly terrain. Not used for characterizing free-stream flow |
3.3. CFD Simulation
Author (Year) | Turbulence Model | Mountain Height/Mesh Size | Mountain Feature | Reduced Scale | Detail | Tool |
---|---|---|---|---|---|---|
Liu et al. [69] (2016) | LES | 540 m/3 m | Real terrain | 1:2000 | The forest canopy | Fluent |
Dhunny et al. [70] (2017) | Steady RANS | 825 m/- | Real terrain | 1:1 | Wind power utilization | WindSim |
Yan et al. [71] (2018) | Steady RANS | 100 m/12 m | 3D hill | 1:1 | Wind power utilization | Fluent |
Huang et al. [72] (2019) | Steady RANS | 700 m/20 m | Real terrain | 1:4000 | ABL | Fluent |
Hu et al. [73] (2021) | LES | 210 m/10 m | 3D hill/Real terrain | 1:1000 | ABL | Fluent |
Zhou et al. [74] (2022) | URANS | 600 m/3 m | 3D hill | 1:1000 | ABL | OpenFOAM |
Cao et al. [75] (2023) | LES | 200 m/10 m | 3D hill | 1:1 | ABL | WRF/OpenFOAM |
Huang et al. [76] (2024) | LES | 288 m/8 m | 2D hill/3D hill | 1:1 | Wind–snow coupling | - |
Zhou et al. [77] (2024) | LES | 60 m/2 m | 3D hill | 1:1000 | Thermally driven flow | Fluent |
Model | Typical Topography | Advantages | Disadvantages and Quantitative Limitations | Computational Cost and Feasibility |
---|---|---|---|---|
RANS | Gentle slopes, no significant flow separation | Extremely high computational efficiency | Severely underestimates the size of separation zones; unable to capture unsteady effects. | Mesh Size: 5–20 million Time Step:/(steady-state) CPU Hours: 10–100 Cost Factor: 1× |
URANS | Moderately complex terrain | Balances efficiency and accuracy; capable of capturing large-scale unsteady structures | Still relies on modeling assumptions; cannot resolve small-scale turbulence. | Mesh Size: 5–25 million Time Step: Large (e.g., 10−2–10−1 s) CPU Hours: 100–1500 Cost Factor: ~10–50× |
LES | Steep hills/cliffs/canyons | High-fidelity prediction of turbulent structures and separated flows | Computationally very expensive; sensitive to inflow boundary conditions. | Mesh Size: 50 million—500+ million Time Step: Very small (e.g., 10−5–10−3 s) CPU Hours: 10,000–100,000+ Cost Factor: ~1000× |
3.4. Data-Driven Models
Author (Year) | Sources of Data | Method | Detail | Key Quantitative Results | Mountain Feature |
---|---|---|---|---|---|
Lou et al. [90] (2020) | CFD | POD | Design of transmission tower lines | Not provided in text; focused on extracting wind field characteristics for design. | 3D valley |
Zhou et al. [91] (2020) | CFD | POD | Time–space characteristic | Not provided in text; focused on spatio-temporal analysis of wake turbulence. | 3D hill |
Wang et al. [92] (2023) | CFD | POD | Dynamic soaring | Not provided in text; used for prediction of wind shear layers. | 2D ridge |
Zhou et al. [93] (2021) | CFD | DMD | High-rise buildings | Not provided in text; focused on analyzing the flow field around buildings. | - |
Quiroga et al. [94] (2021) | Field measurement | k-NN | Wind power utilization | MAE: 1.29% | Actual complex terrain |
Lee et al. [96] (2022) | Field measurement | MLP | Wind power utilization | Mean error: 2.92 m/s | Actual complex terrain |
Sheehan et al. [97] (2022) | CFD | CNN | ABL | MAE: 0.03–0.04 m/s | Actual complex terrain |
Wold et al. [99] (2024) | Low-resolution wind field data | GAN | ABL | No error metric provided; successfully upscaled low-res data to high-res 3D fields. | Actual complex terrain |
Zhang et al. [102] (2025) | ERA5 reanalyzes the data | MFWPN | Wind speed forecast | RMSE: 0.42 m/s | Actual complex terrain |
Lin et al. [101] (2025) | CFD | U-Net | ABL | RMSE: 0.4 m/s. Reduced inference time by over three orders of magnitude (hours to <1 s) with accuracy “comparable to CFD” | Actual complex terrain |
Hidalgo et al. [100] (2025) | CFD | PINN | Rebuild the wind farm | MRMSE: 0.32 m/s | Actual complex terrain |
- AI as Surrogate/Accelerator: This mode uses deep learning models (e.g., CNNs, GANs) to learn and replace high-cost CFD solving processes. After training on massive CFD datasets, the model can achieve rapid, second-level predictions at a very low cost, supporting large-scale parameter studies and real-time forecasting.
- AI as Data Fusion & Corrector: This mode uses AI to resolve the conflict between the systematic biases of CFD models and the sparsity of field observation data. Its core is the “CFD pre-training + observation fine-tuning” strategy, where the AI model corrects CFD’s predictive biases using sparse but real observation points, generating a more accurate wind field than either source alone.
- AI as Physics-Informed Extractor: Represented by PINNs, this mode embeds governing equations like Navier–Stokes into the AI’s loss function as a physical constraint. This forces the model to generate physically self-consistent solutions even with sparse data, making it ideal for reconstructing flow fields or discovering unknown parameters from limited measurements.
- AI as Intelligent Boundary Generator: In multi-scale coupled simulations (e.g., WRF-CFD), AI acts as a bridge connecting models of different scales. By learning the complex relationship between mesoscale meteorological fields and microscale boundary conditions, the model can rapidly generate dynamic and realistic inlet conditions for CFD, thereby replacing or accelerating the slower mesoscale model and improving the efficiency of the entire simulation chain.
4. Wind Field Modeling over Hilly Terrain: Current Practical Issues
4.1. Wind Resource Development
Evaluation Method | Performance Metric | Reported Error Range | Key Influencing Factors |
---|---|---|---|
Steady RANS [86] | Mean wind speed deviation | 10–20% | Terrain complexity, turbulence model selection |
Wind direction deviation | <30° | Inflow boundary conditions | |
Lidar + CFD Correction [127] | Mean wind speed deviation | Reduced from −2.4% to −0.1% | Correction algorithm, terrain curvature |
Meso-microscale Coupling (WRF-CFD) [128] | Mean wind speed deviation | +0.05 m/s | Coupling scheme, physics parameterization |
Variability of prediction error | Reduced by 35% | Mesoscale model accuracy |
- Establish global, open benchmark datasets: Foster international collaborations to conduct large-scale field campaigns in data-sparse but representative mountainous regions (e.g., the Andes, African highlands), similar to the New European Wind Atlas (NEWA) project. The resulting high-quality data should be integrated into standardized, open-access datasets to provide a “gold standard” for researchers worldwide to validate and develop models.
- Develop physics-constrained transfer learning models: Utilize data-driven methods, particularly transfer learning, to adapt models trained in data-rich regions to target areas with sparse observational data. By incorporating physical constraints (e.g., the continuity equation), these models can ensure that predictions in data-scarce regions remain physically plausible.
- Initiate systematic model intercomparison projects: Define standardized simulation cases for different types of mountain geomorphologies and organize multiple research groups globally to perform simulations using various models. Comparing these results against high-quality observational data will help quantify model uncertainties and lead to the development of more robust parameterization schemes.
4.2. The Design of Buildings and Infrastructure
- Developing Parameterized Terrain Influence Models: Use extensive CFD and data-driven methods for parametric studies on typical mountain terrains to develop universal terrain influence prediction models or correction factors, providing a basis for revising design codes.
- Establishing a Standardized Validation Database: An international collaborative effort is needed to establish an open-access benchmark database for “Wind Loads on Infrastructure in Mountainous Terrain.” Similar to the “Askervein Hill” project, this would involve long-term monitoring of representative infrastructure to create a public validation platform for researchers to systematically evaluate and improve models.
- Promoting a Universal Framework for Multi-Hazard Coupling Research: Develop a universal numerical simulation framework capable of coupling multiple physical effects. This should be used to systematically investigate the amplification mechanisms of these coupled effects on structural loads, enabling a more comprehensive and safer resilience design for infrastructure in mountainous regions.
4.3. Pollutant Dispersion
5. Discussion and Challenges
- The Validation Chasm—A Fundamental Contradiction Between High-Fidelity Models and Sparse Validation Data
- 2.
- The Inherent Complexity of Multi-physics and Multi-scale Coupling
- 3.
- The Pervasive Lack of Data Standardization, Quality Control, and Heterogeneity Management
- 4.
- The Generalizability and Geographical Limitations of Research Findings
Core Challenge | Key Quantitative Evidence | Impact | Corresponding Research Direction |
---|---|---|---|
1. Validation Chasm | RANS overestimates leeward zone size by up to 30% [64]. LES shows RMSE up to 5.65 m/s on Perdigão ridge [67]. | Increases AEP prediction uncertainty (P50–P90), directly raising project financing risk [153]. | Establish global benchmark datasets similar to IEA Task [182]. Conduct Perdigão-level high-density observation experiments in data-sparse regions [67]. |
2. Multi-physics Coupling Complexity | Micro-topography can amplify typhoon wind speeds by over 25% [143]. Nocturnal thermal inversions “trap” pollutants [162,163]. | Current design codes (e.g., ASCE 7) underestimate loads by over 30%, posing significant safety risks [139]. | Develop multi-scale coupled models like WRF-LES. Study the impact of multi-hazard coupled effects (wind-rain-ice) on infrastructure [66]. |
3. Data-Driven Paradigm Dilemma | AI reduces inference time from hours to seconds, but training data from unvalidated CFD risks “bias amplification” [102]. | Fast, incorrect predictions are more dangerous than slow, accurate ones. | Develop Physics-Informed Neural Networks (PINNs) to integrate physical constraints into AI. Establish data quality and validation protocols as advocated by IEA Task 51 [183]. |
4. Generalizability Limitations | Most benchmark experiments are concentrated in specific regions (e.g., Europe, North America, like NEWA) [40], questioning model applicability in other climate zones (e.g., tropical, polar). | Limited global generalizability of models and parameterization schemes. | Conduct new international collaborative field observations in representative regions (e.g., Andes, African highlands) to fill “climate mechanism gaps.” |
- Integration of Multi-scale, Multi-physics Models: Efficiently couple mesoscale meteorological models (e.g., WRF) with microscale CFD simulations (especially LES) to balance computational cost and accuracy, while utilizing emerging high-resolution NWP (e.g., ICON, IFS regional configurations [184,185]) to provide higher-precision inflow boundary conditions. Furthermore, explore methods such as Physics-Informed Neural Networks (PINNs), which embed physical information as soft constraints, to generate physically plausible wind field predictions in data-sparse scenarios.
- Standardization of Data and Validation Protocols: Through international collaboration (drawing from initiatives like IEA Tasks [182,183]), conduct large-scale, multi-physics field observation campaigns in globally representative regions (e.g., the Perdigão experiment), especially by making full use of ground-based scanning LIDAR and SODAR technologies [24,25]. These remote sensing technologies are key weapons to bridge the “validation chasm,” providing 4D (3D space + time) wind field structures rather than sparse point data. Future standardized protocols must require models not only to match meteorological mast data but also to quantitatively reproduce key flow structures observed by LIDAR (e.g., recirculation region dimensions, wake morphology) and turbulence statistics (e.g., spectra and PDFs).
- Deepening and Broadening of Application Domains: Conduct in-depth research on the interaction mechanisms between extreme wind events (e.g., downbursts, typhoons) and complex terrain. Investigate the response of infrastructure to multi-physics coupling effects (wind–rain, wind–ice–snow) to enhance structural safety and energy security in mountainous regions. The ultimate future application form is the construction of “Wind Engineering Digital Twins” for mountainous areas [181]. For example, building a wind farm digital twin that fuses high-resolution NWP, live LIDAR scanning data, and sensor information in real-time, utilizing AI and reduced-order models to predict turbulence and loads for every turbine in the next hour, thereby enabling active dynamic yaw control and extreme load avoidance. Apply wind field models to assess pollutant dispersion in mountains (environmental health) and turbine noise propagation to provide a scientific basis for sustainable energy and environmental planning.
6. Conclusions
- Establishing global benchmark datasets and advancing observational technologies: Foster international collaborations to create high-quality, open access benchmark datasets from representative terrains worldwide to bridge the “validation gap.”
- Developing efficient multi-scale coupled and physics-data fusion models: Focus on developing frameworks that efficiently couple mesoscale models (e.g., WRF) with microscale LES, and explore cutting-edge methods like PINNs to generate physically plausible predictions in data-scarce scenarios.
- Deepening application-oriented research with a focus on extreme events: Enhance the investigation of interaction mechanisms between extreme wind events, such as downbursts and typhoons, and complex terrain, and apply high-fidelity wind field models to the resilience assessment of infrastructure under multi-hazard coupling effects.
Author Contributions
Funding
Institutional Review Board Statement
Informed Consent Statement
Data Availability Statement
Acknowledgments
Conflicts of Interest
References
- Körner, C. Global statistics of “mountain” and “alpine” research. Mt. Res. Dev. 2009, 29, 97–102. [Google Scholar] [CrossRef]
- Wani, A.H.; Varma, R.K.; Ahuja, A.K. Experimental investigation of wind flow characteristics over hills and escarpments-A review. Wind Struct. 2021, 32, 393–403. [Google Scholar]
- Liao, H.; Jing, H.; Ma, C.; Tao, Q.; Li, Z. Field measurement study on turbulence field by wind tower and Windcube Lidar in mountain valley. J. Wind Eng. Ind. Aerodyn. 2020, 197, 104090. [Google Scholar] [CrossRef]
- Kaimal, J.C.; Finnigan, J.J. Atmospheric Boundary Layer Flows: Their Structure and Measurement; Oxford University Press: Mumbai, India, 1994. [Google Scholar]
- Ishihara, T.; Hibi, K.; Oikawa, S. A wind tunnel study of turbulent flow over a three-dimensional steep hill. J. Wind Eng. Ind. Aerodyn. 1999, 83, 95–107. [Google Scholar] [CrossRef]
- Milla-Val, J.; Montañés, C.; Fueyo, N. Economical microscale predictions of wind over complex terrain from mesoscale simulations using machine learning. Model. Earth Syst. Environ. 2024, 10, 1407–1421. [Google Scholar] [CrossRef]
- Chen, M.; Yang, H.; Mao, B.; Xie, K.; Chen, C.; Dong, Y. An ensemble forecast wind field correction model with multiple factors and Spatio-Temporal features. Atmosphere 2023, 14, 1650. [Google Scholar] [CrossRef]
- Serafin, S.; Adler, B.; Cuxart, J.; De Wekker, S.F.; Gohm, A.; Grisogono, B.; Zardi, D. Exchange processes in the atmospheric boundary layer over mountainous terrain. Atmosphere 2018, 9, 102. [Google Scholar] [CrossRef]
- Giovannini, L.; Ferrero, E.; Karl, T.; Rotach, M.W.; Staquet, C.; Trini Castelli, S.; Zardi, D. Atmospheric pollutant dispersion over complex terrain: Challenges and needs for improving air quality measurements and modeling. Atmosphere 2020, 11, 646. [Google Scholar] [CrossRef]
- Bradley, S.; Strehz, A.; Emeis, S. Remote sensing winds in complex terrain—A review. Meteorol. Z 2015, 24, 547–555. [Google Scholar] [CrossRef]
- Finnigan, J.; Ayotte, K.; Harman, I.; Katul, G.; Oldroyd, H.; Patton, E.; Taylor, P. Boundary-layer flow over complex topography. Bound.-Layer Meteorol. 2020, 177, 247–313. [Google Scholar] [CrossRef]
- Farina, S.; Zardi, D. Understanding thermally driven slope winds: Recent advances and open questions. Bound.-Layer Meteorol. 2023, 189, 5–52. [Google Scholar] [CrossRef]
- Defant, F. Local winds. Compend. Meteorol. 1951, 655–672. [Google Scholar]
- Jackson, P.S.; Hunt, J.C.R. Turbulent wind flow over a low hill. Q. J. R. Meteorol. Soc. 1975, 101, 929–955. [Google Scholar] [CrossRef]
- Bowen, A.J.; Lindley, D. A wind-tunnel investigation of the wind speed and turbulence characteristics close to the ground over various escarpment shapes. Bound.-Layer Meteorol. 1977, 12, 259–271. [Google Scholar] [CrossRef]
- Meroney, R.N. Wind-tunnel simulation of the flow over hills and complex terrain. J. Wind Eng. Ind. Aerodyn. 1980, 5, 297–321. [Google Scholar] [CrossRef]
- Jenkins, G.J.; Mason, P.J.; Moores, W.H.; Sykes, R. Measurements of the flow structure around Ailsa Craig, a steep, three-dimensional, isolated hill. Q. J. R. Meteorol. Soc. 1981, 107, 833–851. [Google Scholar] [CrossRef]
- Mickle, R.E.; Cook, N.J.; Hoff, A.M.; Jensen, N.O.; Salmon, J.R.; Taylor, P.A.; Teunissen, H.W. The Askervein Hill Project: Vertical profiles of wind and turbulence. Bound.-Layer Meteorol. 1988, 43, 143–169. [Google Scholar] [CrossRef]
- Mouzakis, F.N.; Bergels, G.C. Numerical prediction of turbulent flow over a two-dimensional ridge. Int. J. Numer. Methods Fluids 1991, 12, 287–296. [Google Scholar] [CrossRef]
- Kim, H.G.; Lee, C.M.; Lim, H.C.; Kyong, N.H. An experimental and numerical study on the flow over two-dimensional hills. J. Wind. Eng. Ind. Aerodyn. 1997, 66, 17–33. [Google Scholar] [CrossRef]
- Pozdnoukhov, A.; Foresti, L.; Kanevski, M. Data-driven topo-climatic mapping with machine learning methods. Nat. Hazards 2009, 50, 497–518. [Google Scholar] [CrossRef]
- He, Y.C.; Chan, P.W.; Li, Q.S. Standardization of raw wind speed data under complex terrain conditions: A data-driven scheme. J. Wind Eng. Ind. Aerodyn. 2014, 131, 12–30. [Google Scholar] [CrossRef]
- Yu, C.; Li, Y.; Zhang, M.; Zhang, Y.; Zhai, G. Wind characteristics along a bridge catwalk in a deep-cutting gorge from field measurements. J. Wind Eng. Ind. Aerodyn. 2019, 186, 94–104. [Google Scholar] [CrossRef]
- Song, J.L.; Li, J.W.; Flay, R.G.J. Field measurements and wind tunnel investigation of wind characteristics at a bridge site in a Y-shaped valley. J. Wind. Eng. Ind. Aerodyn. 2020, 202, 104199. [Google Scholar] [CrossRef]
- Tsai, C.L.; Kim, K.; Liou, Y.C.; Lee, G.; Yu, C.K. Impacts of topography on airflow and precipitation in the Pyeongchang area seen from multiple-Doppler radar observations. Mon. Weather Rev. 2018, 146, 3401–3424. [Google Scholar] [CrossRef]
- Afshar-Mohajer, N.; Wu, C.Y. Use of a drone-based sensor as a field-ready technique for short-term concentration mapping of air pollutants: A modeling study. Atmos. Environ. 2023, 294, 119476. [Google Scholar] [CrossRef]
- Babić, N.; Adler, B.; Gohm, A.; Lehner, M.; Kalthoff, N. Exploring the daytime boundary layer evolution based on Doppler spectrum width from multiple coplanar wind lidars during CROSSINN. EGUsphere 2023, 2023, 1–40. [Google Scholar] [CrossRef]
- Barber, S.; Schubiger, A.; Koller, S.; Eggli, D.; Radi, A.; Rumpf, A.; Knaus, H. The wide range of factors contributing to wind resource assessment accuracy in complex terrain. Wind Energy Sci. 2022, 2022, 1503–1525. [Google Scholar] [CrossRef]
- Menke, R.; Vasiljević, N.; Wagner, J.; Oncley, S.P.; Mann, J. Multi-lidar wind resource mapping in complex terrain. Wind Energy Sci. 2020, 5, 1059–1073. [Google Scholar]
- Chen, W.; Qian, G.; Qi, W.; Luo, G.; Zhao, L.; Yuan, X. Layout Method of Met Mast Based on Macro Zoning and Micro Quantitative Siting in a Wind Farm. Processes 2022, 10, 1708. [Google Scholar] [CrossRef]
- Russell, E.S.; Liu, H.; Gao, Z.; Lamb, B.; Wagenbrenner, N. Turbulence dependence on winds and stability in a weak-wind canopy sublayer over complex terrain. J. Geophys. Res. Atmos. 2016, 121, 11502–11515. [Google Scholar] [CrossRef]
- Fenerci, A.; Øiseth, O.; Rønnquist, A. Long-term monitoring of wind field characteristics and dynamic response of a long-span suspension bridge in complex terrain. Eng. Struct. 2017, 147, 269–284. [Google Scholar] [CrossRef]
- Chaurasiya, P.K.; Ahmed, S.; Warudkar, V. Comparative analysis of Weibull parameters for wind data measured from met-mast and remote sensing techniques. Renew. Energy 2018, 115, 1153–1165. [Google Scholar] [CrossRef]
- Huang, G.; Jiang, Y.; Peng, L.; Solari, G.; Liao, H.; Li, M. Characteristics of intense winds in mountain area based on field measurement: Focusing on thunderstorm winds. J. Wind Eng. Ind. Aerodyn. 2019, 190, 166–182. [Google Scholar] [CrossRef]
- Zhang, J.; Zhang, M.; Li, Y.; Qin, J.; Wei, K.; Song, L. Analysis of wind characteristics and wind energy potential in complex mountainous region in southwest China. J. Clean. Prod. 2020, 274, 123036. [Google Scholar] [CrossRef]
- Radünz, W.C.; Sakagami, Y.; Haas, R.; Petry, A.P.; Passos, J.C.; Miqueletti, M.; Dias, E. Influence of atmospheric stability on wind farm performance in complex terrain. Appl. Energy 2021, 282, 116149. [Google Scholar] [CrossRef]
- Radünz, W.C.; de Almeida, E.; Gutiérrez, A.; Acevedo, O.C.; Sakagami, Y.; Petry, A.P.; Passos, J.C. Nocturnal jets over wind farms in complex terrain. Appl. Energy 2022, 314, 118959. [Google Scholar] [CrossRef]
- Jiang, F.; Zhang, J.; Zhang, M.; Li, Y.; Qin, J. Field measurement study on classification for mixed intense wind climate in mountainous terrain. Measurement 2023, 217, 113064. [Google Scholar] [CrossRef]
- Adler, B.; Gohm, A.; Kalthoff, N.; Babić, N.; Corsmeier, U.; Lehner, M.; Georgoussis, G. CROSSINN: A field experiment to study the three-dimensional flow structure in the Inn Valley, Austria. Bull. Am. Meteorol. Soc. 2021, 102, E38–E60. [Google Scholar] [CrossRef]
- Coimbra, I.L.; Mann, J.; Palma, J.M.L.M.; Batista, V.T. Exploring dual-lidar mean and turbulence measurements over Perdigão’s complex terrain. Atmos. Meas. Tech. 2025, 18, 287–303. [Google Scholar] [CrossRef]
- Desnijder, K.; Ali, A.S.; Vandewal, M. Real-time atmospheric characterization using drone swarms as sensor platforms. In Proceedings of the Environmental Effects on Light Propagation and Adaptive Systems VII, Edinburgh, UK, 18–19 September 2024; SPIE: Bellingham, WA, USA, 2024; Volume 13194, p. 1319402. [Google Scholar]
- Utnes, T.; Eidsvik, K.J. Turbulent flows over mountainous terrain modelled by the Reynolds equations. Bound.-Layer Meteorol. 1996, 79, 393–416. [Google Scholar] [CrossRef]
- Hancock, P.E.; Hayden, P. Wind-tunnel simulation of weakly and moderately stable atmospheric boundary layers. Bound.-Layer Meteorol. 2018, 168, 29–57. [Google Scholar] [CrossRef]
- Demarco, G.; Martins, L.G.N.; Bodmann, B.E.J.; Puhales, F.S.; Acevedo, O.C.; Wittwer, A.R.; Degrazia, G.A. Analysis of thermal and roughness effects on the turbulent characteristics of experimentally simulated boundary layers in a wind tunnel. Int. J. Environ. Res. Public Health 2022, 19, 5134. [Google Scholar] [CrossRef]
- Pospíšil, S.; Kuznetsov, S.; Kozmar, H.; Michalcová, V. Wind-tunnel simulation of thermally unstable atmospheric flow in complex terrain. Procedia Eng. 2017, 190, 575–580. [Google Scholar] [CrossRef]
- Zou, Y.; Yue, P.; Liu, Q.; He, X.; Wang, Z. Wind field characteristics of complex terrain based on experimental and numerical investigation. Appl. Sci. 2022, 12, 5124. [Google Scholar] [CrossRef]
- Mattuella, J.M.L.; Loredo-Souza, A.M.; Oliveira, M.G.K.; Petry, A.P. Wind tunnel experimental analysis of a complex terrain micrositing. Renew. Sustain. Energy Rev. 2016, 54, 110–119. [Google Scholar] [CrossRef]
- Li, Y.; Hu, P.; Xu, X.; Qiu, J. Wind characteristics at bridge site in a deep-cutting gorge by wind tunnel test. J. Wind Eng. Ind. Aerodyn. 2017, 160, 30–46. [Google Scholar] [CrossRef]
- Hyvärinen, A.; Lacagnina, G.; Segalini, A. A wind-tunnel study of the wake development behind wind turbines over sinusoidal hills. Wind. Energy 2018, 21, 605–617. [Google Scholar] [CrossRef]
- Ikegaya, N.; Kikumoto, H.; Sasaki, K.; Yamada, S.; Matsui, M. Applications of wide-ranging PIV measurements for various turbulent statistics in artificial atmospheric turbulent flow in a wind tunnel. Build. Environ. 2022, 225, 109590. [Google Scholar] [CrossRef]
- Stidworthy, A.; Carruthers, D. FLOWSTAR-Energy: A high resolution wind farm wake model. Wind Energy Sci. 2016, 2016, 1–24. [Google Scholar]
- Ma, G.; Tian, L.; Song, Y.; Zhao, N. Effects of Turbulence Modeling on the Simulation of Wind Flow over Typical Complex Terrains. Appl. Sci. 2024, 14, 11438. [Google Scholar] [CrossRef]
- Lystad, T.M.; Fenerci, A.; Øiseth, O. Evaluation of mast measurements and wind tunnel terrain models to describe spatially variable wind field characteristics for long-span bridge design. J. Wind Eng. Ind. Aerodyn. 2018, 179, 558–573. [Google Scholar] [CrossRef]
- Tian, W.; Ozbay, A.; Hu, H. An experimental investigation on the aeromechanics and wake interferences of wind turbines sited over complex terrain. J. Wind Eng. Ind. Aerodyn. 2018, 172, 379–394. [Google Scholar] [CrossRef]
- Kamada, Y.; Li, Q.; Maeda, T.; Yamada, K. Wind tunnel experimental investigation of flow field around two-dimensional single hill models. Renew. Energy 2019, 136, 1107–1118. [Google Scholar] [CrossRef]
- Shen, Z.; Li, J.; Li, R.; Gao, G. Nonuniform wind characteristics and buffeting response of a composite cable-stayed bridge in a trumpet-shaped mountain pass. J. Wind Eng. Ind. Aerodyn. 2021, 217, 104730. [Google Scholar] [CrossRef]
- Zhu, S.; Li, Y.; Xu, X. Wind tunnel test on the aerodynamic admittance of a rail vehicle in crosswinds. J. Wind Eng. Ind. Aerodyn. 2022, 226, 105052. [Google Scholar] [CrossRef]
- Raffaele, L.; Glabeke, G.; van Beeck, J. Wind-sand tunnel experiment on the windblown sand transport and sedimentation over a two-dimensional sinusoidal hill. Wind Struct. 2023, 36, 75–90. [Google Scholar]
- Wu, S.; He, J.; Gu, W.; Chen, D.; Nie, B.; Wang, D. Modification to aerosols dispersion algorithms over two-dimensional hilly terrain based on wind tunnel atmospheric experiments. Ann. Nucl. Energy 2025, 213, 111165. [Google Scholar] [CrossRef]
- Uchida, T.; Li, G. Comparison of RANS and LES in the prediction of airflow field over steep complex terrain. Open J. Fluid Dyn. 2018, 8, 286–307. [Google Scholar] [CrossRef]
- Han, Y.; Stoellinger, M.K.; Peng, H.; Zhang, L.; Liu, W. Large eddy simulation of atmospheric boundary layer flow over complex terrain in comparison with RANS simulation and on-site measurements under neutral stability condition. J. Renew. Sustain. Energy 2023, 15, 023301. [Google Scholar] [CrossRef]
- Wenz, F.; Langner, J.; Lutz, T.; Krämer, E. Impact of the wind field at the complex-terrain site Perdigão on the surface pressure fluctuations of a wind turbine. Wind Energy Sci. 2022, 7, 1321–1340. [Google Scholar] [CrossRef]
- Golaz, J.C.; Doyle, J.D.; Wang, S. One-way nested large-eddy simulation over the Askervein Hill. J. Adv. Model. Earth Syst. 2009, 1. [Google Scholar] [CrossRef]
- Temel, O.; Bricteux, L.; van Beeck, J. Coupled WRF-OpenFOAM study of wind flow over complex terrain. J. Wind Eng. Ind. Aerodyn. 2018, 174, 152–169. [Google Scholar] [CrossRef]
- Bao, J.; Chow, F.K.; Lundquist, K.A. Large-eddy simulation over complex terrain using an improved immersed boundary method in the Weather Research and Forecasting Model. Mon. Weather Rev. 2018, 146, 2781–2797. [Google Scholar] [CrossRef]
- Fernando, H.J.S.; Mann, J.; Palma, J.; Lundquist, J.K.; Barthelmie, R.J.; Belo-Pereira, M.; Wang, Y. The Perdigão: Peering into microscale details of mountain winds. Bull. Am. Meteorol. Soc. 2019, 100, 799–819. [Google Scholar] [CrossRef]
- Al Oqaily, D.; Giani, P.; Crippa, P. Evaluating WRF multiscale wind simulations in complex terrain: Insights from the Perdigão field campaign. J. Geophys. Res. Atmos. 2025, 130, e2025JD044055. [Google Scholar] [CrossRef]
- Hammer, F.; Barber, S.; Remmler, S.; Bernardoni, F.; Venkatraman, K.; Díez Sánchez, G.A.; Giljarhus, K.E. Comparison metrics microscale simulation challenge for wind resource assessment. Wind Energy Sci. 2023, 2023, 1–27. [Google Scholar]
- Liu, Z.; Ishihara, T.; He, X.; Niu, H. LES study on the turbulent flow fields over complex terrain covered by vegetation canopy. J. Wind Eng. Ind. Aerodyn. 2016, 155, 60–73. [Google Scholar] [CrossRef]
- Dhunny, A.Z.; Lollchund, M.R.; Rughooputh, S. Wind energy evaluation for a highly complex terrain using Computational Fluid Dynamics (CFD). Renew. Energy 2017, 101, 1–9. [Google Scholar] [CrossRef]
- Yan, S.; Shi, S.; Chen, X.; Wang, X.; Mao, L.; Liu, X. Numerical simulations of flow interactions between steep hill terrain and large scale wind turbine. Energy 2018, 151, 740–747. [Google Scholar] [CrossRef]
- Huang, W.; Zhang, X. Wind field simulation over complex terrain under different inflow wind directions. Wind Struct. 2019, 28, 239–253. [Google Scholar]
- Hu, W.; Yang, Q.; Chen, H.P.; Yuan, Z.; Li, C.; Shao, S.; Zhang, J. Wind field characteristics over hilly and complex terrain in turbulent boundary layers. Energy 2021, 224, 120070. [Google Scholar] [CrossRef]
- Zhou, T.; Yang, Q.; Yan, B.; Deng, X.; Yuan, Y. Detached eddy simulation of turbulent flow fields over steep hilly terrain. J. Wind Eng. Ind. Aerodyn. 2022, 221, 104906. [Google Scholar] [CrossRef]
- Cao, Y.; Tao, T.; Shi, Y.; Cao, S.; Zhou, D.; Chen, W.L. Large-eddy simulation of separated turbulent flows over a three-dimensional hill using WRF and OpenFOAM. J. Wind Eng. Ind. Aerodyn. 2023, 236, 105357. [Google Scholar] [CrossRef]
- Huang, N.; Yu, Y.; Shao, Y.; Zhang, J. Numerical Simulation of Falling-Snow Deposition Pattern Over 3D-Hill. J. Geophys. Res. Atmos. 2024, 129, e2023JD039898. [Google Scholar] [CrossRef]
- Zhou, T.; Ishihara, T. LES study of turbulent flow fields over a three-dimensional steep hill considering the effects of thermal stratification. Comput. Fluids 2025, 288, 106521. [Google Scholar] [CrossRef]
- Chen, F.; Wang, W.; Gu, Z.; Zhu, Y.; Li, Y.; Shu, Z. Investigation of hilly terrain wind characteristics considering the interference effect. J. Wind Eng. Ind. Aerodyn. 2023, 241, 105543. [Google Scholar] [CrossRef]
- Solbakken, K.; Birkelund, Y.; Samuelsen, E.M. Evaluation of surface wind using WRF in complex terrain: Atmospheric input data and grid spacing. Environ. Model. Softw. 2021, 145, 105182. [Google Scholar] [CrossRef]
- Vuorinen, V.; Chaudhari, A.; Keskinen, J.P. Large-eddy simulation in a complex hill terrain enabled by a compact fractional step OpenFOAM® solver. Adv. Eng. Softw. 2015, 79, 70–80. [Google Scholar] [CrossRef]
- Ricci, A. Review of OpenFOAM applications in the computational wind engineering: From wind environment to wind structural engineering. Meccanica 2024, 60, 1695–1735. [Google Scholar] [CrossRef]
- Onel, H.C.; Tuncer, I.H. Short-Term Numerical Forecasting of Near-Ground Wind Fields Using OpenFOAM Coupled With WRF. In Proceedings of the AIAA SCITECH 2023 Forum, Online, 23–27 January 2023; p. 1737. [Google Scholar]
- Durán, P.; Meiβner, C.; Casso, P. A new meso-microscale coupled modelling framework for wind resource assessment: A validation study. Renew. Energy 2020, 160, 538–554. [Google Scholar] [CrossRef]
- Cheng, X.; Yan, B.; Zhou, X.; Yang, Q.; Huang, G.; Su, Y.; Jiang, Y. Wind resource assessment at mountainous wind farm: Fusion of RANS and vertical multi-point on-site measured wind field data. Appl. Energy 2024, 363, 123116. [Google Scholar] [CrossRef]
- Zhou, Q.; Zhu, Y.; Wang, Y.; Han, J. CFD-based wind field correction method for terrain wind tunnel tests. J. Phys. Conf. Ser. 2021, 2083, 032083. [Google Scholar] [CrossRef]
- Blocken, B.; van der Hout, A.; Dekker, J.; Weiler, O. CFD simulation of wind flow over natural complex terrain: Case study with validation by field measurements for Ria de Ferrol, Galicia, Spain. J. Wind Eng. Ind. Aerodyn. 2015, 147, 43–57. [Google Scholar] [CrossRef]
- Dujardin, J.; Lehning, M. Wind-Topo: Downscaling near-surface wind fields to high-resolution topography in highly complex terrain with deep learning. Q. J. R. Meteorol. Soc. 2022, 148, 1368–1388. [Google Scholar] [CrossRef]
- Premaratne, P.; Tian, W.; Hu, H. A proper-orthogonal-decomposition (POD) study of the wake characteristics behind a wind turbine model. Energies 2022, 15, 3596. [Google Scholar] [CrossRef]
- Hu, W.; Yang, Q.; Zhou, T.; Lu, B.; Qian, G.; Shan, W.; Wang, Y. Turbulent wake dynamics and flow characteristics over typical hilly terrain: A proper orthogonal decomposition and dynamic mode decomposition analysis. Phys. Fluids 2025, 37, 055120. [Google Scholar] [CrossRef]
- Lou, W.; Bai, H.; Huang, M.; Duan, Z.; Bian, R. Wind field generation for performance-based structural design of transmission lines in a mountainous area. Wind Struct. 2020, 31, 165–183. [Google Scholar]
- Zhou, T.; Yan, B.; Yang, Q.; Hu, W.; Chen, F. POD analysis of spatiotemporal characteristics of wake turbulence over hilly terrain and their relationship to hill slope, hill shape and inflow turbulence. J. Wind Eng. Ind. Aerodyn. 2022, 224, 104986. [Google Scholar] [CrossRef]
- Wang, D.; Xie, F.; Ji, T.; Zhang, X.; Lu, Y.; Zheng, Y. Prediction of wind shear layer for dynamic soaring by using proper orthogonal decomposition and long short term memory network. Phys. Fluids 2023, 35, 085103. [Google Scholar] [CrossRef]
- Zhou, L.; Hu, G.; Tse, K.T.; He, X. Twisted-wind effect on the flow field of tall building. J. Wind Eng. Ind. Aerodyn. 2021, 218, 104778. [Google Scholar] [CrossRef]
- Quiroga-Novoa, P.; Cuevas-Figueroa, G.; Preciado, J.L.; Floors, R.; Peña, A.; Probst, O. Towards better wind resource modeling in complex terrain: A k-nearest neighbors approach. Energies 2021, 14, 4364. [Google Scholar] [CrossRef]
- Qiao, D.; Wu, S.; Li, G.; You, J.; Zhang, J.; Shen, B. Wind speed forecasting using multi-site collaborative deep learning for complex terrain application in valleys. Renew. Energy 2022, 189, 231–244. [Google Scholar] [CrossRef]
- Lee, D.; Jeong, S.Y.; Kang, T.H.K. Consideration of terrain features from satellite imagery in machine learning of basic wind speed. Build. Environ. 2022, 213, 108866. [Google Scholar] [CrossRef]
- Sheehan, H.; Traiger, E.; Poole, D.; Landberg, L. Predicting linearised wind resource grids using neural networks. J. Wind Eng. Ind. Aerodyn. 2022, 229, 105123. [Google Scholar] [CrossRef]
- Fu, X.; Gao, F.; Wu, J.; Wei, X.; Duan, F. Spatiotemporal attention networks for wind power forecasting. In Proceedings of the 2019 International Conference on Data Mining Workshops (ICDMW), Beijing, China, 8–11 November 2019; IEEE: New York, NY, USA, 2019; pp. 149–154. [Google Scholar]
- Wold, J.W.; Stadtmann, F.; Rasheed, A.; Tabib, M.; San, O.; Horn, J.T. Enhancing wind field resolution in complex terrain through a knowledge-driven machine learning approach. Eng. Appl. Artif. Intell. 2024, 137, 109167. [Google Scholar] [CrossRef]
- Cobelli, P.; Shukla, K.; Nesmachnow, S.; Draper, M. Physics informed neural networks for wind field modeling in wind farms. J. Phys. Conf. Ser. 2023, 2505, 012051. [Google Scholar] [CrossRef]
- Lin, C.; Tie, R.; Yi, S.; Zhong, X.; Li, H. Terrain-aware Deep Learning for Wind Energy Applications: From Kilometer-scale Forecasts to Fine Wind Fields. arXiv 2025, arXiv:2505.12732. [Google Scholar]
- Zhang, Z.; Lin, L.; Gao, S.; Wang, J.; Zhao, H.; Yu, H. A machine learning model for hub-height short-term wind speed prediction. Nat. Commun. 2025, 16, 3195. [Google Scholar] [CrossRef]
- Lu, Q.; Cao, Y.; Xie, P.; Chen, Y.; Lin, Y. A Scalable Data-Driven Surrogate Model for 3D Dynamic Wind Farm Wake Prediction Using Physics-Inspired Neural Networks and Wind Box Decomposition. Energies 2025, 18, 3356. [Google Scholar] [CrossRef]
- Liu, Y.; Wang, R.; Gu, Y.; Li, C.; Wang, G. Physics-inspired and data-driven two-stage deep learning approach for wind field reconstruction with experimental validation. Energy 2024, 298, 131230. [Google Scholar] [CrossRef]
- Pereira, R.; Guedes, R.; Santos, S. Comparing WAsP and CFD wind resource estimates for the regular user. In Proceedings of the European Wind Energy Conference, Warsaw, Poland, 20–23 April 2010. [Google Scholar]
- Piotrowski, P.; Rutyna, I.; Baczyński, D.; Kopyt, M. Evaluation metrics for wind power forecasts: A comprehensive review and statistical analysis of errors. Energies 2022, 15, 9657. [Google Scholar] [CrossRef]
- Abdelsattar, M.; Ismeil, M.A.; Menoufi, K.; AbdelMoety, A.; Emad-Eldeen, A. Evaluating Machine Learning and Deep Learning models for predicting Wind Turbine power output from environmental factors. PLoS ONE 2025, 20, e0317619. [Google Scholar] [CrossRef]
- Badarinath, K.; Hoebeke, P.; Schillebeeckx, D.; Yazicioglu, H. Data-driven wind turbine sensor health validation. J. Phys. Conf. Ser. 2024, 2767, 032032. [Google Scholar] [CrossRef]
- Somashekar, V.; Selwynraj, I. Numerical study of unmanned aerial vehicle (UAV) in dry microburst environment using large eddy simulation (LES) model. Aircr. Eng. Aerosp. Technol. 2022, 94, 1170–1179. [Google Scholar] [CrossRef]
- Spence, S.M.J.; Lombardo, F.T.; McCormick, J.P. Recent Advances in the Modeling of Wind Hazards and Performance Assessment of Wind-Excited Systems. J. Struct. Eng. 2022, 148, 02021003. [Google Scholar] [CrossRef]
- Errigo, I.M.; Abbott, B.W.; Mendoza, D.L.; Mitchell, L.; Sayedi, S.S.; Glenn, J.; Wilson, D. Human health and economic costs of air pollution in Utah: An expert assessment. Atmosphere 2020, 11, 1238. [Google Scholar] [CrossRef]
- Clifton, A.; Barber, S.; Stökl, A.; Frank, H.; Karlsson, T. Research challenges and needs for the deployment of wind energy in hilly and mountainous regions. Wind Energy Sci. 2022, 7, 2231–2254. [Google Scholar] [CrossRef]
- Zhou, H.; Luo, Q.; Yuan, L. Downscaling and wind resource assessment of climatic wind speed data based on deep learning: A case study of the Tengger Desert wind farm. Atmosphere 2024, 15, 271. [Google Scholar] [CrossRef]
- Antonini, E.G.A.; Romero, D.A.; Amon, C.H. Optimal design of wind farms in complex terrains using computational fluid dynamics and adjoint methods. Appl. Energy 2020, 261, 114426. [Google Scholar] [CrossRef]
- Calautit, K.; Aquino, A.; Calautit, J.K.; Nejat, P.; Jomehzadeh, F.; Hughes, B.R. A review of numerical modelling of multi-scale wind turbines and their environment. Computation 2018, 6, 24. [Google Scholar] [CrossRef]
- Sanz Rodrigo, J.; Chavez Arroyo, R.A.; Moriarty, P.; Churchfield, M.; Kosović, B.; Réthoré, P.E.; Rife, D. Mesoscale to microscale wind farm flow modeling and evaluation. Wiley Interdiscip. Rev. Energy Environ. 2017, 6, e214. [Google Scholar] [CrossRef]
- Bodini, N.; Optis, M. Operational-based annual energy production uncertainty: Are its components actually uncorrelated? Wind. Energy Sci. 2020, 5, 1435–1448. [Google Scholar] [CrossRef]
- Zhang, Z.; Huang, P.; Bitsuamlak, G.; Cao, S. Large-eddy simulation of wind-turbine wakes over two-dimensional hills. Phys. Fluids 2022, 34, 065123. [Google Scholar] [CrossRef]
- Chen, Y.; Yan, B.; Yu, M.; Huang, G.; Qian, G.; Yang, Q.; Mo, R. Wind tunnel study of wind turbine wake characteristics over two-dimensional hill considering the effects of terrain slope and turbine position. Appl. Energy 2025, 380, 125044. [Google Scholar] [CrossRef]
- Lee, J.C.; Fields, M.J. An overview of wind energy production prediction bias, losses, and uncertainties. Wind Energy Sci. 2020, 2020, 1–82. [Google Scholar] [CrossRef]
- Cruz, L.E.B.; Carmo, B.S. Wind farm layout optimization based on CFD simulations. J. Braz. Soc. Mech. Sci. Eng. 2020, 42, 433. [Google Scholar] [CrossRef]
- Song, M.; Chen, K.; Zhang, X.; Wang, J. Optimization of wind turbine micro-siting for reducing the sensitivity of power generation to wind direction. Renew. Energy 2016, 85, 57–65. [Google Scholar] [CrossRef]
- Han, X.; Liu, D.; Xu, C.; Shen, W.Z. Atmospheric stability and topography effects on wind turbine performance and wake properties in complex terrain. Renew. Energy 2018, 126, 640–651. [Google Scholar] [CrossRef]
- de Sá Sarmiento, F.I.P.; Oliveira, J.L.G.; Passos, J.C. Impact of atmospheric stability, wake effect and topography on power production at complex-terrain wind farm. Energy 2022, 239, 122211. [Google Scholar] [CrossRef]
- Radünz, W.C.; Sakagami, Y.; Haas, R.; Petry, A.P.; Passos, J.C.; Miqueletti, M.; Dias, E. The variability of wind resources in complex terrain and its relationship with atmospheric stability. Energy Convers. Manag. 2020, 222, 113249. [Google Scholar] [CrossRef]
- Dhunny, A.Z.; Lollchund, M.R.; Rughooputh, S. Numerical analysis of wind flow patterns over complex hilly terrains: Comparison between two commonly used CFD software. Int. J. Glob. Energy Issues 2016, 39, 181–203. [Google Scholar] [CrossRef]
- Tang, X.Y.; Zhao, S.; Fan, B.; Peinke, J.; Stoevesandt, B. Micro-scale wind resource assessment in complex terrain based on CFD coupled measurement from multiple masts. Appl. Energy 2019, 238, 806–815. [Google Scholar] [CrossRef]
- Keck, R.E.; Sondell, N. Validation of uncertainty reduction by using multiple transfer locations for WRF-CFD coupling in numerical wind energy assessments. Wind Energy Sci. 2020, 2020, 997–1005. [Google Scholar] [CrossRef]
- Mann, J.; Angelou, N.; Arnqvist, J.; Callies, D.; Cantero, E.; Arroyo, R.C.; Rodrigues, C.V. Complex terrain experiments in the new european wind atlas. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2017, 375, 20160101. [Google Scholar] [CrossRef]
- de Souza, N.B.P.; Nascimento, E.G.S.; Santos, A.A.B.; Moreira, D.M. Wind mapping using the mesoscale WRF model in a tropical region of Brazil. Energy 2022, 240, 122491. [Google Scholar] [CrossRef]
- Shen, W.Z.; Zhu, W.J.; Barlas, E.; Li, Y. Advanced flow and noise simulation method for wind farm assessment in complex terrain. Renew. Energy 2019, 143, 1812–1825. [Google Scholar] [CrossRef]
- Ma, B.; Yang, J.; Chen, X.; Zhang, L.; Zeng, W. Revealing the ecological impact of low-speed mountain wind power on vegetation and soil erosion in South China: A case study of a typical wind farm in Yunnan. J. Clean. Prod. 2023, 419, 138020. [Google Scholar] [CrossRef]
- Martínez-Martínez, Y.; Dewulf, J.; Aguayo, M.; Casas-Ledón, Y. Sustainable wind energy planning through ecosystem service impact valuation and exergy: A study case in south-central Chile. Renew. Sustain. Energy Rev. 2023, 178, 113252. [Google Scholar] [CrossRef]
- Comanducci, G.; Ubertini, F.; Materazzi, A.L. Structural health monitoring of suspension bridges with features affected by changing wind speed. J. Wind Eng. Ind. Aerodyn. 2015, 141, 12–26. [Google Scholar] [CrossRef]
- Yu, E.; Xu, G.; Han, Y.; Hu, P.; Townsend, J.F.; Li, Y. Bridge vibration under complex wind field and corresponding measurements: A review. J. Traffic Transp. Eng. (Engl. Ed.) 2022, 9, 339–362. [Google Scholar] [CrossRef]
- He, P.; Yu, C.; Li, Y.; Chen, X.; Wei, Z. Validation of a novel transition curve for simulating wind fields in complex terrain using field measurements. J. Wind Eng. Ind. Aerodyn. 2025, 257, 105967. [Google Scholar] [CrossRef]
- Yu, J.; Chen, X.; Li, M.; Li, M. Application of numerical methods in the estimation of design wind velocity for bridges in mountainous areas. J. Wind Eng. Ind. Aerodyn. 2024, 250, 105764. [Google Scholar] [CrossRef]
- An, L.S.; Alinejad, N.; Jung, S. Experimental study on the influence of terrain complexity on wind pressure characteristics of mid-rise buildings. Eng. Struct. 2024, 321, 118907. [Google Scholar] [CrossRef]
- Kim, S.; Alinejad, N.; Jung, S.; Kim, H.K. The effect of open-to-suburban terrain transition on wind pressures on a low-rise building. J. Build. Eng. 2024, 85, 108651. [Google Scholar] [CrossRef]
- Alinejad, N.; Kim, S.; Jung, S. Wind-tunnel testing of low-and midrise buildings under heterogeneous upwind terrains. J. Struct. Eng. 2024, 150, 04724001. [Google Scholar] [CrossRef]
- Yan, B.; Yuan, Y.; Ma, C.; Dong, Z.; Huang, H.; Wang, Z. Modeling of downburst outflows and wind pressures on a high-rise building under different terrain conditions. J. Build. Eng. 2022, 48, 103738. [Google Scholar] [CrossRef]
- Jiang, F.; Zhang, M.; Li, Y.; Zhang, J.; Qin, J.; Wu, L. Field measurement study of wind characteristics in mountain terrain: Focusing on sudden intense winds. J. Wind Eng. Ind. Aerodyn. 2021, 218, 104781. [Google Scholar] [CrossRef]
- Meng, X.; Tian, L.; Ma, R.; Zhang, L.; Liu, J.; Dong, X. Typhoon-induced failure analysis of electricity transmission tower-line system incorporating microtopography. Eng. Fail. Anal. 2024, 163, 108556. [Google Scholar] [CrossRef]
- Zhou, C.; Yin, J.; Liu, Y. Effects of wind and rain on the motion of the high-voltage conductor in a simplified valley terrain. Electr. Power Syst. Res. 2019, 173, 153–163. [Google Scholar] [CrossRef]
- Chen, Y.; Cai, Y.; Xie, Q.; Wan, J.; Sun, Q. Combined effects of icing and wind on transmission lines in mountainous areas. J. Constr. Steel Res. 2024, 223, 109042. [Google Scholar] [CrossRef]
- Zhang, M.; Jiang, F.; Li, Y.; Chen, H.; Qin, J.; Wu, L. Multi-point field measurement study of wind characteristics in mountain terrain: Focusing on periodic thermally-developed winds. J. Wind Eng. Ind. Aerodyn. 2022, 228, 105102. [Google Scholar] [CrossRef]
- Li, Y.; Jiang, F.; Zhang, M.; Dai, Y.; Qin, J.; Zhang, J. Observations of periodic thermally-developed winds beside a bridge region in mountain terrain based on field measurement. J. Wind Eng. Ind. Aerodyn. 2022, 225, 104996. [Google Scholar] [CrossRef]
- Hu, P.; Chen, Y.; Han, Y.; Zhang, F.; Tang, Y. Numerical simulation of characteristics of wind field at bridge sites in flat and gorge terrains under the thunderstorm downburst. Shock. Vib. 2021, 2021, 9962519. [Google Scholar] [CrossRef]
- Abd-Elaal, E.S.; Mills, J.E.; Ma, X. Numerical simulation of downburst wind flow over real topography. J. Wind Eng. Ind. Aerodyn. 2018, 172, 85–95. [Google Scholar] [CrossRef]
- Li, W.; Cui, S.; Zhao, J.; An, L.; Yu, C.; Ding, Y.; Liu, Q. Experimental Study of Wind Characteristics at a Bridge Site in Mountain Valley Considering the Effect of Oncoming Wind Speed. Appl. Sci. 2024, 14, 10588. [Google Scholar] [CrossRef]
- Letson, F.; Barthelmie, R.J.; Hu, W.; Pryor, S.C. Characterizing wind gusts in complex terrain. Atmos. Chem. Phys. 2019, 19, 3797–3819. [Google Scholar] [CrossRef]
- Zhang, Y.; Li, Q.; Cao, S.; Cao, J. Embedded large eddy simulation of typhoon wind field and its effects in a large-scale complex urban area with field validations. Phys. Fluids 2025, 37, 087103. [Google Scholar] [CrossRef]
- Smith, A.B.; Katz, R.W. US Billion-Dollar Weather and Climate Disasters: Data Sources, Trends, Accuracy and Biases. Nat. Hazards 2013, 67, 387–410. [Google Scholar] [CrossRef]
- Gracia, M.A.L. Supply Chain Disruption After Collapse of Bridge at the Port of Baltimore. In Proceedings of the 2024 9th International Engineering, Sciences and Technology Conference (IESTEC), Panama City, Panama, 23–25 October 2024; IEEE: New York, NY, USA, 2024. [Google Scholar]
- Pack, M. Data Prepared: Insights and Impacts from the Francis Scott Key Bridge Collapse. ITE J. Inst. Transp. Eng. 2024, 94, 24–36. [Google Scholar]
- Ghasemian, M.; Amini, S.; Princevac, M. The influence of roadside solid and vegetation barriers on near-road air quality. Atmos. Environ. 2017, 170, 108–117. [Google Scholar] [CrossRef]
- Sabatier, T.; Paci, A.; Lac, C.; Canut, G.; Largeron, Y.; Masson, V. Semi-idealized simulations of wintertime flows and pollutant transport in an Alpine valley: Origins of local circulations (Part I). Q. J. R. Meteorol. Soc. 2020, 146, 807–826. [Google Scholar] [CrossRef]
- Lai, H.C.; Lin, M.C. Characteristics of the upstream flow patterns during PM2. 5 pollution events over a complex island topography. Atmos. Environ. 2020, 227, 117418. [Google Scholar] [CrossRef]
- Wharton, S.; Brown, M.J.; Dexheimer, D.; Fast, J.D.; Newsom, R.K.; Schalk, W.W.; Wiersema, D.J. Capturing plume behavior in complex terrain: An overview of the Nevada National Security Site Meteorological Experiment (METEX21). Front. Earth Sci. 2023, 11, 1251153. [Google Scholar] [CrossRef]
- Szmelter, A.; Szmelter, J. A CFD Study of Pollution Dispersion in a Historic Ventilation Corridor with an Evolving Urban Complex. Sustainability 2025, 17, 7348. [Google Scholar] [CrossRef]
- Guo, W.; Yang, Y.; Zhang, J.; Han, K.; Yang, Y.; Chen, Q.; Zhu, Y. Effects of valley topography on ozone pollution in the Lanzhou valley: A numerical case study. Environ. Pollut. 2024, 363, 125225. [Google Scholar] [CrossRef] [PubMed]
- Wei, N.; Wang, N.; Huang, X.; Liu, P.; Chen, L. The effects of terrain and atmospheric dynamics on cold season heavy haze in the Guanzhong Basin of China. Atmos. Pollut. Res. 2020, 11, 1805–1819. [Google Scholar] [CrossRef]
- Aravind, A.; Srinivas, C.V.; Shrivastava, R.; Hegde, M.N.; Seshadri, H.; Mohapatra, D.K. Simulation of atmospheric flow field over the complex terrain of Kaiga using WRF: Sensitivity to model resolution and PBL physics. Meteorol. Atmos. Phys. 2022, 134, 13. [Google Scholar] [CrossRef]
- Zhang, L.; Guo, X.; Zhao, T.; Gong, S.; Xu, X.; Li, Y.; Yin, X. A modelling study of the terrain effects on haze pollution in the Sichuan Basin. Atmos. Environ. 2019, 196, 77–85. [Google Scholar] [CrossRef]
- Wu, L.; Xie, B.; Wang, W. Quantifying the impact of terrain–wind–governed close-effect on atmospheric polluted concentrations. J. Clean. Prod. 2022, 367, 132995. [Google Scholar] [CrossRef]
- Ahmad, F.; Majumder, D.; Ranjit, R.; Gupta, A.; Manhart, M. Preliminary study on the spread of air-borne pollutants in urban environment: A CFD simulation approach. Sci. Rep. 2025, 15, 18836. [Google Scholar] [CrossRef]
- Nimmatoori, P.; Kumar, A. Dispersion Modeling of Particulate Matter in Different Size Ranges Releasing from a Biosolids Applied Agricultural Field Using Computational Fluid Dynamics. Adv. Chem. Eng. Sci. 2021, 11, 180. [Google Scholar] [CrossRef]
- Brusca, S.; Famoso, F.; Lanzafame, R.; Mauro, S.; Messina, M.; Strano, S. Pm10 dispersion modeling by means of cfd 3d and Eulerian–Lagrangian models: Analysis and comparison with experiments. Energy Procedia 2016, 101, 329–336. [Google Scholar] [CrossRef]
- Ma, X.; Zhong, W. CFD-DPM simulation on the atmospheric pollutant dispersion in industrial park. Atmosphere 2024, 15, 298. [Google Scholar] [CrossRef]
- He, J.; Kang, Y.; Wang, Y.; Gu, Y.; Zhong, K. Effects of sea-land breeze on air pollutant dispersion in street networks with different distances from coast using WRF-CFD coupling method. Sustain. Cities Soc. 2024, 115, 105757. [Google Scholar]
- Xin, B.; Dang, W.; Yan, X.; Yu, J.; Bai, Y. Dispersion characteristics and hazard area prediction of mixed natural gas based on wind tunnel experiments and risk theory. Process Saf. Environ. Prot. 2021, 152, 278–290. [Google Scholar] [CrossRef]
- Yao, M.; Liu, Z.; Xiao, P.; Xie, D. Numerical Modeling and Wind Tunnel Experimental Study of Pollutant Dispersion in Coastal Continent. Nucl. Eng. Technol. 2025, 57, 103693. [Google Scholar]
- Pantusheva, M.; Mitkov, R.; Hristov, P.O.; Petrova-Antonova, D. Air pollution dispersion modelling in urban environment using CFD: A systematic review. Atmosphere 2022, 13, 1640. [Google Scholar] [CrossRef]
- Auvinen, M.; Izbassarov, D.; Grönholm, T.; Hakala, J.; Kuula, J.; Asmi, E.; Hellsten, A. Quantitative validation in indoor dispersion modeling: Comparing large-eddy simulation results with experimental measurements. Phys. Fluids 2025, 37, 085107. [Google Scholar] [CrossRef]
- Rzeszutek, M.; Szulecka, A. Assessment of the AERMOD dispersion model in complex terrain with different types of digital elevation data. IOP Conf. Ser. Earth Environ. Sci. 2021, 642, 012014. [Google Scholar] [CrossRef]
- Li, Q.; Cai, X.; Kang, L. CFD Simulation of Atmospheric Flow and Diffusion under Building Disturbance Conditions. Environ. Sci. Res. 2013, 26, 829–837. (In Chinese) [Google Scholar]
- Fernández-Pacheco, V.M.; Fernández-Tena, A.; Ackermann, T.; Blanco-Marigorta, E.; Álvarez-Álvarez, E. Physical and CFD model used in the analysis of particles dispersion. Heliyon 2023, 9, e21330. [Google Scholar] [CrossRef] [PubMed]
- Owkes, M.; Homan, T.; Benson, M.; Banko, A. High fidelity simulations of contaminant dispersion in an urban environment with comparison to magnetic resonance imaging measurements. Environ. Fluid Mech. 2025, 25, 5. [Google Scholar] [CrossRef]
- Chen, S.; Li, W.; Zheng, W.; Liu, F.; Zhou, S.; Wang, S.; Zhang, T. Application of Optical Communication Technology for UAV Swarm. Electronics 2025, 14, 994. [Google Scholar] [CrossRef]
- Huang, Y.; Li, J.; Zheng, H. Modeling of Wildfire Digital Twin: Research Progress in Detection, Simulation, and Prediction Techniques. Fire 2024, 7, 412. [Google Scholar] [CrossRef]
- Cowell, N.H.; Chapman, L.; Topping, D.; James, P.; Bell, D.; Bannan, T.; Birkin, M. Moving from monitoring to real-time interventions for air quality: Are low-cost sensor networks ready to support urban digital twins? Front. Sustain. Cities 2025, 6, 1500516. [Google Scholar] [CrossRef]
- Gaertner, E.; Rinker, J.; Sethuraman, L.; Zahle, F.; Anderson, B.; Barter, G.E.; Viselli, A. IEA Wind TCP Task 37: Definition of the IEA 15-Megawatt Offshore Reference wind Turbine; National Renewable Energy Lab. (NREL): Golden, CO, USA, 2020. [Google Scholar]
- Kainz, S.; Quick, J.; Souza de Alencar, M.; Sanchez Perez Moreno, S.; Dykes, K.; Bay, C.; Bortolotti, P. IEA Wind TCP Task 55: The IEA Wind 740-10-MW Reference Offshore Wind Plants; National Renewable Energy Laboratory (NREL): Golden, CO, USA, 2024. [Google Scholar]
- Adamidis, P.; Pfister, E.; Bockelmann, H.; Zobel, D.; Beismann, J.O.; Jacob, M. The real challenges for climate and weather modelling on its way to sustained exascale performance: A case study using ICON (v2. 6.6). Geosci. Model Dev. 2025, 18, 905–919. [Google Scholar] [CrossRef]
- Williams, R.S.; Hogan, R.J.; Polichtchouk, I.; Hegglin, M.I.; Stockdale, T.N.; Flemming, J. Evaluating the Impact of Prognostic Ozone in IFS NWP Forecasts; European Centre for Medium-Range Weather Forecasts: Reading, UK, 2021. [Google Scholar]
Driver Category | Specific Driving Mechanism | Impact on Research Paradigm |
---|---|---|
Technological | Proliferation of High-Performance Computing (HPC) | Enabled high-fidelity Large Eddy Simulation (LES) to transition from a theoretical tool to a feasible research instrument, addressing the accuracy bottlenecks of RANS models. |
Rise of Artificial Intelligence (AI)/Machine Learning (ML) | Enabled rapid wind field prediction (from hours to seconds), fostering new research directions focused on real-time control and optimization. | |
Economic | Rapid Growth of the Global Wind Energy Market | The massive market size amplified the impact of model accuracy on economic benefits. |
Financial Impact of Annual Energy Production (AEP) Uncertainty | To secure favorable financing terms (lower cost of capital), project developers must reduce AEP prediction risks, creating strong demand for high-precision models. | |
Geopolitical & Policy | The Paris Agreement and Nationally Determined Contributions (NDCs) | Established a global political framework for decarbonization, using top-down pressure to compel nations to set renewable energy targets. |
National-Level Incentive pOlicies (e.g., U.S. Inflation Reduction Act) | Translated global goals into concrete financial incentives (e.g., tax credits), directly reducing the investment cost and risk of wind power projects. | |
Energy Independence and Security | Positioned domestic renewable energy development as a national security strategy, reducing reliance on external fossil fuels and providing political assurance for long-term investment. |
ANSYS Fluent | OpenFOAM | WRF | |
---|---|---|---|
License Model | Commercial license, high cost | Open-source, free | Open-source, free |
Primary Simulation Scale | Microscale (10−2–103 m) | Microscale (10−2–103 m) | Mesoscale (103–105 m) |
User Interface | Graphical User Interface (GUI), user-friendly | Command-line | Command-line |
Core Strengths | Industry standard, reliable and easy to use | Flexible, free, suitable for large-scale parallel computing | Provides realistic atmospheric background fields |
Core Limitations | High cost, poor customizability | High technical barrier to entry | Coarse resolution, unable to resolve local flow details |
Role | Microscale wind field solver | Microscale wind field solver | Provides boundary conditions for microscale simulations |
Method | Advantages | Limitations | Optimal Application Scenarios |
---|---|---|---|
Field Measurement | Provides authentic “ground-truth” data, serving as a benchmark for validation. Captures long-term wind climate characteristics at specific locations. | High cost and sparse measurement points hinder comprehensive capture of spatial variability in the wind field. Instrument installation and maintenance are challenging in remote areas. | Micro-siting for critical turbine locations. Supplying boundary conditions and validation data for CFD and wind tunnel models. |
Wind Tunnel Experiment | Allows for efficient and repeatable testing of terrain effects on the wind field in a controlled environment. Aids in understanding the physical mechanisms of complex flows. | Reynolds number mismatch makes it difficult to fully simulate the atmospheric boundary layer and thermal effects. High cost and model scaling ratios limit the precision of fine details. | Fundamental research on physical mechanisms (e.g., flow over hills, wake model development). Preliminary validation of parameterization schemes in CFD models. |
CFD Simulation | Relatively low-cost; provides high-resolution, full-domain wind field data. Offers flexibility to simulate various terrains and atmospheric conditions. | Accuracy is highly dependent on mesh quality, turbulence model selection, and boundary conditions. Requires rigorous validation and calibration against high-quality field measurement data. | Macro-siting, micro-siting, and layout optimization. Analysis of wake effects and energy yield prediction in complex terrain. |
Phenomenon/Application | Key Quantitative Finding | Implications for Design/Codes | Source |
---|---|---|---|
Wind loads on low-rise buildings | Due to abrupt changes in terrain roughness, local pressure coefficients can exceed ASCE 7 code values by more than 30%. | Correction factors in current codes, which are based on idealized terrain, are insufficient to reflect the influence of real 3D topography. | Kim et al. [139] |
Typhoon amplification effects at mountain gaps | Specific gap topographies can amplify typhoon wind speeds by over 25%. | Extreme event load assessment must consider the amplification effects of micro-topography. | Meng et al. [143] |
Wind environment mitigation at mountain tunnel portals | A novel transition section design can reduce local wind speed amplification effects by 15–20%. | Targeted aerodynamic shape design can significantly enhance traffic safety. | Peng et al. [136,137] |
Peak pressure prediction on building surfaces | Neural networks can improve the prediction accuracy of peak pressures on building surfaces in complex terrain by approximately 25%. | Data-driven models provide a new pathway for rapid, site-specific load assessment. | An et al. [138] |
Method | Core Principle | Main Advantages | Key Limitations | Best Application Scenarios |
---|---|---|---|---|
Field Measurement | Direct measurement in real environments | Provides “ground truth” data for validation | High cost; difficult to capture 3D structure | Long-term monitoring; numerical model validation |
Wind Tunnel | Scaled-down lab simulation | High controllability and repeatability | Significant thermal effect errors | Basic physical research; model verification |
RANS-CFD | Solves time-averaged equations | High computational efficiency, low cost | Poor accuracy for peak concentrations | Preliminary screening and long-term average concentration assessment for industrial site selection and environmental impact |
LES-CFD | Solves large-scale eddies directly | High fidelity; captures turbulent structure | High computational cost and demands | Consequence assessment of sudden accidents; refined risk assessment in high-risk areas |
WRF-CFD | Couples meteorological and CFD models | High accuracy for forecasting | High computational complexity | Regional air quality forecasting; pollutant diffusion research under complex weather conditions |
Feature | Intensive Observation Period (IOP) | Long-Term Routine Monitoring |
---|---|---|
Main Objective | Understanding physical mechanisms | Regulatory compliance and statistical validation |
Time Coverage | Short-term (days to weeks) | Long-term (months to years) |
Measurement Frequency | High (e.g., 1 Hz to 100 Hz) | Low (e.g., hourly averages) |
Typical Instruments | Multi-wavelength LiDAR, UAVs, sodar, research aircraft | Fixed-site gas/particle analyzers |
Core Advantages | High spatiotemporal resolution data | Cost-effective; strong statistical representativeness |
Core Limitations | Limited time coverage, high cost | Low spatial resolution, missing turbulent data |
Ideal Model Match | LES, high-fidelity CFD models | RANS, Gaussian, operational forecast models |
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Wang, W.; Chen, F. Wind Field Modeling over Hilly Terrain: A Review of Methods, Challenges, Limitations, and Future Directions. Appl. Sci. 2025, 15, 10186. https://doi.org/10.3390/app151810186
Wang W, Chen F. Wind Field Modeling over Hilly Terrain: A Review of Methods, Challenges, Limitations, and Future Directions. Applied Sciences. 2025; 15(18):10186. https://doi.org/10.3390/app151810186
Chicago/Turabian StyleWang, Weijia, and Fubin Chen. 2025. "Wind Field Modeling over Hilly Terrain: A Review of Methods, Challenges, Limitations, and Future Directions" Applied Sciences 15, no. 18: 10186. https://doi.org/10.3390/app151810186
APA StyleWang, W., & Chen, F. (2025). Wind Field Modeling over Hilly Terrain: A Review of Methods, Challenges, Limitations, and Future Directions. Applied Sciences, 15(18), 10186. https://doi.org/10.3390/app151810186