Next Article in Journal
Effectiveness of Erythrocyte Morphology Observation as an Indicator for the Selection and Qualification of Blood in a Mechanically Induced Hemolysis Test
Previous Article in Journal
Selenium Content of Goose Breast Meat Depending on the Type of Heat Processing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical and Analytical Estimation of the Wind Speed Causing Overturning of the Fast-Erecting Crane—Part II

Department of Machine Design and Composite Structure, Faculty of Mechanical Engineering, Cracow University of Technology, Al. Jana Pawła II 37, 31-864 Cracow, Poland
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(11), 4694; https://doi.org/10.3390/app14114694
Submission received: 8 May 2024 / Revised: 25 May 2024 / Accepted: 28 May 2024 / Published: 29 May 2024
(This article belongs to the Special Issue Structural Wind Engineering, 2nd Edition)

Abstract

:

Featured Application

The presented results of the current work can be exploited to ensure safer exploitation of the tower cranes concerning the strong wind and the possibility of overturning them.

Abstract

The currently presented work is a continuation of the previous one, where the estimation of the forces induced by the wind flow acting on the fast-erecting crane. In that work, the values of the aerodynamic forces were determined experimentally and numerically for the sectional models of the tower and jib. Next, the obtained results were compared with the appropriate standards. Now, the main aim is to determine the critical wind speed causing the overturning of the whole structure. At the very beginning, the numerical analysis of the simplified model of the crane on the real scale is studied. The computations are performed with the use of the ANSYS FLUENT R22. The simulations are performed for three different wind profiles, namely: urban terrain, village terrain, and open terrain. Moreover, the various geometric configurations of the crane in the wind direction are studied. The k-ε model of turbulent flow is exploited. The obtained critical values of the wind speed are confronted with those that are obtained from standards and estimations based on the results obtained from previous investigations performed for sectional models. The influence of the load carried by the crane is also taken into consideration in the overturning of the structure.

1. Introduction

With rapid climate change, strong, unexpected winds or gusts of wind have become a real danger for the various high-slender structures. Examples of such structures are the tower cranes used practically at all construction sites. Moreover, these kinds of weather phenomena take place in countries like Poland, where they have very rarely been reported before. Two tragic accidents involving tower cranes have occurred in Cracow (the southern part of Poland) for the last two years. These accidents were caused by a strong, unexpected gust of wind. In the first case [1,2], two people died. In the second case [3], one person died and three were injured. It shows how important this problem is.
Generally, the wind is not a steady-state phenomenon. It is rather a dynamic phenomenon accompanied by turbulence and sudden changes in direction. Moreover, the gusts of wind can be even two times greater in comparison to the mean wind speed. The frequency of the gusts most frequently is about 1 Hz [4]. The impact of the wind on high-slender lattice structures, like various tower cranes, could be completely different depending on the wind speed and strength. The wind can cause the vibrations of the lifted load or crane itself. Modern slender steel lattice structures, like tower cranes, consist of members that carry high-tension forces. It causes solid profiles to possess low-damping properties. Therefore, the tower cranes are sensitive to wind-induced vibrations with increasing amplitude, so-called galloping vibrations. Such a phenomenon is reported in the case of the tension bars of slender tower cranes. It leads to fatigue damage to the individual elements or even total failure of the whole structure [5]. The analysis of the dynamic response of the tower cranes is more complicated due to wind excitation, which possesses rather random characteristics [6]. Next, Jiang and Li [7] investigated the dynamic reliability of the tower cranes with the wind load. They used a linear autoregressive model to simulate the time history of multidimensional fluctuating wind samples in conjunction with the finite element method. and concluded that for a wind speed of 20 m/s, the structural strength is relatively sufficient. Further detail analysis reveals that the most dangerous geometry configuration of the crane is when the horizontal jib is not strictly perpendicular to the wind direction [8]. Hechmi El Ouni et al. [9] perform a numerical analysis of the tower carne loaded by the turbulent wind with the use of the finite element method to design the active system of vibration damping, which prevents the crane from collapsing. Oliveira and Correia [10] investigated the vibration of two different tower cranes induced by seismic and wind excitation. They used the finite element method, namely SAP2000 software. Finally, the tower cranes are also installed in super-high-rise buildings. A separate issue is the phenomenon related to wind-affecting tower cranes used during the construction of these kinds of super-high-rise buildings [11,12].
The effect of the wind is not limited only to the crane itself but also to the load being lifted. Depending on the size, shape, and weight of the load, this influence may have a decisive impact on the safe operation of the whole structure. Skelton et al. [13] designed a special container of wing shape for lifted loads for tower cranes, which operate on the construction sites of high buildings. This container significantly reduces the aerodynamic drag force caused by the wind. Monteiro and Moreira [14] analyzed an offshore oil and gas platform module lifted by a crane. They assumed the pendulum-like displacement of the lifted load was caused by the wind. The pendulum-like displacement is determined via a finite difference scheme, whereas the drag coefficients of the platform module are estimated through CFD analysis. Cekus et al. [15,16] studied the dynamic behavior of the lifted load of cuboid shapes by crane. The lifted load is subjected to wind. Finally, Jin et al. [17] investigated the impact of the lifted load subjected to wind on the tower crane. The load under wind pressure is modeled as a pendulum. Nguyen [18] designed a special spring-damper device that can reduce the payload vibration caused by wind.
If the wind or gust of wind is sufficiently strong, the whole tower crane can be overturned. This kind of accident is the most dangerous because of the potential deaths and relatively significant material losses. However, works devoted strictly to the problem of tipping over cranes of different kinds or other similar structures caused by a strong wind are very rare. Here can be quoted the following works concerning the overturning of the gantry container crane with payload [19], gantry cranes [20,21,22], or scissor lift [23].
Finally, it is worth noting that the surroundings (for example, buildings and other cranes) of the operating crane-like structures can have a significant impact on the wind load acting on the investigated structure. This phenomenon is known as an interference effect. As before in the case of the overturning of the crane-like structures due to wind load, the works devoted to this problem are also very rare. Here can be quoted the work by Wu et al. [24], where the system of three container cranes is studied. Each other’s impact as well as the dynamic response of the cranes are studied. Furthermore, the antenna system and its mutual impact are investigated by Holmes et al. [25], Martin et al. [26], and Carril et al. [27]. It has been found that the interference factor of microwave antenna dishes is greater than one for specific wind directions. It is worth noting that the lattice structure of the antenna is very similar to that of most crane structures. This phenomenon is also studied in the case of scaffolding [28], low-rise buildings [29,30,31,32], tall buildings [33], and cooling towers [34,35].
The currently presented work is a continuation of the problem discussed in the previous one [36]. In that work, we tried to determine the magnitude of the aerodynamic forces induced by the wind acting on the sectional model of the tower lattice and the horizontal jib lattice. The experimental tests in the aerodynamic tunnel and the CFD simulations were carried out. The obtained results were compared with appropriate standards and codes. The main aim of the current work is to investigate the impact of the wind on the fast-erecting crane, namely determining the most dangerous geometric configuration of the structure against the wind direction. We are also going to take into account the potential external load that is carried by the crane. To deal with this problem, we will exploit the results reported in the previous work as well as the CFD simulations of the whole structure on a real scale with the various wind profiles considered.

2. Materials and Methods

2.1. Object of Study

As in the previous work [36], the object of investigation is a fast-erecting 63 K crane by Liebherr. Its detailed description and the technical parameters are described in [36,37]. Here, it is worth noting that this structure is very similar to that that fell in Cracow in 2022 [1]. In Figure 1, the studied crane is shown. Here we only mention that the total height of the crane is equal to H = 31 m, and the length between the axis of rotation and the end of the jib is equal to L = 35 m.
To perform the CFD simulations of the airflow around the investigated crane, an appropriate simplified model of the real structure should be prepared. It is worth noting that the geometry model should enable the automatic generation of a good-quality mesh of the finite elements. Therefore, in such a complicated object as a lattice structure, some simplification must be introduced. Further, all simplifications will be discussed concerning the real structure. For the creation of the CAD model, the ANSYS R22 DesignModeler software is exploited. The simplifications that are introduced are as follows:
  • The CAD model does not contain: rope immobilizing the horizontal jib, rear lashing, guy support I, crane jib—lashing rope II, crane jib—lashing rope III, crane jib—lashing rope I, jib—assembly rope, guy support II—head, lifting rope with hook.
  • The jib extension at the end of the jib is made of the same lattice as other parts of the horizontal part of the crane.
  • The wider and narrower parts of the crane tower truss are connected directly. These parts do not overlap each other like in the real structure.
  • The dimensions of the transverse sections of some elements of the lattice are slightly changed to avoid problems and errors during automatic mesh generation.
In Figure 2, the details of the CAD model of the studied fast-erecting crane are depicted.

2.2. CFD Simulation

The CFD simulations of the airflow around the studied lattice structure of the studied tower crane were performed with the use of the ANSYS R22 Fluent with Meshing module. The computations were carried out for various configurations of the crane geometry concerning the wind direction as well as for three different wind profiles (for urban terrain, village terrain, and open terrain). In Figure 3, it is depicted how the geometry of the crane is defined concerning the wind direction. The assumed wind direction is parallel to the X-axis of the global Cartesian coordinate system. The geometry configuration is defined by the angle θ between the X-axis of the coordinate system and the horizontal jib. The computations were performed for the following angles, namely: θ = 0°, 15°, 30°, 45°, 60°, 75°, and 90° for each assumed wind profile.
The investigated structure is immersed in the cuboid volume filled with an air of geometrical dimensions: width × length equal to 125 m × 130 m, and height equal to 65 m. One wall is assumed to be an inlet (blue wall in Figure 4), and the opposite one is the outlet, as shown in Figure 4. The ground (the wall on which the crane is installed) is stationary, while the rest of the walls are assumed to be movable. The wind speed on these walls varies according to the predefined wind profile for the inlet. The external dimensions of the cuboid seem to be enough. The lattice structures, like tower cranes, cause a limited disturbance of the airflow due to the relatively small area of the whole structure. Standard air properties at sea level (temperature T = 15 °C, ambient pressure p0 = 101,325.25 Pa) were assumed, namely: kinematic viscosity ν = 1.7894 × 105 kg/(m·s), density ρ = 1.225 kg/m3.
For simulation of the turbulent flow of air, the standard k-ε model with a standard wall function is exploited. According to the authors’ experience so far [22,23,36], this model allows obtaining sufficiently precise results in a situation where the magnitude of the resultant aerodynamic forces (static pressure) is mainly determined by the drag force, while the lift force is several orders of magnitude smaller. Furthermore, the k-ε model of turbulent flow is still successfully exploited by other authors, for example [8,11,38]. Here, it is worth noting that the other available models of turbulent flow are rather difficult to use in this case. For example, the use of the Reynolds stress model [24] causes the convergence of the numerical solution to be unacceptably slow. Another possibility is to involve the k-ω turbulent flow [14,39,40]. In the ANSYS R22 Fluent software, this model is available in several different variants. Depending on the variant of this model used, the results in the form of aerodynamic force values may vary significantly. Moreover, when using the k-ω model, it is necessary to generate an appropriate FEM mesh in the boundary layer. This results in a significant increase in the number of elements, which seems unacceptable in the case of the currently considered structure.

2.2.1. Wind Profile

The investigated structure is subjected to a wind load. Generally, the wind profile of Davenport is described by the following formula [41]:
z = v g z z g α
where vg is the wind velocity measured at the height zg over the ground level. The exponent α takes values of 0.4, 0.28, and 0.16 depending on the type of terrain, namely urban terrain, village terrain, or open terrain, respectively. In the current work, we arbitrarily assumed that in all studied cases vg = 6 m/s at the height zg = 10 m. In Figure 5, the assumed wind profiles are shown.
Based on the wind profiles depicted above, the reference wind velocity is also computed for the three wind velocity profiles according to the following formula:
v r e f = 0 H z V z d z H Z ,
where HZ is the total height of the crane, where Hz = 34.5 m. The values of the reference wind speed, computed according to Formula (2) for assumed wind profiles, are equal as follows: urban terrain vref = 7.0321 m/s, village terrain vref = 6.62928 m/s, and open terrain vref = 6.30482 m/s. These values will be used to compute the aerodynamic force and moment coefficients for the whole crane structure. Moreover, it is assumed that the turbulent intensity equals 9% with the turbulent length scale LTURB = 3.5 m. These values are assumed to be arbitrary. The choice was made according to the parameters discussed in our previous work.

2.2.2. Finite Cell Mesh

In Figure 6, the finite cell mesh is generated automatically by the ANSYS Fluent Mesh module. It is assumed that the minimal length of the cell edge is equal to lmin = 0.028 m (value independent of the geometry configuration), and the maximal length of the cell edge is generally greater than lmax > 6 m, depending on the geometry configuration.
Such a minimal cell edge length ensures that during the automatic mesh generation of tetrahedral elements, there are no degenerated or highly distorted cells. The existence of such cells makes it very difficult to obtain a convergence solution to the numerical problem. For considered geometry configurations, the number of nodes is over 2.5 million, and the number of cells is over 14 million. Depending on the geometry configuration, the number of nodes and cells varies not significantly.
To verify the precision of the numerical solutions, which are obtained for the assumed minimal length of the cell edge, the convergence test was performed. In Table 1 and Table 2, the total number of nodes, faces, and cells generated for the following minimal lengths of the cell edges is reported: lmin = 0.056 m, 0.028 m, and 0.014 m. The convergence test was performed for the geometry configuration, for which the angle θ = 90°. The overturning moment Mtip is the moment that is measured concerning the tipping line, which is shown in Figure 7.
Taking into consideration the results presented above in Table 1 and Table 2, the assumed value of lmin = 0.028 m is quite reasonable. All other calculations will be performed for this value of lmin.
Finally, it is worth noting that le = 0.03 m is the largest element size, which enables automatic finite cell generation without warnings or errors. On the other hand, le > 0.02 m leads to an enormous, large number of nodes and cells. Moreover, considering the mentioned limitation connected with element size, the performed convergence test should be rather threaded as a sensitivity analysis.

3. Results of CFD Simulation

Figure 8a–c presents the results of the carried-out CFD simulations for different kinds of terrain and geometry configurations of the crane. It should be noted that FZ aerodynamic force components are one order of magnitude less in comparison with the FY components. Therefore, they are not reported here. As can be observed in all investigated cases, the maximum of the FX component of the aerodynamic force and the overturning moment CM are observed for the angle θ = 75°. For larger values of the angle θ, these values decrease. The value of the FY aerodynamic force is one order of magnitude less in comparison with the FX component. It is worth stressing here that for such defined wind profiles, the greatest values of the aerodynamic forces are obtained for urban terrain and the smallest ones for open terrain. Finally, in Table 3, the aerodynamic force coefficients are collected, which are computed according to Formula (3), where vref is determined with the use of (1) and (2).
From the obtained values of the FX and FY forces, the aerodynamic coefficients CX and CY were determined, which served as the basis for stability calculations of the 63 K crane. The appropriate values of the aerodynamic force and moment coefficients are determined with the use of the following formulas:
C X = 2 F X ρ v r e f 2 A r e f ,   C Y = 2 F Y ρ v r e f 2 A r e f ,   C M = 2 M ρ v r e f 2 A r e f B r e f
where Aref is the area of the shadow normally projected by the crane lattice members on a plane parallel to the wall; Aref = 23.31 m2; and Bref is the conventionally accepted characteristic dimension. In our case, Bref is assumed to be equal to the total height of the crane; Bref = HZ. The convergence test is carried out for urban terrain, where vref = 7.0321 m/s. Other parameters are as discussed above. The values of the obtained aerodynamic coefficients are shown in Table 3.

3.1. Validation of Numerical Results Obtained from CFD Simulation with Experimental Results for Sectional Models

The results of the CFD simulations have been verified by comparing the overturning moments that act on the structure of the tower crane 63 K concerning one of the tipping edges for the two characteristic positions of the tower jib, namely for the angles θ = 0° and θ = ±90°. As a basis for the verification, it is assumed that the experimental results reported in [36]. These results concern the values of the aerodynamic forces FX and FY obtained for the sectional models of the tower truss and jib truss of the tower crane 63 K. With the values of these forces, it is possible to estimate the overturning moment caused by the wind according to the appropriate standard [41]. In the case of standard, the most dangerous position of the crane jib concerning the wind direction is taken into consideration, namely, the jib position is perpendicular to the wind direction θ = 90°.
The overturning moments are equal to the sum of the moments caused by the masses of the particular parts of the crane jib G1–G4 without the external load but with the aerodynamic force generated by the wind taken into account, as shown in Figure 9. The geometrical dimensions as well as the masses of the structure elements are assumed according to the technical and commissioning documentation [37]. Therefore, it should be noted here that the mentioned parameters are estimated with a certain approximation, and they should not be treated as data provided by the producer. The test object itself is treated as an example for calculating the stability of a low-slewing crane.
The schema of the force distribution acting on the studied crane is shown in Figure 9 and Table 4. The aerodynamic forces induced by wind are marked with the symbol W. The values of these forces are determined in the next step of the analysis according to the standard Eurocode for θ = 90° [41], experimental results for θ = 0° and θ = ±90° [36], and the results obtained from the CFD simulations for the θ range from −90° to +90° with steps equal to 15°.

3.2. The Wind Force Estimation According to the Standard

The estimation of the wind force W according to the standard [41] is based on the reference area Aref of the whole supporting structure of the tower crane and the air pressure, which acts on the structure under assumed wind speed vg = 6 m/s at the height H = 10 m above the ground level for the urban terrain. Due to the significant changes in the wind force together with the increasing height of the tower crane, the reference area Aref was split into three parts, namely: Aref1 = 11.04 m2 of the crane jib, Aref2 = 12.35 m2 of the tower, and Aref3 = 3.91 m2 of the counterweight. It should be stressed that according to various standards [39,40,41,42,43,44,45,46], the load induced by the wind, which acts on structures, is treated as a static normal pressure subjected to the external surface of a designed structure. Therefore, it is assumed that the wind pressure p = 245 N/m2 corresponds to the wind speed equal to v = 20 m/s. Thus, for a particular reference area and wind speed vg = 6 m/s at height h = 10 m the following forces are obtained, namely: for Aref1, where h1 = 33.81 m, W1 = 359.6 N; for Aref2, where h2 = 20.33 m, W2 = 341.84 N; and for Aref3, where h3 = 2.83 m, W3 = 57.5 N. Regarding the fact that the value of W3 is relatively small in comparison with the other ones (as well as the small height and large mass of the counterweight), in further computations, the overturning moment induced by this wind force was neglected. The obtained results MO ECode90° as a sum of the overturning moments MO (Table 4) and the moments induced by wind MW ECode90° [41] with the wind force taken into account are shown in Figure 10 as a constant value, which is independent of the rotation of the jib.

3.3. The Wind Force Estimation Based on the Experiment

In Figure 10, there are three points that correspond to the overturning moments MO exp 0°; 90°. These moments are the sum of the values of the moments estimated based on the wind force for the position of the crane jib, corresponding to θ = 0° and θ = ±90°. The appropriate values are determined with the use of force coefficients CX obtained from experimental analysis [36] of the sectional models and the moments MO (Table 4). In the experimental investigations, two sectional models were analyzed, namely part of the crane tower and part of the crane jib. At the very beginning of the current analysis, the reference area of the tower ArefT and the reference area of the jib ArefJ should be determined with the model scale considered. It should be noted here that the determination of the total wind force acting on the crane jib is relatively easy because all parts of the jib are located at the same height, namely h9 in Figure 11. However, the crane tower consists of eight parts—sectional models. Thus, it is necessary to determine the value of the wind force that acts on each section (it is assumed that the force is applied in the middle of each section) and compute the appropriate moment. It results in eight forces acting on the eight heights. The resultant overturning moment induced by the wind is equal to the sum of the product of these eight pairs of force W1–W8 and corresponding height h1–h8, which is shown in Figure 11.
The values of the wind speed at the assumed height can be determined according to Formula (1) for urban terrain. Next, the total wind force acting on the crane jib is determined using the following equation:
W J = 0.5 · v 9 ( h 9 ) 2 · A r e f J · ρ · C X J ,
In the case of the tower the particular wind forces are estimated for i = 1, 2, …, 8 as:
W T i = 0.5 · v i ( h i ) 2 · A r e f T · ρ · C X T ,
and, finally, moments for jib MWJ and tower MWT:
M W J = W J · h 9 ,   M W T = i = 1 8 W T i · h i
The numerical data and values of the calculated forces and moments for both angles are shown in Table 5. Similar computations were also carried out in the case of the laminar flow of the air. The results are presented in Table 6.
Concerning the above analysis, the total overturning moment, together with the impact of the wind MO exp 0°; 90° considered, can be computed. This moment is a sum of the moment Mo presented in Table 4, induced by the mass of the particular parts of the structure, and the moments MWJ and MWT shown in Table 5.
M O   exp   0 ° ;   90 ° = M o + M W J + M W T .
The values of the obtained overturning moments are shown in Figure 10. The symbol vref* denotes the arithmetic average wind speed, computed according to Figure 11.

3.4. The Wind Force Estimation According to the CFD

Based on the above-determined coefficients CX and CY (as a result of the CFD simulation), the values of the aerodynamic forces are evaluated. These quantities are a function of the wind speed. Next, it is assumed that the wind forces are applied and computed at the height, equal to jib height h1. Because in the case of the real structure, the measurement of the wind speed is performed at the top of the tower crane, for the determination of the wind forces WX, WY, it is assumed that the height is h1. It is worth noting that WX and WY act relative to the tipping edge of the investigated crane (63 K). These forces are defined as a function of wind speed.
W X = F X cos θ ,       W Y = F Y sin θ ,   W = W X 2 + W Y 2
In the next step, the reduced forces W and overturning moment MoCFD63K are described as a function of the rotation angle theta of the jib position.
M O   CFD   63 K = M O + W · h 1 .
As is shown in Figure 10, the three isolated points from the experiment (turbulent and laminar flow). which correspond to the jib position for jib rotation angle theta equal to 0° and 90°, almost overlap each other with the characteristic of the overturning moment. The overturning moment is based on the values of the wind forces obtained from CFD simulations.
The determination of the wind forces in all cases was performed according to the standards [41,46].

4. Discussion

4.1. The Maximum Wind Force

A comparison of overturning and stabilizing moments obtained from numerical tests with calculations based on Mo ECode90° [41] for different wind speeds acting on the jib height at h1 = 33.81 m is shown in Figure 12 and Figure 13.
Taking into account the above results, it is possible to determine the characteristics of the overturning moment as a function of wind speed for an angle of θ = 15° with the maximum load on the boom and for an angle of θ = 45° with no load on the boom. The results obtained are shown in Figure 14.
Using the data shown in Figure 14 and Figure 15, we can determine the maximum wind speed at which crane toppling will occur (Table 7).

4.2. Trace of the Center of Gravity

The trace of the center of gravity is a perpendicular projection of the center of gravity (in this case, tower crane 63 K) on the rectangular plane limited by the tipping edges. The geometrical dimensions of this plane are determined by the four supports of the investigated crane. In the current analysis, it is assumed that the overturning of the crane is possible only concerning the edge, which is oppositely located in the wind direction (Figure 3). It should be noted that the wind direction is constant to the contrary of the crane jib, which can rotate about its axis by the angle theta and the value of the wind speed. The two cases have been investigated. In the first case, the crane does not carry an external load; Q1 = 0. In the second case, the crane carries the maximal load Q1 = 3050 kg (Figure 9).
The trace of the center of gravity, depending on the angle θ of rotation of the crane jib, generates the closed contour. This curve can be described with the help of the guiding ray. Below are the formulas describing the guiding ray in the case of a lack of wind rc and in the case where the wind is present rcw. The guiding ray is defined by its coordinates xc, yc and xcw, ycw:
x c θ = r c · cos θ ,   y c θ = r c · sin θ ,
where:
r c = M 7 + M 7 Q 1 M 1 M 2 M 3 M 4 M 5 Q 1 + G 1 + G 2 + G 3 + G 4 + G 5 + G 6 + G 7 + G 8 ,
and the trace of the center of gravity at the time of the wind rcw:
x c w θ = r c · cos θ + r c + r c w θ ,   y c w θ = y c θ ,
where:
r c w θ = M 7 + M 7 Q 1 M 1 M 2 M 3 M 4 M 5 M o C F D 63 K θ Q 1 + G 1 + G 2 + G 3 + G 4 + G 5 + G 6 + G 7 + G 8 + F x C F D 63 K θ .
The obtained contours are presented in Figure 15 for the trace of the center of gravity in the case of maximal load, and the second case with no load is shown in Figure 16.

5. Conclusions

The presented work is devoted to the problem of the loss of stability of the tower cranes caused by strong wind or gusts of wind. In the first step of the analysis, the numerical simulations where the CFD ANSYS R22 Fluent software is exploited were performed for three different wind profiles. It occurred that the most dangerous conditions are met in urban terrain. It is caused by the significant gradient of the wind speed value along the vertical direction. Next, the obtained results, mainly the values of the aerodynamic force coefficients, are confronted with those obtained from previously performed experimental investigations. This experimental analysis was reported in our previous work concerning the studies of the sectional models of the tower crane truss and the jib truss performed in the aerodynamic tunnel. The results are also verified with the data provided by the appropriate standards. A relatively good agreement is observed in the case of the jib position, which corresponds to the values of the theta angle at 0° and 90°. However, the maximal values of the aerodynamic forces are observed for θ = 15° or 30° degrees. Finally, the influence of the external load is also taken into account. The new analytical formula is provided, which enables estimation of the trace of the center of gravity in the case when the crane is subjected to external load and without the load. The presence of the wind causes the characteristic “heart-like” shape of the curves, which describes the trace of the center of gravity as a function of the jib rotation.
Finally, it is worth noting that the most dangerous configuration of the tested crane is the one for which the angle θ = 15° with load and θ = 45° without load. These configurations correspond to a critical wind speed of v = 14.9 m/s (53.6 km/h) and v = 28.3 m/s (101.9 km/h), respectively. The obtained values agree with the wind speed that caused the crane disaster in Cracow (Poland) in 2023 [1,2]. This is particularly important because during the safe configuration of the crane (the boom is aligned parallel to the wind direction), a sudden, strong gust of wind is capable of toppling the entire crane, as can be seen in the figures, which show the trace of the center of gravity.

Author Contributions

Conceptualization, M.A.; methodology, M.A. and M.B.; software, M.A. and M.B; validation, M.A. and M.B.; numerical analysis, M.B.; experimental investigation, M.A.; data curation, M.A. and M.B.; writing—original draft preparation, M.A. and M.B.; writing—review and editing, M.A. and M.B.; visualization, M.A. and M.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wiosło, M.; Grochot, A. The Gale Overturned a Crane in Krakow. Two People Are Dead, RMF24. 2022. Available online: https://www.rmf24.pl/regiony/krakow/news-wichura-przewrocila-dzwig-w-krakowie-dwie-osoby-nie-zyja-fil,nId,5838306#crp_state=1 (accessed on 27 February 2024). (In Polish).
  2. Release from the Press Office of the Institute of Meteorology and Water Management, Warsaw, 16 February 2022. Available online: https://danepubliczne.imgw.pl/data/arch/ost_meteo/2022/ (accessed on 27 February 2024). (In Polish).
  3. Ciryt, B. Fatal Accident at the Construction Site of the Court in Wieliczka. The Crane Fell onto an Employee’s Container, and One Person Died. Watch the Video, Wieliczka Nasze Miasto. 2023. Available online: https://wieliczka.naszemiasto.pl/smiertelny-wypadek-na-budowie-sadu-w-wieliczce-dzwig/ar/c1-9511255 (accessed on 27 February 2024). (In Polish).
  4. König, G.; Zilch, K.; Lappas, G. Wind Loading of Shipyard Gantry Cranes—A Comparison of Full-Scale Measurement, Wind Tunnel Test and Gust Factor Approach, Wind Engineering. In Proceedings of the Fifth International Conference, Fort Collins, CO, USA, 8–13 July 1979; Volume 2, pp. 911–923. [Google Scholar]
  5. Klinger, C. Failures of cranes due to wind induced vibrations. Eng. Fail. Anal. 2014, 43, 198–220. [Google Scholar] [CrossRef]
  6. Jiang, H.; Li, S. The Wind-Induced Vibration Response for Tower Crane Based on Virtual Excitation Method. Open Mech. Eng. J. 2014, 8, 201–205. [Google Scholar] [CrossRef]
  7. Jiang, H.; Li, Y. Dynamic Reliability Analysis of Tower Crane with Wind Loading. IOP Conf. Ser. Mater. Sci. Eng. 2019, 677, 052031. [Google Scholar] [CrossRef]
  8. Chen, W.; Qin, X.R.; Yang, Z.; Zhan, P. Wind-induced tower crane vibration and safety evaluation. J. Low Freq. Noise Vib. Act. Control. 2020, 39, 297–312. [Google Scholar] [CrossRef]
  9. Hechmi El Ouni, M.; Ben Kahla, N.; Islam, S.; Jameel, M. A Smart Tower Crane to Mitigate Turbulent Wind Loads. Struct. Eng. Int. 2021, 31, 18–29. [Google Scholar] [CrossRef]
  10. Oliveira, S.C.; Correia, P.M.B. Comparison of the seismic and wind analyses of two tower cranes. J. Vibroeng. 2021, 23, 956–975. [Google Scholar] [CrossRef]
  11. Lu, Y.; Gao, M.; Liang, T.; He, Z.; Feng, F.; Pan, F. Wind-induced vibration assessment of tower cranes attached to high-rise buildings under construction. Automat. Constr. 2022, 135, 104132. [Google Scholar] [CrossRef]
  12. Lu, Y.; Zhang, L.; He, Z.; Feng, F.; Pan, F. Wind-induced vibration fragility of outer-attached tower crane to super-tall buildings: A case study. Wind. Struct. 2021, 32, 405–421. [Google Scholar]
  13. Skelton, I.; Demian, P.; Glass, J.; Bouchlaghem, D.; Anumba, C. Lifting Wing in Constructing Tall Buildings—Aerodynamic Testing. Buildings 2014, 4, 245–265. [Google Scholar] [CrossRef]
  14. Monteiro, F.A.; Moreira, R.M. A CFD Analysis of Wind Effects on Lifted Loads. Int. J. Adv. Eng. Res. Sci. 2019, 6, 365–371. [Google Scholar] [CrossRef]
  15. Cekus, D.; Gnatowska, R.; Kwiatoń, P. Impact of Wind on the Movement of the Load Carried by Rotary Crane. Appl. Sci. 2019, 9, 3842. [Google Scholar] [CrossRef]
  16. Cekus, D.; Kwiatoń, P.; Geisler, T. The dynamic analysis of load motion during the interaction of wind pressure. Meccanica 2021, 56, 785–796. [Google Scholar] [CrossRef]
  17. Jin, L.; Liu, H.; Zheng, X.; Chen, S. Exploring the Impact of Wind Loads on Tower Crane Operation. Hindawi Math. Probl. Eng. 2020, 2020, 2807438. [Google Scholar] [CrossRef]
  18. Nguyen, T.K. Combination of feedback control and spring-dumper to reduce the vibration of crane payload. Arch. Mech. Eng. 2021, 68, 165–181. [Google Scholar] [CrossRef]
  19. Lee, S.W.; Shim, J.J.; Han, D.S.; Han, G.J.; Lee, K.S. An Experimental Analysis of the Effect of Wind Load on the Stability of a Container Crane. J. Mech. Sci. Technol. 2007, 21, 448–454. [Google Scholar] [CrossRef]
  20. Frendo, F. Gantry crane derailment and collapse induced by wind load. Eng. Fail. Anal. 2016, 66, 479–488. [Google Scholar] [CrossRef]
  21. Su, J.-C.; Li, L.; Chan, P.W.; Zhou, Q.-J.; Yang, H.-L. Numerical simulation research on the overturning of gantry crane by downbursts. Heliyon 2023, 9, e18641. [Google Scholar] [CrossRef] [PubMed]
  22. Augustyn, M.; Barski, M. Estimation of the Wind Load Required to Cause the Overturning of a Gantry Crane, Comparing Different Structures of the Main Horizontal Girder. Appl. Sci. 2024, 14, 1092. [Google Scholar] [CrossRef]
  23. Augustyn, M.; Barski, M.; Chwał, M.; Stawiarski, A. Numerical and Experimental Determination of the Wind Speed Value Causing Catastrophe of the Scissor Lift. Appl. Sci. 2023, 13, 3528. [Google Scholar] [CrossRef]
  24. Wu, X.; Sun, Y.; Wu, Y.; Su, N.; Peng, S. The Interference Effects of Wind Load and Wind-Induced Dynamic Response of Quayside Container Cranes. Appl. Sci. 2022, 12, 10969. [Google Scholar] [CrossRef]
  25. Holmes, J.D.; Banks, R.W.; Roberts, G. Drag and aerodynamic interference on Microwave dish antennas and their supporting towers. J. Wind. Eng. Ind. Aerod. 1993, 50, 263–270. [Google Scholar] [CrossRef]
  26. Martín, P.; Elena, V.; Loredo-Souz, A.M.; Camaño, E.B. Experimental study of the effects of dish antennas on the wind loading of telecommunication towers. J. Wind. Eng. Ind. Aerod. 2016, 149, 40–47. [Google Scholar] [CrossRef]
  27. Carril, C.F., Jr.; Isyumov, N.; Brasil, R.M.L.R.F. Experimental study of the wind forces on rectangular latticed communication towers with antennas. J. Wind. Eng. Ind. Aerod. 2003, 91, 1007–1022. [Google Scholar] [CrossRef]
  28. Feng, W.; Tamura, Y.; Yoshida, A. Interference effects of a neighboring building on wind loads on scaffolding. J. Wind. Eng. Ind. Aerod. 2014, 125, 1–12. [Google Scholar]
  29. Holmes, J.D. Wind pressure on tropical housing. J. Wind. Eng. Ind. Aerod. 1994, 53, 105–123. [Google Scholar] [CrossRef]
  30. Li, G.; Gan, S.; Li, Y.X.; Wang, L. Wind-induced interference effects on low-rise buildings with gable roof. J. Wind. Eng. Ind. Aerod. 2017, 170, 94–106. [Google Scholar] [CrossRef]
  31. Quan, Y.; Gu, M.; Yukio, T.; Huang, P. Aerodynamic interference of wind loads on roofs of low-rise buildings. J. Tongji Univ. Nat. Sci. 2009, 37, 1576–1580. [Google Scholar]
  32. Pindado, S.; Meseguer, J.; Franchini, S. Influence of an upstream building on the wind-induced mean suction on the flat roof of a low-rise building. J. Wind. Eng. Ind. Aerod. 2011, 99, 889–893. [Google Scholar] [CrossRef]
  33. Kareem, A.; Kijewski, T.; Lu, P.C. Investigation of interference effects for a group of finite cylinders. J. Wind. Eng. Ind. Aerod. 1998, 77–78, 503–520. [Google Scholar] [CrossRef]
  34. Li, G.; Cao, W.B. Structural analysis and optimization of large cooling tower subjected to wind loads based on the iteration of pressure. Struct. Eng. Mech. 2013, 46, 735–753. [Google Scholar] [CrossRef]
  35. Ke, S.T.; Wang, H.; Ge, Y.J. Interference effect and the working mechanism of wind loads in super-large cooling towers under typical four-tower arrangements. J. Wind. Eng. Ind. Aerod. 2017, 170, 197–213. [Google Scholar] [CrossRef]
  36. Augustyn, M.; Barski, M.; Chwał, M.; Stawiarski, A. Experimental and Numerical Estimation of the Aerodynamic Forces Induced by the Wind Acting on a Fast-Erecting Crane. Appl. Sci. 2023, 13, 10826. [Google Scholar] [CrossRef]
  37. Cranes, Hoists & Material Handlers: Tower Cranes, Technical Data Tower Crane LIEBHERR Turmdrehkran 63K. Liebherr-Werk Biberach GMBH, Postfach 1663, D-7950 Biberach an der Riss 1, Germany.
  38. Zan, Y.F.; Guo, R.N.; Bai, X.; Ma, Y.; Yuan, L.H.; Huasng, F.X. Wind and current loads on a pipelaying crane vessel. IOP Conf. Ser. Earth Environ. Sci. 2020, 612, 012059. [Google Scholar] [CrossRef]
  39. Yeon, S.M.; Kwon, C.S.; Kim, Y.C.; Kim, K.S. Study of the lift effect on wind load estimation for a semi-submersible rig using the maritime atmospheric boundary layer model. Int. J. Nav. Arch. Ocean 2022, 14, 100419. [Google Scholar] [CrossRef]
  40. He, Z.; Gao, M.; Liang, T.; Lu, Y.; Lai, X.; Pan, F. Tornado-affected safety assessment of tower cranes outer-attached to super high-rise buildings in construction. J. Build. Eng. 2022, 51, 104320. [Google Scholar] [CrossRef]
  41. ISO 4302:2016 EN; Cranes-Wind Load Assessment. ISO Copyright Office: Geneva, Switzerland, 2016.
  42. JIS B 8830-2001; Cranes-Wind Assessment. Japanese Industrial Standards Committee: Tokyo, Japan, 2001.
  43. BS 2573-1; British Standard. Rules for the Design of Cranes Part 1: Specifications for Classification, Stress Calculations, and Design Criteria for Structures (4th Revision). BSI: San Jose, CA, USA, 1983.
  44. GB/T 3811-2008; Design Rules for Cranes. General Administration of Quality Supervision. Inspection and Quarantine of the People’s Republic of China: Beijing, China, 2008.
  45. ASCE. Minimum Design Loads and Associated Criteria for Buildings and Other Structures; ASCE 7: Reston, VA, USA, 2016. [Google Scholar]
  46. ISO 8686-1:2012; Cranes. Design principles for loads and load combinations. Part 1: General. ISO Copyright Office: Geneva, Switzerland, 2012.
Figure 1. The object of study: 63 K crane: (a) real structure; (b) pressure distribution as a result of the CFD simulation.
Figure 1. The object of study: 63 K crane: (a) real structure; (b) pressure distribution as a result of the CFD simulation.
Applsci 14 04694 g001
Figure 2. Details of the CAD model of the fast-erecting crane: (a) crane basis; (b) tower lattice; (c) join of the tower and jib; (d) jib lattice.
Figure 2. Details of the CAD model of the fast-erecting crane: (a) crane basis; (b) tower lattice; (c) join of the tower and jib; (d) jib lattice.
Applsci 14 04694 g002
Figure 3. The assumed configuration of the crane geometry.
Figure 3. The assumed configuration of the crane geometry.
Applsci 14 04694 g003
Figure 4. The investigated crane inside the cuboid volume filled with air.
Figure 4. The investigated crane inside the cuboid volume filled with air.
Applsci 14 04694 g004
Figure 5. Applied wind profiles.
Figure 5. Applied wind profiles.
Applsci 14 04694 g005
Figure 6. The finite cell mesh: (a) basis of the tower; (b) tower lattice; (c) join of the tower and jib; (d) jib lattice.
Figure 6. The finite cell mesh: (a) basis of the tower; (b) tower lattice; (c) join of the tower and jib; (d) jib lattice.
Applsci 14 04694 g006
Figure 7. The location of the tipping line.
Figure 7. The location of the tipping line.
Applsci 14 04694 g007
Figure 8. The values of aerodynamic forces (a) FX component, (b) FY component, (c) overturning moment M, estimated concerning the tipping line (Figure 7), for the wind speed vg = 6 [m/s] measured at 10 m over the ground.
Figure 8. The values of aerodynamic forces (a) FX component, (b) FY component, (c) overturning moment M, estimated concerning the tipping line (Figure 7), for the wind speed vg = 6 [m/s] measured at 10 m over the ground.
Applsci 14 04694 g008
Figure 9. The distribution of loads acting on the 63 K crane, which was considered in the calculations according to Eurocode (θ = 90°).
Figure 9. The distribution of loads acting on the 63 K crane, which was considered in the calculations according to Eurocode (θ = 90°).
Applsci 14 04694 g009
Figure 10. The distribution of loads acting on the 63 K crane, which was considered in the calculations according to Eurocode [41] (θ = 90°) in the case of the maximal load.
Figure 10. The distribution of loads acting on the 63 K crane, which was considered in the calculations according to Eurocode [41] (θ = 90°) in the case of the maximal load.
Applsci 14 04694 g010
Figure 11. The distribution of loads acting on the 63 K crane, which was considered in the calculations according to experimental results.
Figure 11. The distribution of loads acting on the 63 K crane, which was considered in the calculations according to experimental results.
Applsci 14 04694 g011
Figure 12. Overturning moment characteristics obtained from CFD compared with calculated overturning moment and stability moment according to Eurocode as a function of jib rotation angle θ for different wind speeds and no load on the jib.
Figure 12. Overturning moment characteristics obtained from CFD compared with calculated overturning moment and stability moment according to Eurocode as a function of jib rotation angle θ for different wind speeds and no load on the jib.
Applsci 14 04694 g012
Figure 13. Overturning moment characteristics obtained from CFD compared with calculated overturning moment and stability moment according to Eurocode as a function of jib rotation angle θ for different wind speeds and maximum jib load.
Figure 13. Overturning moment characteristics obtained from CFD compared with calculated overturning moment and stability moment according to Eurocode as a function of jib rotation angle θ for different wind speeds and maximum jib load.
Applsci 14 04694 g013
Figure 14. Characteristics of the overturning moment obtained from CFD for selected jib rotation angles θ (considered unsafe) compared with the calculated overturning and stability moment according to Eurocode for angle θ = 0° but for Aref equal to the maximum rectangular projection area for angle 90: (a) θ = +/−15°, maximum load on the jib and wind speed v = 10–27.5 m/s; (b) θ = +/−45°, no load on the jib and wind speed v = 15–50 m/s.
Figure 14. Characteristics of the overturning moment obtained from CFD for selected jib rotation angles θ (considered unsafe) compared with the calculated overturning and stability moment according to Eurocode for angle θ = 0° but for Aref equal to the maximum rectangular projection area for angle 90: (a) θ = +/−15°, maximum load on the jib and wind speed v = 10–27.5 m/s; (b) θ = +/−45°, no load on the jib and wind speed v = 15–50 m/s.
Applsci 14 04694 g014
Figure 15. Trace of the center of gravity for the case where there is a maximum load on the boom Q1 = 3050 kg at 21 m from the crane tower. Numbers 1 to 4 are the tipping edges of the crane, considering the actual support spacing of 4.2 m. The wind direction was assumed to be from left to right.
Figure 15. Trace of the center of gravity for the case where there is a maximum load on the boom Q1 = 3050 kg at 21 m from the crane tower. Numbers 1 to 4 are the tipping edges of the crane, considering the actual support spacing of 4.2 m. The wind direction was assumed to be from left to right.
Applsci 14 04694 g015
Figure 16. Trace of the center of gravity for the case where there is no load on the boom Q1 = 0 kg at 21 m from the crane tower. Numbers 1 to 4 are the tipping edges of the crane, considering the actual support spacing of 4.2 m. The wind direction was assumed to be from left to right.
Figure 16. Trace of the center of gravity for the case where there is no load on the boom Q1 = 0 kg at 21 m from the crane tower. Numbers 1 to 4 are the tipping edges of the crane, considering the actual support spacing of 4.2 m. The wind direction was assumed to be from left to right.
Applsci 14 04694 g016
Table 1. The total number of nodes, faces, and cells.
Table 1. The total number of nodes, faces, and cells.
Cell Size [m]Faces 1NodesCells
0.030750,9222,290,76212,676,871
0.028852,2162,593,29914,342,502
0.0201,705,8884,813,38226,393,269
1 The number of faces generated on the crane surface.
Table 2. Results of the converged test.
Table 2. Results of the converged test.
Cell Size [m]FX [N]FY [N]Mtip [Nm]Number of Iter.
0.0302227.649−43.41161,278.52591
0.0282221.892−34.00061,043.12287
0.0202196.509−31.65260,187.75651
Table 3. The values of aerodynamic coefficients for urban terrain.
Table 3. The values of aerodynamic coefficients for urban terrain.
Urban Terrain
vref = 7.03210 [m/s]
Village Terrain
vref = 6.62928 [m/s]
Open Terrain
vref = 6.340482 [m/s]
Angle α [°]CXCYCMCXCYCMCXCYCM
01.5170.0000.9001.5110.0020.8311.5300.030.762
152.077−0.0491.3701.990−0.0291.2291.939−0.0051.095
302.520−0.2951.7332.395−0.2461.5512.319−0.2021.385
452.956−0.4522.1362.780−0.3931.8962.636−0.3431.664
603.247−0.4582.4633.007−0.3982.1612.827−0.3521.888
753.272−0.3412.5803.004−0.3082.2522.774−0.2811.944
903.147−0.0482.5062.856−0.0742.1782.613−0.1111.870
Aref = 23.31 m2, Bref = 34.5, m, ρ = 1.225 kg/m3.
Table 4. Selected geometrical dimensions of the Liebherr 63 K crane and the values of moments used for calculations. The maximum load weight at a height of 21 m is 3050 kg [41].
Table 4. Selected geometrical dimensions of the Liebherr 63 K crane and the values of moments used for calculations. The maximum load weight at a height of 21 m is 3050 kg [41].
Overturning Moments from
Crane Weights Only
ΣMO = 1228.90 kNm
Stability Moments from Crane Weights Only
ΣMS = 1380.85 kNm
JibMax Load TowerSupport Base of TowerCounterweight
load
G1G2G3G4G5Q1 G6G7G8
130330144014402003050kg7600766026,000
1275323714,12614,126196229,921N74,55675,145255,060
distance from the tipping line
r1r2r3r4r5rQ5 r6r7r8
39332262121m0.872.104.54
moments
M1M2M3M4M5MQ5 M6M7M8
49.17106.75313.5488.5241.29629.63kNm64.67157.801158.38
Table 5. Turbulent flow [36].
Table 5. Turbulent flow [36].
iv(i) [m/s]h(i) [m]Aref [m2]J Jib/
T Tower
WCx0(i)MWCx0(hi)WCx90(i)MWCx90(hi)
99.7733.818.08J00907.0130,664.68
ΣMWJ0ΣMWJ30,664.68
- CxT90 = 1.92
89.4230.8510.20T1579.0848,717.111534.8247,351.64
78.9226.93 T1416.4238,146.371376.7237,077.18
68.3723.01 T1248.9428,739.891213.9327,934.35
57.7719.09 T1075.6120,535.081045.4719,959.51
47.0915.18 T895.2513,587.66870.1613,206.82
36.2911.25 T704.627927.99684.877705.78
25.307.33 T500.193667.17486.173564.39
13.903.41 T271.23925.31263.63899.37
vref*7.13 ΣMWT162,246.57ΣMWT157,699.04
CxT0 = 2.85 CxT90 = 2.77
Table 6. Laminar flow [36].
Table 6. Laminar flow [36].
iv(i) [m/s]h(i) [m]Aref [m2]J Jib/
T Tower
WCx0(i)MWCx0(hi)WCx90(i)MWCx90(hi)
99.7733.818.08J00628.3221,242.64
ΣMWJ0ΣMWJ21,242.64
- CxT90 = 1.31
89.4230.8510.20T1141.3735,212.871169.1236,069.18
78.9226.93 T1023.7927,572.301048.6928,242.81
68.3723.01 T902.7320,773.27924.6921,278.44
57.7719.09 T777.4614,842.81796.3615,203.76
47.0915.18 T647.099821.20662.8310,060.03
36.2911.25 T509.305730.37521.685869.72
25.307.33 T361.542650.64370.332715.10
13.903.41 T196.05668.81200.81685.08
vref*7.13 ΣMWT117,272.28ΣMWT120,124.12
CxT0 = 2.06 CxT90 = 2.11
Table 7. The maximum wind speed v for which overturning moments cause loss of stability of the 63 K crane.
Table 7. The maximum wind speed v for which overturning moments cause loss of stability of the 63 K crane.
v 
[m/s]
With Max Load on jib [kNm]v 
[m/s]
Without Max Load on jib [kNm]
h1 = 33.8 mMO ECode 90°MO CFD 63K +/−15°h1 = 33.8 mMO ECode 90°MO CFD 63K +/−45°
14.88-1380.6728.28-1380.81
23.391380.67-48.571380.52-
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.

Share and Cite

MDPI and ACS Style

Augustyn, M.; Barski, M. Numerical and Analytical Estimation of the Wind Speed Causing Overturning of the Fast-Erecting Crane—Part II. Appl. Sci. 2024, 14, 4694. https://doi.org/10.3390/app14114694

AMA Style

Augustyn M, Barski M. Numerical and Analytical Estimation of the Wind Speed Causing Overturning of the Fast-Erecting Crane—Part II. Applied Sciences. 2024; 14(11):4694. https://doi.org/10.3390/app14114694

Chicago/Turabian Style

Augustyn, Marcin, and Marek Barski. 2024. "Numerical and Analytical Estimation of the Wind Speed Causing Overturning of the Fast-Erecting Crane—Part II" Applied Sciences 14, no. 11: 4694. https://doi.org/10.3390/app14114694

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop