You are on page 1of 16

Competitive Assessment of Ice and Frozen Silt Mat for

Crane Ground Support Using Finite-Element Analysis


Ghulam Muhammad Ali, S.M.ASCE 1; Joe Kosa 2; Ahmed Bouferguene 3; and
Mohamed Al-Hussein, M.ASCE 4
Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Abstract: High capacity cranes are the backbone of the heavy construction industry, which, over the last few years, has embraced mod-
ularization. Consequently, ground stability has become a critical issue for their safe utilization. In practice, ground stability includes laying
several thicknesses of adequately compacted construction aggregates and one or more layers of timber/steel mats on top. In the present study,
a novel alternative is explored whereby an artificially created layer of ice or frozen silt constitutes the base upon which timber or steel mats can
be stacked for ancillary crane support. However, to understand the challenges and feasibility of the proposed technology, a theoretical study
using finite-element analysis (FEA) is carried out in order to gain insight into the factors that can affect the structural behavior of ice/frozen silt
and their comparison with commonly used mat materials, timber (Coastal Douglas fir), and steel (G40.21-44W). The comparison is built
using mechanical properties under identical boundary conditions. The results show that the performance of frozen silt is on par with that of
Coastal Douglas fir. A preliminary cost comparison is also established to develop the value proposition of using frozen silt as a crane mat.
DOI: 10.1061/(ASCE)CO.1943-7862.0002046. © 2021 American Society of Civil Engineers.
Author keywords: Ground bearing pressure; Finite-element analysis (FEA); Crawler crane; Hydraulic crane; Crane mats; Frozen silt; Ice.

Introduction construction sites, the first order of business is to ensure that the soil
bearing capacity of the ground can accommodate the pressure ex-
The modern heavy construction industry is increasingly adopting erted by the crane due to the compounded weights of the crane and
modularization as its primary design and project delivery paradigm. payload. Beavers et al. (2006) conducted a study on crane-related
The projects are broken down into elements (known as modules) accidents on construction sites in the United States, and according
that are fabricated independently in a controlled environment, to the research, per Occupational Safety and Health Administration
leaving a single activity to be carried out outdoors, namely, the (OSHA) data, the fatality rate of the construction industry (US) in
assembly of these modules. Over time, designing and fabricating 2001 reached 13.3 fatalities per 100,000 construction workers. In
modules has become more efficient and thus has encouraged en- addition, the same research also stated that between 1997 and 2003,
gineers to front-load these modules with an increasing number approximately 84% of crane and derrick fatalities were of the work-
of functionalities so to minimize onsite work. However, the draw- ers involved in crane operations (such as crane operators, signal
back of maximizing the functionality of a module is the resulting men, and riggers). A crane tip-over accounted for approximately
weight of the modules, which, a few decades ago, measured a few 11% of crane accidents connected with onsite mobile crane oper-
tons and has jumped in recent years to tens if not hundreds of tons. ation, with the major cost driver being poor ground support
With the emergence of this new weight constraint, a new generation (Beavers et al. 2006). To prevent crane accidents related to ground
of lifting equipment characterized by a high lifting capacity has failure, systematic education systems for those working with cranes
been developed. However, this higher capacity is synonymous with also need to be developed (Lee et al. 2020). In practice, the design
increased structural complexity and heavier weights. Furthermore, of ground support for mobile crane stability requires (1) the evalu-
the surge in demand for crane work has caused an increase in the ation of the ground bearing pressure (GBP) (Pa) under the tracks of
number of crane accidents, which in return has increased the num- the crawler crane or pad load (PL) (kN) under the outriggers in the
ber of fatalities. To ensure that such heavy cranes operate safely on case of a hydraulic truck crane; and (2) the design of one or more
layers, usually in the form of mats, to maintain an overall pressure
1
Ph.D. Candidate, Dept. of Civil and Environmental Engineering, exerted by crane tracks/outriggers below the soil bearing capacity.
Univ. of Alberta, Edmonton, AB, Canada T6G 1H9 (corresponding author).
Traditionally, the calculation of GBP/PL under crane tracks/
ORCID: https://orcid.org/0000-0002-1335-6191. Email: gmali@ualberta.ca
2
Engineering Manager, NCSG Crane & Heavy Haul Corporation,
outriggers is carried out according to a procedure in which the
28765 Acheson Rd., Acheson, AB, Canada T7X 6A8. Email: joe.kosa@ weights of the crane-payload system components are assumed to
ncsg.com be static forces that generate pressure on the ground. The carrier
3
Professor, Campus Saint-Jean, Univ. of Alberta, Edmonton, AB, has a fixed center of gravity (COG), but the superstructure COG
Canada T6C 4G9. Email: ahmed.bouferguene@ualberta.ca revolves based on the lifting radius and the angle of the boom.
4
Professor, Dept. of Civil and Environmental Engineering, Univ. of Shapiro and Shapiro (2010) present in-depth, detailed calculations
Alberta, Edmonton, AB, Canada T6G 1H9. ORCID: https://orcid.org/0000 for GBP and PL covering all crane weights and their COGs. It is
-0002-1774-9718. Email: mohameda@ualberta.ca
important to note that although these applications have proven to be
Note. This manuscript was submitted on April 9, 2020; approved on
December 8, 2020; published online on March 19, 2021. Discussion period useful to heavy construction practitioners, their output is often lim-
open until August 19, 2021; separate discussions must be submitted for ited to pressure values that are calculated at specific points selected
individual papers. This paper is part of the Journal of Construction En- beforehand by the user (Al-Hussein et al. 2000; Hasan et al. 2010;
gineering and Management, © ASCE, ISSN 0733-9364. Shapiro and Shapiro 2010; Wu et al. 2011). As a result, using an

© ASCE 04021038-1 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


alternative procedure built upon finite-element analysis (FEA) not because, on average, the daily maximum temperature is below 0°C
only allows the pressure to be mapped in a quasi-continuous for most of the days in winter (December–February) (Environment
manner, but more importantly, it provides a means to check the and Climate Change 2019). Two of the many planned future oil and
accuracy of previously developed calculation tools. In the present gas projects in that area are (1) the Meadow Creek East SAGD
research, the calculation of the GBP under the crawlers (or PL Project (total estimated value: $1.5 billion; scheduled for 2019–
under outriggers) is carried out using FEA with the aim of under- 2023), and (2) Aspen Oil Sands Project (total estimated value:
standing the geometric conditions under which four different $4.0 billion; scheduled for 2018–2022) (Alberta Government
supports, i.e., ice, frozen silt, Coastal Douglas fir, and steel 2019).
(G40.21-44W), exhibit similar structural behavior. In practical Another cold region in North America for potential application
terms, there exists a significant gap between mat production and of this technology is the state of Alaska, where many regions
mat demand, and the market demand for mats continues to outpace exhibit similar temperatures to those in northern Canada. For in-
Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

mat production (Ali et al. 2019a, b). stance, in Anchorage (61.2167N, 149.9000W), the average mini-
As a result, the aim of this research is to propose an alternative mum ambient temperature, based on historical data spanning the
(or a complementary) solution for traditional crane ground support period 1952–2005, is below −10°C in the winter months of
used for heavy lifting projects. The proposed technology uses ar- December, January, and February, and the temperature is below
tificially frozen material, i.e., ice or frozen silt, to improve the 0°C most of the time during the months of November through
ground bearing capacity, which in turn will help reduce the number February (Western Regional Climate Center 2020). One example
of layers of traditional mats required. Although ground freezing of a potential application is the $2 billion Port of Alaska upgrade
may not seem economically viable because of its requirements project, specifically the petroleum and cement terminal facility
in terms of energy, artificial freezing technology can be suitable (Slowey 2020).
in cold regions, such as the northwestern region of Canada, where Although AGF as a stabilization technology for high-capacity
the ambient temperature in winter is already close to freezing. For cranes is explored in this study in the context of northern Canada
instance, the Encana Cutbank Ridge project, in Dawson Creek, (and other regions of North America with similar climates), where
British Columbia, Canada (N55.7596, W120.2377), was a fast- low ambient temperatures (for extended periods) can aid the freez-
tracked project with 120 major heavy lifts for which the Manitowoc ing process, this technology can also be applied in regions with
18000 and Liebherr LR1300 cranes were used. The project was warmer temperatures. However, for these cases, the cost of freezing
completed ahead of schedule, and the plant was operational in the ground and maintaining it at the targeted temperature is ex-
2018 (Newton 2017). Most of the lifts were performed in winter pected to be costly. As a result, this factor is likely to steer
(November, December, January, February, and March), as shown practitioners toward traditional technologies that can be more
in Fig. 1. For this project, 3,500 timber mats (4 ft × 20 ft × 12 in:) cost-effective. At this juncture, to be a viable solution for industry,
were used for ground stability under crawler cranes. According to the proposed alternative (ice and frozen silt) needs to satisfy at least
metrological data from Environment Canada for the period 1981– two conditions: (1) the materials must have structural viability that
2010, on average, approximately 13.5 days in November are can ensure safety; and (2) the technology needs to be cost-effective.
below freezing (i.e., the daily maximum temperature is below 0° This research focuses mainly on the first aspect in the context of
C). Similarly, based on the same weather data, 18.8 days in Decem- which structural properties of ice and frozen silt are modeled and
ber, 20 days in January, 16.4 days in February, and 11.8 days in compared with those of Douglas fir and steel (G40.21-44W).
March are below freezing, which means those days may be suitable Subsequently, in this contribution, a preliminary cost comparison
for the use of frozen silt or ice mat (Environment and Climate to estimate the cost-effectiveness of this new technology, provided
Change 2019). that the ambient temperature satisfies the requirement for the
Furthermore, the city of Fort McMurray (N56.7264, freezing process, is presented.
W111.3803), Canada, which has seen major heavy construction To build on the findings reported in this study, the research pro-
projects over the last few years, could be an excellent candidate ceeds according to the following roadmap. First, the FEA model of

Fig. 1. Winter crane lifts at Encana Cutbank Ridge Project, Dawson Creek, Canada. (Images by authors.)

© ASCE 04021038-2 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


the crane, i.e., Manitowoc 18000 (crawler) and Grove GMK 7550 heaviest loads because some of the crawler cranes have a capacity
(hydraulic) are validated for GBP/PL by comparing the FEA out- exceeding 2,000 metric tons. Despite the complexity of the tax-
puts for GBP/PL with those provided by the manufacturer’s soft- onomy tree shown in Fig. 2, mobile cranes are typically composed
ware. The goal of this comparison is to ensure that the parameters of three major elements: (1) the boom; (2) the superstructure (or
of FEA crane models, e.g., the weight of each element and the lo- upper structure), which rotates during the lifting process; and
cation of the center of gravity, are all sound. The next step is to (3) the carrier (or chassis), which includes the traction system
model the dynamic ground pressure for each of the selected cranes (wheels or tracks) that is in direct contact with the ground through
and support materials as the superstructure is rotating under similar the tracks or outriggers. The latter components are at the interface
boundary conditions (such as ground properties, payload, crane between the ground and the crane-payload system, which is the
configuration, crane rotation, crane part weights and COGs, and area where the forces and reactions are exerted. The properties
mat placement) as used for the GBP and PL calculations. From this of these forces, i.e., magnitude and directions, are investigated
Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

analysis, the thickness of a slab of ice and frozen silt is designed in by means of FEA, which allows the determination of the lower
such a way as to achieve a similar structural behavior (stress and bounds of the GBP required for a safe lifting process. In the liter-
deflection as the main parameters) to that of Coastal Douglas fir ature that examined applications of FEA to the modeling of crane-
and steel (G40.21) with standard dimensions. Subsequently, based related issues, the emphasis has primarily been put on the determi-
on the results of frozen silt and timber mats, the cost comparison nation of the conditions under which boom buckling occurs (Anlin
for the use of frozen silt instead of timber mart is developed. The et al. 2012; Huang and Ji 2013; Marquez et al. 2014; Yao et al.
present study then includes a conclusion that mainly depends on 2015; Zdravković et al. 2019). Although FEA can be applied to
the mechanical properties and a preliminary cost comparison. crane buckling under any loading condition, i.e., static or dynamic,
Finally, the recommendations and future research opportunities FEA is particularly suitable and essential in the context of dynamic
are presented. loading because it makes the Newtonian mechanical complexity of
the problem manageable, which explains why this method has been
at the core of many investigations that were devoted to crane boom
Literature Review buckling (Haniszewski 2017).
Regarding FEA modeling for crane operation, in one particular
Based on the constraints of the lifts, mobile cranes are selected from study, Wu et al. (2011) used FEA software as part of a three-
among various types and shapes; however, the taxonomy of high dimensional (3D) crane operation simulator engine, whose output
capacity mobile cranes, as commonly defined by practitioners, re- includes a series of images showing color-coded stresses on lifting
lies on two major attributes: (1) the type of boom, i.e., telescopic lugs/trunnions. This study provided a framework for analyzing the
(also known as hydraulic) or lattice; and (2) the type of traction, impact of dynamic loading on the failure of these components (Wu
i.e., wheels or tracks (Fig. 2). It is important to note that, world- et al. 2011). Interestingly, while the boom buckling problem has
wide, crawler cranes have always performed the lifts with the received a considerable amount of attention from researchers and

Fig. 2. Taxonomy of cranes.

© ASCE 04021038-3 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


practitioners, theoretical studies and simulations regarding GBP are Another interesting study was conducted by Lin et al. (2017)
numbered—some examples of which include research authored by to determine the distribution of mobile crane loads under the out-
Anlin et al. (2012), Huang and Ji (2013), and Yao et al. (2015). rigger; the study found that area preparation is required to place
The traditional method for calculating GBP under the tracks the crane at the lifting location, wherein a rule of thumb calculation
(or the outriggers) of a crawler (or wheeled) crane is based on is carried out due to the nonavailability of any actual load distri-
the fundamentals of statics in which the load of the crane-payload bution under the hydraulic mat. Furthermore, the investigation
system, expressed as static forces, is transmitted to the ground was performed using a 2D axisymmetric model through FEA.
through the area of the mat layer(s) with which it is in direct con- The outcome of their study indicates that the traditional method
tact. Under static conditions, the GBP calculations used to design of calculating GBP under the crane mat can be used considering
the mat support for cranes are often conducted by means of the the outrigger mat as rigid with an increment of 10% as a safety
equations given by Shapiro and Shapiro (2010) or, alternatively, factor (Lin et al. 2017).
Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

by computer implementation of these equations available from In the context of timber mats, Mahamid et al. (2017) experimen-
crane manufacturers. In this respect, David Duerr (2010) presented tally addressed the stability of cross-laminated timber mats (built
the industry practice for the selection of timber mats. Hasan et al. using southern yellow pine) under various loads and of various di-
(2010) also included the same design criteria to develop the re- mensions. The specimens were modeled in ANSYS with similar
quired mat selection criterion based on the maximum pressure ex- properties. The graphical representation of the results indicated
erted by the crane. David Duerr (2010) and Hasan et al. (2010) used that the finite-element simulation results were similar to the actual
the two most commonly used techniques for the design of mats lab test results (Mahamid et al. 2017). Therefore, the research
(timber/steel): (1) the first depends on the mat length, which is presented in the present study uses the mechanical properties of
based on the allowable soil bearing capacity, and (2) the second various crane mat materials under similar boundary conditions. The
depends on the strength of the mat under loading. These approaches comparison in this paper provides the competitive position of ice
have been adapted by many researchers to determine the GBP and frozen silt with respect to other mat materials case by case,
under crane mats, not only for crawler cranes but also for hydraulic rather than calculating them individually using complex mat stabil-
cranes (PL calculations with slight modification) (Romanello 2020; ity equations and material properties.
Shahnavaz et al. 2020; Yang et al. 2019). However, a more chal- The major focus of the present contribution is the use of frozen
lenging issue when designing crane support (i.e., mats) to satisfy silt or ice as a crane support, an application that cannot be properly
the GBP constraint is the quantification of the effect of dynamic understood without taking into consideration soil mechanics. In
loading associated with abrupt changes in the rotation speed of this regard, previous research has been carried out in which a
the payload or by wind gusts acting on GBP (Korytov et al. numerical model is developed in ANSYS in order to investigate
the performance of foundation under loading, with the authors
2020; Yang et al. 2019). A simplified method using linear optimi-
of the study observing that tensile strength plays an important role
zation was formulated to simulate the effect of dynamic loading
in determining the strength of the soil. The ANSYS values were
(Zaretsky and Shapiro 1997). Although the traditional approach
compared with the experimental values, with the results showing
is considered common knowledge among crane practitioners, its
that ANSYS can be effective in simulating the behavior of soil
use for developing the equations describing static or dynamic load-
under footing (Elshesheny et al. 2015).
ing requires simplifying the mechanical system under study in or-
It should be noted that the seasonality of frozen ground also
der to keep the mathematical efforts manageable. As a result, to
influences the stability of a structure. In this regard, a recent study
avoid the risk of oversimplifying the mechanical system of interest
simulated the impact of ground freezing on the mechanical behavior
(or using unjustifiable simplifications), analysts often favor the ap-
of bridge pile foundations. In particular, a finite-element model was
plication of FEA, which not only can expedite modeling the propa- developed to observe the impact of frozen silt on the bridge piles
gation of stress due to loading in complex mechanical machines during lateral loading and found favorable agreement with the field
but, more importantly, can map (and quantify) their effect in a con- data (Wenping et al. 2019). In a similar study, the impact of season-
tinuous manner. While FEA is commonly used in the context of ality of frozen ground on the Harbin–Dalian Passenger Dedicated
boom buckling, its application as an alternative (or complement) Line railway (China) was observed at different soil layers (24 loca-
to the traditional statics-based method for investigating GBP ap- tions) (Miao et al. 2020). The authors of this study came to the
pears to be only marginally addressed in the literature devoted conclusion that the most frost heave occurs in the top (well-graded)
to studying this problem in connection to crane stability (Anlin gravel layer, representing 66% of the total deformation.
et al. 2012; Huang and Ji 2013; Marquez et al. 2014; Yao et al. One of the main uses of frozen soil is the construction of winter
2015). The work of Liu et al. (2008) can be considered an excep- roads in North America and Eurasia. For example, in Canada, the
tion. According to their study, the soil type and the configuration of road system in the northern territories nearly doubles in winter due
the crane and mat determine the lower limit of the soil capacity to the addition of seasonal roads that are only available in winter
needed for the safe use of the crane. Interestingly, calculations (Sladen et al. 2020). In considering the possibility of using frozen
based on the equations for soil bearing capacity provide a con- silt for a similar application, the present study proposes a novel
servative design criterion (Meyerhof 1956; Terzaghi Terzaghi approach using frozen silt for the purpose of crane matting.
1943; Vesić 1973). Liu et al. (2008) performed computer simula-
tion using the two-dimensional (2D) finite-difference method and
subsequently compared the simulation with field studies using Methodology and Case Examples
crane and soil deflection under loading. It was determined that us-
ing the standard factor of safety (FS) of 2 or 3 for the traditional Because the objective of this research is to design a slab of ice or silt
calculations of soil bearing capacity can be misleading in the case to be used as part of the ground support system (which may include
of crane use due to load variation and mat configuration. A new timber or steel mats) for a high capacity crane, the first part of the
method was proposed to calculate the soil bearing capacity methodology involves determining the dynamics of the load trans-
under crawler tracks based on the soil type and crane and mat fer from the crane and payload to the supporting layers. This quan-
configurations (Liu et al. 2008). titative analysis is then used to determine the parameters of the

© ASCE 04021038-4 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Parameters used for (a) crawler crane GBP (Pa); and (b) hydraulic crane PL (kN) calculations.

frozen layer (ice or silt) such that it has strength similar to a layer superstructure rotational axis; θcj = angle (degrees) of respective
of timber or steel mats of nominal sizes (e.g., 3.65 m × 2.44 m × part COG with the x-axis; W l = weight (kN) of the payload; Rl =
203 mm). The strength of the frozen layer is referred to in terms of crane radius (m); and θl = crane superstructure slew angle (degrees).
its timber mat equivalent. In the next section, the mathematical All these variables are shown in Fig. 3.
models addresses two elements: (1) the GBP (under crawler cranes) The variables in Eqs. (2) and (3), Ry and Rx , are calculated using
and the PL (under hydraulic cranes); and (2) the linear and non- Eqs. (4) and (5)
linear stress analysis for crane mats under identical boundary  n
conditions. 1 X
Ry ¼ W R sinðθsi þ θl Þ
W i¼1 si si
Xm 
GBP under Crawler Crane Tracks and PLs under þ W cj Rcj sin θcj þ W l Rl sin θl ð4Þ
Hydraulic Crane Outriggers j¼1
To operate mobile cranes safely, it is paramount to have a clear  n
understanding of the GBP. Traditionally, the calculations of the 1 X
Rx ¼ W R cosðθsi þ θl Þ
GBP under the supporting components of a crawler crane and of W i¼1 si si
the PL under hydraulic cranes are conducted by applying funda-
X
m 
mentals of statics, as given in the book Cranes and Derricks, þ W cj Rcj cos θcj þ W l Rl cos θl ð5Þ
4th edition, by Shapiro and Shapiro (2010). To calculate the j¼1
GBP under crawler tracks and the PL under hydraulic crane out-
riggers, the first step involves calculating the sum weight (kN) of Based on the preceding equations, the GBP under the crawler
the crane, including the payload and the location of the center of crane tracks and the PL under hydraulic crane outriggers can be
gravity (COG) using Eqs. (1)–(3), as shown subsequently calculated using Eqs. (6) and (7), respectively, in a unified matrix
X X    
n m
PLF PRF 4Gc νλ ð1 − νÞλ ð1 − λÞν
W¼ W si þ W cj þ W l ð1Þ ¼
i¼1 j¼1 PLR PRR aLe Be ð1 − νÞλ νλ ð1 − νÞð1 − λÞ
2 3
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ðabÞð2 − aÞ ðacÞð2 − aÞ
R¼ R2y þ R2x ð2Þ 6 7
× 4 ðabÞða − 1Þ ðacÞða − 1Þ 5 ð6Þ
b c
Ry
θ¼ tan−1 ð3Þ     
Rx FRR FLR ν 1−ν bð1 þ 2dÞ cð1 þ 2dÞ
¼ Gc
where W si = weight (kN) of n parts in crane superstructure; Rsi = FRF PLF 1−ν ν bð1 − 2dÞ cð1 − 2dÞ
distance (m) between respective part COG and the superstructure ð7Þ
rotational axis; θsi = angle (degrees) of respective part COG with
the x-axis when θl ¼ 0°; W cj = weight (kN) of m parts in the crane where PLF =PLR = GBP under the left front/rear track of crawler
carrier; Rcj = distance (m) between respective part COG and the crane; PRF =PRR = GBP under the right front/rear track of crawler

© ASCE 04021038-5 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


crane; FRR =FRF = PL on the right rear/front outrigger of the hy- directly under the crane track, based on the dimensions of the track,
draulic crane; and FLR =FLF = PL on left rear/front outrigger of a and are further distributed using crane mats (Al-Hussein et al. 2005;
hydraulic crane. The factors in Eqs. (6) and (7) are defined by Hasan et al. 2010; Shapiro and Shapiro 2010). For comparison pur-
Eqs. (8)–(16) poses, for hydraulic cranes, the PL is taken into consideration for
  mat selection, and for crawler cranes, the GBP values under the
1 j cosð270 þ θÞj track are taken into consideration for mat selection. Moreover,
μ¼ þ1 ð8Þ
2 cosð270 þ θÞ in the context of crawler cranes, the effective/bearing length and
  width of the crane track (contact area) is smaller than the actual
1 j cos θj physical length and the width of the crawler tracks, as shown in
ν¼ þ1 ð9Þ
2 cos θ Fig. 5 (Al-Hussein et al. 2005; Hasan et al. 2010; Shapiro and
Shapiro 2010).
Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

r ¼ Rj sin θj ð10Þ Traditionally, for crawler cranes, the crawler crane track width-
wise pressure distribution under the track of the crane is considered
W uniform (Shapiro and Shapiro 2010). However, in reality, the
Gc ¼ ð11Þ
4s widthwise pressure distribution diagram is trapezoidal/triangular
  in nature (Fig. 5). For this reason, FEA provides two values for
1 jLe − 6Rj cos θjj each side of the track—one maximum and one minimum. The aver-
λ¼ þ1 ð12Þ
2 Le − 6Rj cos θj age of these values is taken for comparison with those obtained
from manual calculations and with those obtained from the crane
3 manufacturers, e.g., Manitowoc 18000 and Grove GMK 7550
a¼ ðL − 2Rj cos θjÞ ð13Þ (Al-Hussein et al. 2005, 2000; Grove 2019; Hasan et al. 2010;
2Le e
Liu et al. 2008; Manitowoc 2019; Shapiro and Shapiro 2010).
b ¼ s − r þ 2μr ð14Þ In the case of the crawler crane Manitowoc 18000, the results
of the comparative analysis are obtained for three different crane
c ¼ s þ r − 2μr ð15Þ configurations (summarized in Tables 1–3). The variation of these
configurations is mainly of the payload and the lifting radius. As
Rj cos θj the lifting radius changes, the COG of the boom also changes. From
d¼ ð16Þ a computer implementation perspective, running the FEA on the
Le
problem at hand requires developing a 3D model of the crawler
where s = distance from the crane center to the center of crawler crane. To keep the complexity of this design within reasonable
width/outrigger; Be = width of the track; and Le = effective track bounds and for the ease of FEA simulation, the crane components
length/distance between two outriggers (lengthwise). are assigned to three major parts. The components of the first part
Alternatively, GBP and PL may be determined by means of FEA include the superstructure, counterweights, crane cabin, boom,
modeling to provide a continuous map, which was the approach mast (if available), superlift-counterweights (if available), and the
used in the present study. To calibrate and to check the validity load. The second part includes the crane carbody, crawler tracks,
of the developed FEA models, the GBP/PL values at specific points and carbody counterweights. Finally, the third part contains steel
from the FEA simulation are compared with the values calculated plates that act as pressure sensors whose function is to capture the
manually by means of the preceding equations. Further to the val- pressure profiles under the track. These plates are of the same area
idation of the data used, the manually calculated values are be com- as the effective/bearing area of the track for the GBP.
pared with those calculated by software provided by the crane Before moving to the FEA simulation, the same geometrical
manufacturer (Grove 2019; Manitowoc 2019). data is used to obtain a manual set of GBP values for the boom
The FEA-based simulations are generated for two types of mo- slew angle in the range 0°–90° (crane front to crane left-side)
bile cranes: (1) a crawler crane (Manitowoc 18000) for GBP in (Al-Hussein et al. 2005; Hasan et al. 2010; Shapiro and Shapiro
which the ground-machine interface consists of two parallel rectan- 2010). The same configuration is used to obtain the GBP values
gular tracks; and (2) a hydraulic crane (Grove GMK 7550) for PL in from the manufacturer’s software, Manitowoc Ground Bearing
which the load transfer occurs at the outriggers (see flow chart Pressure Estimator Version: 1.0.6.0, which is available as freeware
shown in Fig. 4). from Manitowoc’s official website (Manitowoc 2019) as a set of
Following the flowchart (Fig. 4), the finite-element model of values for counterchecking both FEA and manual calculations
each crane is loaded into the FEA software (ANSYS 17.1). These (Tables 1–3 and Fig. 4). It can be noted that the differences between
models can be sketched in the FEA drafting module or can be im- the manual calculations and the values from the Manitowoc soft-
ported from any other 3D source compatible with ANSYS. The ware are minimal. The percentage variation error between the
main inputs required are the weights of the various components manual calculations and values from the Manitowoc software is
of the crane and the exact location of their center of gravity found to be less than 1% for all three cases. Further, the percentage
(COG), which are obtained from the manufacturer of the selected error between the FEA and the manual calculations is found to be
crane. Likewise, the same data are used for the manual calculations less than 1% (Table 4). The GBP values are presented in Fig. 6 for
of GBP/PL under the crane. The GBP values under the crawler the left front of the crawler crane for all three case examples
crane or the PL values for hydraulic cranes are obtained for crane (Manitowoc 2019).
superstructure slew from 0° to 90°. It is worth noting in this study After obtaining the GBP validation using FEA for the crawler
that the traditional procedure of calculating the GBP values under crane FEA model, the next step is to obtain the same validation
the crawler crane varies slightly from that used to calculate the PL for PL using FEA for the hydraulic crane FEA model, Grove
under the hydraulic crane. In the case of a hydraulic crane, the PL GMK7550 [Fig. 11(b)], which is an all-terrain crane. As the re-
under the outriggers are obtained and then distributed through out- sults of crawler cranes are favorable and create confidence in the
rigger mats per the mat surface area to get the GBP values. On the procedure adopted, only one case example for hydraulic cranes
other hand, the GBP values under the crawler crane are taken is performed to validate the PL. The hydraulic crane selected is

© ASCE 04021038-6 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Flowchart for the crawler/hydraulic crane data flow for FEA crane model validation.

configured to have a boom length of 38.13 m (125.1 ft), a boom COGs are not available from the manufacturer, although the PL
configuration of [0-100-100-0], superstructure counterweights of values are available from the online Grove application at various
119,975 kg (264,500 lb), a lifting radius of 19.812 m (65 ft), a angles. By utilizing these PL values against different angles and
lifting load of 40,000 kg (88,185 lb), an outrigger span of 28.7 × the use of GBP equations with weights and COGs as unknown var-
29.2 ft, and a surface operational condition of solid. In the case of iables, the required crane data is obtained (Al-Hussein et al. 2005;
hydraulic cranes, as stated previously, the manufacturer’s software Hasan et al. 2010; Shapiro and Shapiro 2010).
and manual calculations only provide the PL in the form of the Subsequently, the top surface of the mat under the hydraulic
force on each outrigger, so only the PLs are compared. An online crane is utilized as a sensor to obtain the values of PL. The values
application for PL calculations is available from the hydraulic crane from the FEA are plotted against manual calculations as presented
manufacturer (Grove 2019); however, this online application only in Fig. 7(b), and it is found that the variation in the values is
provides outrigger PL at particular angles with increments of 45°. approximately 0.37% (with respect to the largest value), which
To facilitate manual calculations with respect to the online appli- is less than 1%.
cation, the superstructure slew angle is taken from 0° to 360° from
the rear to the right side with increments of 45° [Fig. 7(a)] (Grove
2019). The percentage variation between the manual calculations Linear and Nonlinear Stress Analysis
and the Grove online application (Grove 2019) values are found The GBP and PL crane analyses pave the way for the mat strength
to be approximately 1.66% [Fig. 7(a)]. The reason for this higher study. The analysis conducted in the previous section develops a set
variation is that the weights of crane parts and their respective of boundary conditions that can subsequently be used for the linear

© ASCE 04021038-7 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Table 3. Case examples for Manitowoc 18000
Weight (ton) Lifting radius) Weight (ton) Lifting radius
Description (manual) (manual) (FEA) (FEA)
Case-1 50.000 25.917 49.998 25.917
Case-2 75.000 21.336 74.996 21.358
Case-3 100.000 19.812 99.996 19.812

Table 4. Variation (%) of GBP (Pa) values between manual, Manitowoc,


Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

and FEA calculation (Manitowoc 18000)


Description Case-1 Case-2 Case-3
Variation (%) between manual and 0.054 0.049 0.036
Manitowoc software calculations
Variation (%) between manual and 0.518 0.506 0.473
FEA calculations

simulation is to display the results. The results can be displayed


in many ways, including normal stresses, principal stresses, forces,
Fig. 5. Effective/bearing: (a) length; (b) width; (c) traditional; and
energy, deformation, and so forth, along with the crane superstruc-
(d) FEA load distribution. (Images by authors.)
ture slew. For the material comparison in this research, only the
normal stress (Pa) distribution (negative due to compression) under
the mat and the mat deflection (mm) are considered. First, the mats
Table 1. Manitowoc 18000 crane configuration
under the crawler crane (Manitowoc 18000) are analyzed, and sub-
sequently, the mats under the hydraulic crane (GMK 7550) are
Crane component Configuration analyzed.
Boom 85.344 m (280 ft) #55 OR #55A As mentioned previously, for the mat strength analysis, four mat
Carbody counterweight 145,150 kg (320,000 lb) materials (G40.21-44W, Coastal Douglas fir, ice, and frozen silt)
Superstructure counterweight 239,500 kg (528,000 lb) are selected. For each of these materials, five linear mechanical
properties and one nonlinear mechanical property are selected for
FEA. To build the comparison, ice and frozen silt are considered as
proposed materials for use as crane mats. The temperature of the
and nonlinear stress analyses. Following the same boundary con- frozen silt and ice is −10°C for analysis purposes (Andersland et al.
ditions as verified in the previous section, the idea is to compare the 2003). The mechanical properties of the frozen silt mat under in-
behavior of different crane mat materials rather than calculating the vestigation are taken from the research done by Yang et al. (2015)
values and further verifying them with manual calculations and lab on specimens composed of silt from the Campbell Creek Bridge
testing values. This decreases the requirement for the verification of case project in Anchorage, Alaska (seasonally frozen soil), and
results, as the timber mat can serve as the benchmark for the frozen the CRREL Permafrost Tunnel in Fox, Alaska (permafrost, frozen
silt. Because the value proposition is to replace timber mats with for two consecutive seasons). The values of the mechanical proper-
frozen silt in winter, the stress and deflection variation with respect ties of these materials are presented in Table 5.
to the timber mat is observed. For the mat strength and deformation To simplify the analysis, all of the mat models developed for
analysis, the overall simulation process is divided into six steps, as FEA simulation are carried out, assuming that the mechanical prop-
shown in Fig. 8. The main concern for the FEA simulation is the erties of the four materials are isotropic in nature. The mat model is
accuracy of the crane model and the output results under the same placed under the crane track/outrigger. There are various sizes of
boundary conditions. The crane GBP calculations shown in the crane mats available in the market, and most of the time, they are
“GBP” section are used to determine the accuracy of the FEA crane custom-made (Matrax 2020). The selection of mats size is deter-
model, which shows that the pressure that the FEA crane model is mined, based on the GBP and the soil bearing capacity, as demon-
exerting on the crane mats is realistic. The boundary conditions strated in the literature review. Applying the same selection criteria,
need to be defined to simulate the crane mat bending and the stress for the research described in this paper, the suitable mat size for
distribution under the crane loading. The final step in this the traditional mat materials (Coastal Douglas fir and steel) and

Table 2. Manitowoc 18000 crane parts and respective COGs


Weight (ton) COG distance (m) from Weight COG distance (m) from
Description (manual) CL of rotation (manual) (ton) (FEA) CL of rotation (FEA)
Boom 118.358 11.256 118.360 11.270
Upper 346.351 −5.985 346.350 −5.985
Crawler 249.719 −0.176 249.720 −0.176
Length of bearing crawler (m) 8.839 (manual—center to center) 8.840 (FEA—center to center)
Distance between crawler (m) 8.534 (manual—center to center) 8.534 (FEA—center to center)

© ASCE 04021038-8 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. GBP (Pa) (minimum principal stress) values from Manitowoc computer application, manual calculations, and FEA simulation calculations on
left front crawler crane, Manitowoc 18000.

Fig. 7. (a) PL (kN) values from Grove online application and manual calculations; and (b) PL (kN) values from manual calculations and FEA
simulation under right rear outrigger of the hydraulic crane (Grove GMK7550).

for the frozen silt mat is 12 ft × 8 ft × 8 in., while the suitable size stresses are transferred to the ground through the mat, which results
of the ice mat for the crawler crane is assumed to be 12 ft × 8 ft × in the deflection of the mat and the deflection (settlement) of the
2.62 ft, and for the hydraulic crane, the suitable size of the ice mat ground. Figs. 10(a and b) show the minimum normal stress
is assumed to be 12 ft × 8 ft × 4.3 ft. The thickness of ice mats is (negative due to compression) under the left front mat and the left
different from all other mats due to its low Young’s modulus. The rear mat, respectively.
sizes of the mats can be changed in the future to observe the sim- Besides normal stresses, mat deflection is also a major concern
ulation of FEA with various mat sizes. Along with normal stresses, in the selection of a crane mat. Fig. 11(a) shows the mat deflection
the ground deflection is also considered for the comparative analy- under loading (under similar boundary conditions) under the left
sis, so the ground properties need to be added to simulate the stress front mat. Under loading conditions, the maximum deflection is
distribution. The soil parameters considered for this research are that of the ice mat. The best performer is found to be G40.21-
listed in Table 6 (Yang et al. 2010). The soil under the mats is mod- 44W. An acceptable deflection is 1% per ISO4305:1991(E); how-
eled per the Drucker-Prager theory (Ravishankar and Satyam ever, for this simulation, a 0.75% deflection along the length of the
2013). The model was first presented by D. C. Drucker in 1952 and mat is considered a conservative value (ISO 1991). The results
is based on the elastic–plastic constitutive relationship (Drucker show that all the mat materials fall within an acceptable range ac-
and Prager 1952). It is important to indicate in this study that cording to industry standards. In this mat stress and deflection
for FEA analysis, only one crane configuration for each crawler analysis, the frozen silt and Coastal Douglas fir are comparable
crane and the hydraulic crane is selected. and relatively similar. As the mechanical properties of both are sim-
Following the flowchart for FEA, as shown in Fig. 9(a), a case ilar, this is evidence that frozen silt may behave similarly to timber
example of a Manitowoc 18000 crane (crawler crane) with boom (i.e., Coastal Douglas fir) under loading, provided the conditions
slew angles ranging from 0° to 90° from the front to the left side and are favorable, such as the mat surface temperature, frozen silt com-
with a load of 50,000 kg (110,231.13 lb) at a lifting radius of position, ambient temperature, and soil composition (ground).
25.917 m (85 ft) is considered and simulated. Coastal Douglas fir, Besides considering the normal stress values under mats and
steel mats, and test mats (ice and frozen silt) are placed under mat deflection as mentioned previously, Fig. 12 shows the normal
the respective crawler crane tracks, as shown in Fig. 9(b). During stress variation profile along the length of the mat under left-front
the rotation of the superstructure, the GBP under the crawler track crane tracks. In this regard, Fig. 12 shows how the stresses are
varies with the change in the boom slew angle. With the change transferred from the center of the mat to the edge of the mat.
of the GBP, the stresses developed in the mat also change. These For steel (G40.21-44W), the variation is less in comparison with

© ASCE 04021038-9 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. ANSYS problem-solving sequence (ANSYS Mechanical Workbench 17.1).

Table 5. Linear and nonlinear mechanical properties for FEA


Description Units G40.21-44W Coastal Douglas fir Ice Frozen silt
Young’s modulus MPa 200,000a 13,400b 9.0c 10,000d,e
Poisson’s ratio — 0.3a 0.449a 0.3c 0.3f
Tensile yield strength MPa 350a 2.3c 3.1g 5.1d
Compressive yield strength MPa 152h 5.0b 25c 5.0d
Tensile ultimate strength MPa 450a 5.0b 3.1c 5.1d
Tangent modulus MPa 504a 110a 0 0
a
Data from Green et al. (1999).
b
Data from Kretschmann (2010).
c
Data from Schulson (1999).
d
Data from Yang et al. (2015).
e
Data from Ralph Wilson (1982).
f
Data from Andersland et al. (2003).
g
Data from Petrovic (2003).
h
Data from Boyer and Gall (1985).

other materials, and the stress is uniform over the length of the steel the stress profile moves up and decreases in value along the length
mat. This can be explained by the higher tensile strength, compres- of the mat—like an inverse bell curve with the maximum value in
sive strength, and the value of Young’s modulus of steel in com- the center of the mat. It is noteworthy that the stress profile for
parison to those of Coastal Douglas fir, ice, and frozen silt. Coastal Douglas fir and frozen silt tilts toward the left side of
The stress profiles for Coastal Douglas fir and frozen silt are the crane, which is in accordance with the rotation of the
similar to each other, which is further evidence that the behavior superstructure.
of frozen silt is relatively similar to that of Coastal Douglas fir. The strength of ice is greater when compared to G40.21-44W.
For frozen silt and Coastal Douglas fir, the compressive stresses The main reason is the mat thickness: for all other mats, the
(normal stresses in negative) reach the maximum in the center mat thickness is 8″ (0.2032 m), but to make ice comparable with
of the mat (lengthwise), and as it moves away from the center, respect to deflection and stress distribution to the other materials,

© ASCE 04021038-10 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Table 6. Soil properties for mat investigation the ice mat thickness is 2.62′ (0.8 m). If the thickness of ice is
Property Value Unit 8″ (0.2032 m), the compressive stresses (normal stress in the neg-
ative) would be lower than that of frozen silt mat and Coastal
Density 1,900 kg=m3
Isotropic elasticity
Douglas fir.
Derive from Young’s modulus — Following the same procedure as was followed for the crawler
Young’s modulus 4 × 107 Pa crane stress analysis, the configuration of the hydraulic crane
Poisson’s ratio 0.3 — GMK7550 is provided in the “Ground Bearing Pressure” section.
Bulk modulus 3.33 × 107 Pa For mat strength analysis, the normal stresses under the mat are
Shear modulus 1.538 × 107 Pa obtained by rotating the superstructure of the FEA crane model
Drucker-Prager (using ANSYS), simulating the actual crane operation, creating
Drucker-Prager base — — pressure on the outrigger mats. During the simulation, the boom
Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Uniaxial compressive strength 1 × 105 Pa


slews from the right rear to the right side, and so the stress values
Uniaxial tensile strength 1 × 10−19 Pa
Biaxial compressive strength 1.1 × 105 Pa are taken for the right rear and right front outrigger mats. For the
Failure plane data set-1 right rear outrigger, during the load slew, the stress value increases
Inner friction angle 0.010656 radian in magnitude (compression stress) initially, but decreases later be-
Initial cohesion 10,000 Pa cause the maximum compressive stress value (negative normal
Dilatancy angle 0.0015231 radian stress) occurs when the boom is directly over the outrigger. Beyond
Residual inner friction angle 0.0097478 radian that point, the stress value decreases in magnitude as the load
Residual cohesion 10,000 Pa moves away from the outrigger. Fig. 13 shows the minimum nor-
Source: Data from Yang et al. (2010). mal stress under the right rear and right front outrigger mat.

Fig. 9. (a) Flowchart for crane mat strength analysis; and (b) relative positions of the crane, mat, and soil for analysis.

Fig. 10. Normal minimum stress (Pa) under (a) left front mat; and (b) left rear mat (for Manitowoc 18000).

© ASCE 04021038-11 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. (a) Left front mat deflection (mm) (Manitowoc 18000); and (b) right rear mat (mm) deflection (Grove GMK7550).

Considering the load pattern and mat behavior under the hy- verification for the accuracy of an FEA crane model for mat stress
draulic crane outrigger, the pressure exerted by the outrigger results analysis. The results for the crawler crane GBP values create con-
in the bending of the mat and the deflection (settlement) of the fidence in the use of FEA as an alternative to calculate the GBP
ground under the mat. Fig. 11(b) shows the mat deflection under under the tracks, and the same can be said in the case of the hydraulic
the right rear outrigger; it was found that the conservative industry crane example. It is important to bear in mind that when the part
benchmark is much lower in comparison with all the mat material weights and COGs are equal, it is observed from this research that
specimens considered in the present study (ISO 1991). FEA also provides the same value for GBP/PL as the manufacturer’s
Fig. 14 shows the normal stress variation profile along the width software, and FEA also provides the same values as the manual cal-
of the mat. In contrast to the crawler cranes, the location of the culations (Grove 2019; Manitowoc 2019). If there were a difference
stress profile is taken in the center of the mat because the outrigger in the values, this would indicate that an error exists either in the
exerts maximum pressure in the center of the mat, and the stress software calculations, in the manual calculations, or in the FEA.
profile (line) is taken at the edge of the mat. The stress distribution The results from the GBP/PL verification establish a strong case
profile for Coastal Douglas fir and frozen silt are similar in nature for the use of FEA for crane mat material comparison with the same
and look like inverse bell curves. boundary conditions. Because the benchmark material is timber mat,
In the case of the hydraulic crane, the stress values, the mat de- it is easy to observe the behavior of frozen silt and ice under similar
flection, and the stress profile of ice under the outrigger is also fa- conditions to those when timber mats are used, without going into
vorable in comparison to frozen silt and Coastal Douglas fir. It is detail for actual stress deflection values.
important to mention again that the thickness of the ice mat under From the mechanical property comparison analysis, it is inferred
the hydraulic crane is 4.3 ft (1.3 m) instead of the typical mat thick- that frozen silt mat can be used as an alternative to timber mat for
ness of 8 in. (0.2032 m) due to its low Young’s modulus. crane support provided that the load is fully distributed, and the tem-
perature required for the frozen silt (−10°C or below) is achieved and
Cost Estimation and Comparison for Frozen Silt Mat maintained. Ice can also be used, but due to its low Young’s modulus,
it requires more thickness to compete with the mechanical properties
The next step is to estimate the cost (and, therefore, the practicality)
of Coastal Douglas fir or frozen silt. The mechanical properties of the
of using frozen silt as a crane mat and for the value proposition of
frozen silt mat under investigation are taken from the research done
the frozen mat. For comparison purposes, a hypothetical case sce-
by Ralph Wilson (1982) and Yang et al. (2015). Based on the
nario is assumed for the Cutbank Ridge project in Dawson Creek,
mechanical properties, there is a strong possibility that ice and frozen
British Columbia, Canada. The crane assumption is Manitowoc
silt could be used for crane support, especially in cold regions where
18000 for this hypothetical project. It is a crawler crane requiring
the temperature is favorable to achieve freezing, whether naturally or
eight timber mats (3.6449 × 2.4384 × 0.2032 m)—four for each
by means of artificial ground freezing.
track—or, in the case of using frozen silt mats, two frozen silt mats
Another phenomenon is observed during the FEA simulation of
(14.58 × 2.4384 × 0.2032 m). The cost breakdown is shown in
these mats using ANSYS, wherein in the case of Coastal Douglas fir
Table 7. For the three winter months, it is estimated that a crane mat
and frozen silt, an insufficient mat thickness does not distribute the
company can save approximately $2,276.24. If the surface temper-
GBP completely, and due to the bending of the mat and its deflection,
ature of the frozen silt mat is not at the required level, the AGF can
the edges of the mat leave the ground surface and shows positive
be utilized.
normal stress at the edges. The stress profiles of Coastal Douglas
fir, frozen silt, and ice shown in Fig. 12 indicate that the edges
Conclusion, Recommendations, and Future Work are not taking any load because the stresses at the edges are above
0 Pa (positive). The load distribution is such that the loading length of
The cases presented in the “GBP” section (crawler crane and hy- the mat decreases in comparison to the actual length of the mat. More-
draulic crane case examples) are used to develop a GBP and PL over, the compressive stress profile under the front left mat is tilted

© ASCE 04021038-12 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 12. Normal stress (Pa) variation profile under left front crane mats (for Manitowoc 18000 crane).

Fig. 13. Normal minimum stress (Pa) under (a) right rear; and (b) right front mat (Grove GMK 7550 crane).

© ASCE 04021038-13 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 14. Normal stress (Pa) variation profile under right rear crane mats (Grove GMK 7550).

Table 7. Cost comparison between timber mats and frozen silt mats
Description Units Timber mats (8 mats) Frozen silt (4 mats)
Capital cost for crane mats $ 13,600 0 (naturally frozen ground)
Lifespan of mats Years 3 Not applicable
Capital cost for three winter months $ 1,133 0
Transportation cost (500 km) $ 1,143 0
Mat laying cost $ Assumed to be the same for both
Total cost (savings) $ 2,276 0

toward the load side during crane slew in the case of frozen silt, ice, mats (Andersland et al. 2003; Harris 1994). An estimation of the
and Coastal Douglas fir. It is significant to mention that the compres- cooling needs and costs is required to conduct a comparison be-
sive stress value is at its maximum when the boom is right over the tween traditional matting solutions (steel and timber mats) and fro-
crawler track or outrigger during crane superstructure slew. zen silt or ice mat solution. A complete SWOT (strength, weakness,
The cost comparison (section “Cost Estimation and Comparison opportunity, and threats) analysis would be required to establish a
for Frozen Silt Mat”) gives an overview of the practicability of fro- market study for this new approach (McNutt 1991).
zen silt mat. The crane mat rental company can save capital by us- As the ambient temperature is the critical factor, the present
ing the frozen silt mat in place of timber mats in winter months, study assumes −10°C is the frozen silt/ice mat surface temperature,
provided that the ambient temperature in winter months in the given resulting in the frozen silt mat properties’ being comparable to
region is favorable for the freezing process. those of Coastal Douglas fir; however, the properties change with
This research contribution can provide a starting point for the fluctuations in the frozen silt/ice mat surface temperature. Further
feasibility of using a frozen silt mat for crane support. For practical research is required to define the changes in mat properties with
application, the freezing process needs to be investigated with respect to mat surface temperature.
actual industrial numbers and procedures. In regions where the The results of these FEA simulations for GBP/PL indicate that
temperature is above −10°C, AGF can be used to prepare these the load distribution under the mat is nonuniform in nature. This

© ASCE 04021038-14 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


phenomenon warrants further study to estimate the bending and Int. Conf. on Teaching and Computational Science, Advanced Technol-
stability of crane mats made of frozen silt. ogy in Teaching, edited Y. Wu, 401–406. Berlin: Springer. https://doi
Only a few mechanical properties are compared, but it is ex- .org/10.1007/978-3-642-11276-8_52.
pected that future research can incorporate additional mechanical Beavers, J. E., J. R. Moore, R. Rinehart, and W. R. Schriver. 2006. “Crane-
properties. In the case of frozen silt, cracking and fracking, as major related fatalities in the construction industry.” J. Constr. Eng. Manage.
contributors to eventual structural failure, would need to be inves- 132 (9): 901–910. https://doi.org/10.1061/(ASCE)0733-9364(2006)
132:9(901).
tigated in a lab environment before site testing. Moreover, for
Boyer, H. E., and T. L. Gall. 1985. Metals handbook desk. Edited by J. R.
frozen silt mats, future research can investigate other soil types
Davis. Cleveland, OH: American Society for Metals.
(such as sand, clay, loam, and chalky soil) in order to establish an
David Duerr, P. E. 2010. “Effective bearing length of crane mats.” In
empirical relationship between mat strength, soil composition, Crane & rigging conference, 1–8. Houston: 2DM Associates.
water content, and mat temperature, which could be verified by lab Drucker, D. C., and W. J. Prager. 1952. “Soil mechanics and plastic analysis
Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

and site testing. or limit design.” Q. Appl. Math. 10 (2): 157–165. https://doi.org/10
The US dollar values mentioned in the “Cost Estimation and .1090/qam/48291.
Comparison for Frozen Silt Mat” section are meant to illustrate Elshesheny, A., N. Nagy, and A. Belal. 2015. “Numerical evaluation of
the usefulness of the new technology. Spatial constraints prevent bearing capacity of square footing on geosynthetic reinforced sand.”
us from going into detail about the cost breakdown in this study. In Proc., Int. Conf. on Civil, Structural and Transportation Engineer-
Although a detailed cost breakdown is outside the scope of the ing. Ottawa: International ASET Inc.
present work, a detailed analysis of these results, as well as the Environment and Climate Change. 2019. “1981–2010 climate normals &
assumptions underlying their calculations, will be provided in a fu- averages.” Accessed July 24, 2019. http://climate.weather.gc.ca/climate
ture contribution. Regarding the cost comparison of AGF for frozen _normals/index_e.html.
silt mats, it is also important to mention that the use of AGF for the Green, D. W., J. E. Winandy, and D. E. Kretschmann. 1999. “Mechanical
preparation of frozen silt mats is presented in a previous contribu- properties of wood.” Accessed February 12, 2018. https://www.fpl.fs
tion by the authors (Ali et al. 2019a). The next step in this regard is .fed.us/documnts/fplgtr/fplgtr113/ch04.pdf.
the cost estimation for the preparation of frozen silt mat using AGF, Grove. 2019. “GMK legacy models outrigger pad load calculator.” Ac-
cessed July 24, 2019. https://www.manitowoccranes.com/en/Tools/lift
which involves incorporating all the activities involved in the direct
-planning/Outrigger-Pad-Load-Calculators/All-Terrain.
or indirect AGF process.
Haniszewski, T. 2017. “Modeling the dynamics of cargo lifting process by
overhead crane for dynamic overload factor estimation.” J. Vibroengin-
eering 19 (1): 75–86. https://doi.org/10.21595/jve.2016.17310.
Data Availability Statement Harris, J. S. 1994. Ground freezing in practice. London: Thomas Telford.
Hasan, S., M. Al-Hussein, U. Hermann, and H. Safouhi. 2010. “Interactive
Some or all data, models, or code that support the findings of this and dynamic integrated module for mobile cranes supporting system
study are available from the corresponding author upon reasonable design.” J. Constr. Eng. Manage. 136 (2): 179–186. https://doi.org/10
request. This includes FEA generated data, FEA crane models, .1061/(ASCE)CO.1943-7862.0000121.
FEA boundary conditions, crane details, and so forth. Huang, X. L., and A. M. Ji. 2013. “Analysis of nonlinear local buckling of
crane telescopic boom.” Appl. Mech. Mater. 387: 197–201. https://doi
.org/10.4028/www.scientific.net/AMM.387.197.
References ISO. 1991. Mobile crane: Determination of stability. ISO 4305:1991(E).
Geneva: ISO.
Alberta Government. 2019. “Alberta major projects.” Accessed July 24, 2019. Korytov, M. S., V. S. Shcherbakov, V. V. Titenko, and V. E. Belyakov. 2020.
http://majorprojects.alberta.ca/#map/?sector=Oil-and-Gas,Pipelines&type “Study of the crawler crane stability affected by the length of compen-
=Oil-and-Gas_Gas,Oil-and-Gas_Oil-Sands:-In-Situ,Oil-and-Gas_Oil sating ropes and platform rotation angle in the mode of movement with
-Sands:-Mining,Oil-and-Gas_Upgrader. payload.” J. Phys.: Conf. Ser. 1546: 12135. https://doi.org/10.1088
Al-Hussein, M., S. Alkass, and O. Moselhi. 2000. “D-CRANE: A database %2F1742-6596%2F1546%2F1%2F012135.
system for utilization of cranes.” Can. J. Civ. Eng. 27 (6): 1130–1138. Kretschmann, D. E. 2010. “Mechanical properties of wood.” Accessed
https://doi.org/10.1139/l00-039. February 12, 2018. https://charlespetersonflooring.com/wp-content
Al-Hussein, M., S. Alkass, and O. Moselhi. 2005. “Optimization algorithm /uploads/2014/04/chapter_05.pdf.
for selection and on site location of mobile cranes.” J. Constr. Eng. Lee, J., I. Phillips, and Z. Lynch. 2020. “Causes and prevention of mobile
Manage. 131 (5): 579–590. https://doi.org/10.1061/(ASCE)0733-9364 crane-related accidents in South Korea.” Int. J. Occup. Saf. Ergon.
(2005)131:5(579). 14: 1–10. https://doi.org/10.1080/10803548.2020.1775384.
Ali, G. M., M. Al-Hussein, and A. Bouferguene. 2019a. “Use of finite Lin, M., N. Duong, L. Deng, U. Hermann, T. Zubick, Z. Lei, and S. Adeeb.
element analysis for the estimate of freezing & maintenance phase of 2017. “An investigation of the distribution of mobile crane loads for
indirect & direct artificial ground freezing of proposed frozen silt mat, construction projects.” In CSCE leadership in sustainable infrastruc-
an alternative of timber mat.” In Proc., 2019 Int. Symp. on Automation
ture, 9. Pointe Claire, QC, Canada: Canadian Society of Civil
and Robotics in Construction (ISARC), 846–853. London: The
Engineering.
International Association for Automation and Robotics in Construction.
Liu, X., D. H. Chan, and B. Gerbrandt. 2008. “Bearing capacity of soils for
https://doi.org/10.22260/ISARC2019/0114.
Ali, G. M., M. Al-Hussein, A. Bouferguene, and J. Kosa. 2019b. “Com- crawler cranes.” Can. Geotech. J. 45 (9): 1282–1302. https://doi.org/10
petitive finite element analysis (ANSYS) for the use of ice & frozen silt .1139/T08-056.
as a supporting structural material, an alternative to the traditional Mahamid, M., T. Brindley, N. Triandafilou, and S. Domagala. 2017.
crawler crane mat material (S355, G40.21 & Coastal Douglas-fir).” “Behavior and strength characteristics of cross-laminated timber mats:
In Proc., CSCE Annual Conf. Growing with youth—Croître avec les Experimental and numerical study.” In Structures Congress 2017: Busi-
jeunes, 10. Pointe Claire, QC, Canada: Canadian Society of Civil ness, professional practice, education, research, and disaster manage-
Engineering. ment, 254–268. Reston, VA: ASCE. https://doi.org/10.1061
Andersland, O. B., B. Ladanyi, and ASCE. 2003. Frozen ground engineer- /9780784480427.022.
ing. 2nd ed. Hoboken, NJ: Wiley. Manitowoc. 2019. “Ground bearing pressure estimator.” Accessed
Anlin, W., J. Tao, D. Yaning, and L. Fei. 2012. “Study on structural insta- January 9, 2019. https://www.manitowoccranes.com/en/Tools/lift
bility of large crawler crane boom structure.” In Vol. 116 of Proc., 3rd -planning/ground-bearing-pressure.

© ASCE 04021038-15 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038


Marquez, A., P. Venturino, and J. L. Otegui. 2014. “Common root causes in permafrost peatlands, subarctic Canada.” Cold Reg. Sci. Technol.
recent failures of cranes.” Eng. Fail. Anal. 39: 55–64. https://doi.org/10 170 (2): 102930. https://doi.org/10.1016/j.coldregions.2019.102930.
.1016/j.engfailanal.2014.01.012. Slowey, K. 2020. “Alaska construction spending to near $7B in 2020.”
Matrax. 2020. “One size does not fit all.” Accessed July 6, 2020. https:// Accessed July 6, 2020. https://www.constructiondive.com/news
matraxinc.com/. /alaska-construction-spending-to-near-7b-in-2020/571813/.
McNutt, K. 1991. “SWOT before you start.” Accessed January 12, Terzaghi Terzaghi, K. 1943. Theoretical soil mechanics. New York: Wiley.
2019. https://journals.lww.com/nutritiontodayonline/Abstract/1991/01000 Vesić, A. S. 1973. “Analysis of ultimate loads of shallow foundations.”
/SWOT_Before_You_Start.9.aspx. J. Soil Mech. Found. Div. 99 (1): 45–73. https://doi.org/10.1061
Meyerhof, G. G. 1956. “Penetration tests and bearing capacity of cohesion- /JSFEAQ.0001846.
less soil.” J. Soil Mech. Found. Div. 82 (1): 1–19. https://doi.org/10 Wenping, F., Z. J. Yang, and T. Sun. 2019. “Ground freezing impact on
.1061/JSFEAQ.0000001. laterally loaded pile foundations considering strain rate effect.” Cold
Miao, Q., F. Niu, Z. Lin, J. Luo, and M. Liu. 2020. “Comparing frost heave Reg. Sci. Technol. 157 (Jan): 53–63. https://doi.org/10.1016/j
Downloaded from ascelibrary.org by UNIVERSITY OF ALBERTA LIBRARY on 04/27/21. Copyright ASCE. For personal use only; all rights reserved.

characteristics in cut and embankment sections along a high-speed .coldregions.2018.09.006.


railway in seasonally frozen ground of Northeast China.” Cold Reg. Western Regional Climate Center. 2020. “Anchorage WSCMO AP,
Sci. Technol. 170 (Feb): 102921. https://doi.org/10.1016/j.coldregions Alaska.” Accessed July 6, 2020. https://wrcc.dri.edu/cgi-bin/cliMAIN
.2019.102921. .pl?akanch.
Newton, C. 2017. “Construction complete on three gas plants near Dawson
Wu, D., Y. Lin, X. Wang, X. Wang, and S. Gao. 2011. “Algorithm of crane
Creek.” Accessed January 9, 2020. https://www.energeticcity.ca/2017
selection for heavy lifts.” J. Comput. Civ. Eng. 25 (1): 57–65. https://doi
/11/construction-complete-three-gas-plants-near-dawson-creek/.
.org/10.1061/(ASCE)CP.1943-5487.0000065.
Petrovic, J. 2003. “Review mechanical properties of ice and snow.”
Yang, Q., F. Qu, Z. Yu, and Z. Xie. 2019. “Stress and stability analysis of
J. Mater. Sci. 38 (1): 1–6. https://doi.org/10.1023/A:1021134128038.
slewing motion for crawler crane mounted on flexible ground.” Eng.
Ralph Wilson, C. 1982. “Dynamic properties of naturally frozen Fairbanks
Fail. Anal. 105: 817–827. https://doi.org/10.1016/j.engfailanal.2019
silt.” Master’s thesis, Oregan State Univ. Libraries & Press, Oregon
.07.005.
State Univ. https://ir.library.oregonstate.edu/concern/graduate_thesis
_or_dissertations/5x21th54x. Yang, X., Z. Xiu, L. Zhang, and X. Yan. 2010. “3D simulation of weak
Ravishankar, P., and D. N. Satyam. 2013. “Finite element modeling to foundation for good-sized oil storage tank.” In Proc., 2010 Int. Conf.
study soil structure interaction of asymmetrical tall buildings.” In Proc., on Mechanic Automation and Control Engineering, 1345–1348.
Workshop of TC207 during the 18th Int. Conf. on Soil-Mechanics and New York: IEEE. https://doi.org/10.1109/MACE.2010.5536276.
Geotechnical Engineering, 7. London: International Society for Soil Yang, Z., B. Still, and X. Ge. 2015. “Mechanical properties of seasonally
Mechanics and Geotechnical Engineering. frozen and permafrost soils at high strain rate.” Cold Reg. Sci. Technol.
Romanello, G. 2020. “A graphical approach for the determination of out- 113 (May): 12–19. https://doi.org/10.1016/j.coldregions.2015.02.008.
rigger loads in mobile cranes.” Mech. Based Des. Struct. Mach. 1–14. Yao, J., X. Qiu, Z. Zhou, Y. Fu, F. Xing, and E. Zhao. 2015. “Buckling
https://doi.org/10.1080/15397734.2020.1726184. failure analysis of all-terrain crane telescopic boom section.” Eng. Fail.
Schulson, E. M. 1999. “The structure and mechanical behavior of ice.” Anal. 57 (Nov): 105–117. https://doi.org/10.1016/j.engfailanal.2015.07
JOM 51 (2): 21–27. https://doi.org/10.1007/s11837-999-0206-4. .038.
Shahnavaz, F., H. Taghaddos, R. S. Najafabadi, and U. Hermann. 2020. Zaretsky, A. A., and H. I. Shapiro. 1997. “Overturning stability of a
“Multi crane lift simulation using building information modeling.” free standing crane under dynamic loading.” In Proc., 1997 SAE Int.
Autom. Constr. 118 (Oct): 103305. https://doi.org/10.1016/j.autcon Off-Highway and Powerplant Congress and Exposition, 184–192. War-
.2020.103305. rendale, PA: Society of Automotive Engineers. https://doi.org/10.4271
Shapiro, J., and L. Shapiro. 2010. Cranes and derricks. 4th ed. New York: /972721.
McGraw-Hill Professional. Zdravković, N., M. Savković, G. Marković, D. Pršić, and G. Pavlović.
Sladen, W. E., S. A. Wolfe, and P. D. Morse. 2020. “Evaluation of threshold 2019. “Numerical solution for the deflection of the column-mounted
freezing conditions for winter road construction over discontinuous jib crane structure.” Mech. Transp. Commun. 17 (3): 9.

© ASCE 04021038-16 J. Constr. Eng. Manage.

J. Constr. Eng. Manage., 2021, 147(6): 04021038

You might also like