A Hierarchical Bayesian Model of Workers' Responses to Proximity ...

5 downloads 39579 Views 2MB Size Report
Jan 12, 2015 - Email: [email protected] ... model based on workers' responses to proximity warnings of safety hazards to address the problem. .... Automating hazard identification that is an intuitive approach to addressing the ...
A Hierarchical Bayesian Model of Workers’ Responses to Proximity Warnings of Construction Safety Hazards: Towards Constant Review of Safety Risk Control Measures

1 2 3 4

Xiaochun Luo1; Heng Li2; Fei Dai3; Dongping Cao4; Xincong Yang5; Hongling Guo6

5

1

6

Hong Kong. E-mail: [email protected]

7

2

8

E-mail: [email protected]

9

3

Senior Research Fellow, Dept. of Building and Real Estate, Hong Kong Polytechnic Univ., Hung Hom, Kowloon,

Chair Professor, Dept. of Building and Real Estate, Hong Kong Polytechnic Univ., Hung Hom, Kowloon, Hong Kong.

Assistant Professor, Dept. of Civil and Environmental Engineering, West Virginia Univ., Morgantown, WV 26506-

10

6103, USA. E-mail: [email protected]

11

4

12

Kong. Email: [email protected]

13

5

14

Email:[email protected]

15

6

16

Beijing, China (corresponding author). Email: [email protected]

Postdoctoral Fellow, Dept. of Building and Real Estate, Hong Kong Polytechnic Univ., Hung Hom, Kowloon, Hong

PhD Candidate, Dept. of Building and Real Estate, Hong Kong Polytechnic Univ., Hung Hom, Kowloon, Hong Kong.

Associate Professor, Dept. of Construction Management, Tsinghua University, Qinghuayuan, Haidian District,

17 18

Abstract

19

Quickly changing and complicated workplace conditions, which are typical of construction projects, have

20

always been contributing to the poor safety record of the construction industry. However, few of existing

21

approaches to reviewing control measures take into account or effectively tackle them. This paper

22

introduces a three-level (i.e., individual hazards, hazard types, and generic hazards) hierarchical Bayesian

23

model based on workers’ responses to proximity warnings of safety hazards to address the problem. The

24

proposed model distinguishes between two uses of proximity warnings (i.e., as the primary control measure

25

and as the secondary measure), takes as input observational response rates, takes into account prior

26

knowledge on workers’ responses to similar hazards, and finally produces as output estimated response

27

rates (ERRs). These ERRs are the primary indicator of the effectiveness of a given control measure as well

28

as the validity of a previously identified safety hazard. A 17-day field experiment was conducted to test the

29

proposed approach preliminarily using a location-based proximity warning system, which was featured to

30

warn workers of safety hazards and record their responses to warnings. The experimental results

31

demonstrated the potential of the proposed approach. The limitations of the present study and the

32

directions for future research are also discussed.

33

Keywords

34

Proximity warning system; Safety risk management; Control measure review; Hierarchical Bayesian

35

modeling

36

37

Introduction

38

Safety risk management is traditionally conducted in the form of four sequential tasks: hazard identification,

39

risk evaluation, risk control, and control measure review (Manuele, 2005; Safe Work Australia, 2011). In

40

practice, the approaches to hazard identification are also used in control measure review (Safe Work

41

Australia, 2011), but with a different focus: identification focuses on finding newly-presented hazards, while

42

review focuses on validating those already identified. Hazards can be identified, or control measures

43

reviewed, through a general engineering evaluation based on previous data and operational history (an

44

informal approach), or a formal approach using well-developed identification/review techniques

45

(Kumamoto and Henley, 1996). These approaches are either reactive where information is analyzed after

46

accident occurrence or proactive where identification is based on historical data from similar hazards

47

(Khanzode, et al., 2012).

48 49

Although these reactive and proactive approaches do facilitate construction safety risk management, few

50

of them effectively take into account quickly changing and complicated workplace conditions, which have

51

always been contributing to the poor safety record of the construction industry (Haslam, et al., 2005; Hide,

52

et al., 2003). In Hong Kong, the numbers of fatal accidents in construction in 2013, 2014, and 2015 were

53

respectively 37, 38, and 43 (Occupational Safety and Health Branch, 2015; Occupational Safety and Health

54

Branch, 2016). Quickly changing and complicated workplace conditions were identified to have directly

55

accounted for over 30% of these accidents. There is a need to review safety risk control measures effectively

56

with the consideration of quickly changing and complicated workplace conditions.

57 58

Over the years, numerous variations and applications of proximity warning systems (PWS’s) have been

59

published in the literature (Carbonari, et al., 2011; Luo, et al., 2016b; Teizer and Cheng, 2015). Among them,

60

location-based PWS’s first detect dangerous proximities and then alert those people affected (Luo, et al.,

61

2016b); they acquire locations of both potential hazards and objects to be protected, and then evaluate the

62

relative distances between them to identify dangerous situations. These applications possess the ability to

63

monitor construction processes continuously. This study attempts to introduce a method of constantly

64

reviewing safety risk control measures based on workers’ responses to proximity warnings of safety hazards.

65 66

Review of Construction Safety Risk Management

67

The four steps in safety risk management (i.e., hazard identification, risk assessment, risk control, and

68

control measure review) can be organized into a cycle in that the output of the final step can meaningfully

69

inform and supplement the first step (see Fig. 1). This section reviews the literature concerning the four

70

stages together on account of (1) hazard identification and control review traditionally share most of their

71

methods, (2) the methods for risk assessment were used to enlighten the method design in this study with

72

respect to their dynamic nature, and (3) the hierarchical risk control methods were referenced to distinguish

73

between two primary uses of proximity warnings, which structure the proposed review logic (detailed in

74

next section).

75

76 77

Fig. 1. Safety risk management cycle

78 79

Hazard Identification

80

Hazard means “the inherent property or ability of something to cause harm” (Holt, 2001) and identification

81

is to identify the potential hazards. Khanzode, et al. (2012) group the hazard identification methods into

82

three categories. The first is referred to as the biased reactive approach, which is adopted after accident

83

occurrence. Examples of it are general engineering evaluations. The second is named as the biased

84

proactive approach, which is based on information from similar construction systems or methods of

85

previous data from the same system or method. Examples include fault-tree analysis and event-tree analysis.

86

The last is the unbiased proactive approach, which is adopted without waiting for the events to occur, and

87

without any restrictive assumptions about the presence of specific hazards, but with a focus on searching

88

potential hazardous elements and targets.

89 90

However, hazard identification and definition can become a major challenge due to the large-scale, dynamic,

91

and complex nature of construction projects. In fact, the hazard identification levels are far from ideal and,

92

for example, a studied project within both the railway and general construction industry only reported a

93

66.5% hazard identification level (Carter and Smith, 2006). Efficiency problem becomes outstanding on the

94

traditional identification approaches.

95

96

Automating hazard identification that is an intuitive approach to addressing the efficiency problem above

97

is implemented primarily with two strategies. The first strategy is to identify construction activities’ safety

98

hazards in conjunction with construction schedules. For example, Navon and Kolton (2006) investigated

99

how to automate fall prevention procedures and developed an automated model that identifies the

100

dangerous activities in project schedules, which in turn lead to the identification of fall hazards. Hallowell,

101

et al. (2011) proposed a safety risk interaction matrix to quantify the impact that pairwise spatial and

102

temporal interactions on the base-level risk of 25 common highway construction tasks in the United States.

103

Based on that matrix, Esmaeili and Hallowell (2013) introduced a system to produce predictive plots of

104

safety risk over time based on the temporal and spatial interactions among concurrent activities.

105 106

The second strategy is to utilize building information models (BIM) to identify latent safety hazards,

107

primarily by rule-based reasoning, where geometric information of building elements is checked against

108

safety rules to determine dangerous conditions and factors (Kiviniemi, et al., 2011). For example, rule-based

109

fall hazard identification can be implemented in two steps: the first goes to identify if there are hazardous

110

openings or edges, and the second is to determine if adequate protection facilities (e.g., guardrails) are put

111

in place (Zhang, et al., 2011; Zhang, et al., 2015; Zhang, et al., 2013). Similarly, Hammad, et al. (2012)

112

reported a study to prevent falls from a height by automatically identifying hazardous slab edges and

113

generating corresponding virtual fences. More recently, Kim, et al. (2016) reported a method of automated

114

hazard identification based on the deviation between optimal routes, which were shortest paths

115

determined by BIM, and actual routes of workers, which were collected with a real-time location system.

116 117

Risk Assessment

118

Risk means “the chance or probability of loss”, and an evaluation of the potential for failure (Holt, 2001).

119

Risk assessment methods can be classified as qualitative methods producing as output a qualitative risk

120

level and quantitative methods providing a quantitative score (Luo, et al., 2016a). However, the traditional

121

and prospective methods are criticized as ignoring the actual execution of construction activities

122

(Mitropoulos and Namboodiri, 2010) and being subject to safety officers’ experience and diligence.

123

Furthermore, officers can be overwhelmed due to the changing and complex working place situations and

124

concurrent involvement of numerous resources on construction sites (Behzadan, et al., 2008). To address

125

this problem, Luo, et al. (2016a) introduced a dynamic risk assessment method by quantifying workers’

126

hazard exposure amounts based on real-time location data of workers and construction equipment.

127

Similarly but more specifically, Teizer and Cheng (2015) proposed a leading safety indicator, “proximity

128

hazard indicator,” by automatically gathering and analyzing the spatial–temporal conflicts between

129

workers-on-foot and large construction equipment. Continuous informational feeding and feedback

130

between workplaces and workers are essential to the methods envisaged to monitor dynamic construction

131

processes.

132 133

Risk Control

134

The measures of controlling safety risks can be meaningfully organized into a six-level hierarchy in terms of

135

their effectiveness (Manuele, 2005): (1) eliminate hazards through system design and redesign, (2) reduce

136

risks by substituting less hazardous methods and materials, (3) incorporate safety devices, (4) provide

137

warning systems, (5) apply administrative controls, and (6) provide personal protection equipment. The first

138

three levels are more useful because they are preventive actions that reduce risk by design, substitute

139

methods and materials, and most importantly, rely least on personnel performance (Manuele, 2005). The

140

last three levels rely significantly on human behavior and thus are believed to be less effective.

141 142

However, research attributes the majority of safety accidents to human behavior. Heinrich (1939) stated

143

that unsafety acts cause around 85% of accidents. Blackmon and Gramopadhye (1995) found that unsafe

144

behavior causes 98% of accidents. Therefore, measures at the last three levels are critical to reducing safety

145

accidents on sites. In this context, the authors have developed a PWS, the Proactive Construction

146

Management System (PCMS), which pertains to the fourth level, “provide warning systems”, and has been

147

introduced in detail in (Luo, et al., 2016a; Luo, et al., 2016b).

148 149

Control Measure Review

150

To ensure control measures work as planned, they should be reviewed, and if necessary, revised (Manuele,

151

2005; Safe Work Australia, 2011). Thus, control measure review is an integral step in the effective problem-

152

solving technique (Manuele, 2005). The dynamic nature of construction sites, which can turn the engaged

153

measures ineffective, or give rise to a new safety risk that these means may be not adequate to handle,

154

intensifies the need to examine control measures (Safe Work Australia, 2011). In the former case, the

155

engaged control measures should be revised. In the latter case, the approaches toward hazard identification

156

are used in reviewing control measures (Safe Work Australia, 2011); control measure review focuses on

157

validating the hazards already identified, while hazard identification focuses on finding newly-presented

158

hazards. However, review processes with these methods seldom take into account or effectively tackle

159

quickly changing workplace conditions; they are primarily based on previous information from similar

160

hazards or factual information after accident occurrence.

161 162

In summary, rapidly changing workplace conditions pose the greatest challenge to existing methods for

163

reviewing safety risk control measures, which take the form of either reactive analysis or proactive

164

projection. However, the engaged control measures can quickly become invalid due to an entirely new

165

hazardous situation or a sharp change in the envisaged position. Therefore, continuously reviewing those

166

employed risk control measures is critical for making a prompt and efficient response to situational changes.

167 168

Research Objective, Scope, and Definitions

169

The objective of this paper was to introduce a novel method for continuous review of safety risk control

170

measures according to workers’ responses to proximity warnings of safety hazards. The PCMS was used to

171

track workers and record their answers to proximity warnings. Note that the hazards investigated in this

172

paper were restricted to those that can be addressed by proximity warnings. These hazards are classified

173

into static hazards, e.g., unprotected roof edges, roof and floor openings, and dynamic hazards, e.g., tower

174

crane hooks, excavators and bulldozers (Luo, et al., 2016b). They are defined in the form of virtual fences

175

at the backend of the PCMS to identify dangerous proximities of objects to be protected.

176 177

A 2-second rule was introduced to evaluate if a worker responds to a proximity warning in light of the earlier

178

research on driver brake reaction time (Green, 2000; Summala, 2000), which found that a driver could react

179

to surprise events, such as an object suddenly moving into their path, in roughly 1.5 seconds. In

180

consideration of that, most of the time workers and equipment on construction sites move at a slower

181

speed than cars do on highways, the effective response time threshold was set as 2 seconds (Luo, et al.,

182

2016a). As a result, a reaction is viewed as a response only if a worker does not walk closer to a hazard after

183

2 seconds from receiving a personal hazard-specific warning until they leave the hazard area, or it is viewed

184

as a neglect. The response rate (RR) of a particular hazard can be observational and defined as the ratio of

185

the number of responses to the number of warnings occurred in the hazard area in a unit time, which was

186

a workday in this study. The observational RR is abbreviated as ORR for the sake of convenience.

187 188

However, it is desirable to view RRs from a comprehensive perspective since safety hazards on construction

189

sites are dynamic in nature. Usually, construction activities at different stages by multiple trades are

190

involved, and site situations change quickly. The observational numbers of responses to and neglects of

191

warnings can conceal safety hazards with higher risks or overstate those with lower risks. Prior knowledge

192

on workers’ responses to similar hazards can be utilized to counterbalance the impact of the simplification

193

of only using observational data.

194 195

Hierarchical Bayesian modeling is an appropriate approach to accommodating all these concerns due to its

196

potential to overrule classical frequentist statistics in applications where respondents give multiple

197

observational data (Gelman, et al., 2014). It describes a statistical model in a hierarchical form of various

198

levels estimating the parameters of the posterior distribution using the Bayesian method (Allenby, et al.,

199

2005). To this end, this paper introduces a three-level (i.e., individual hazards, hazard types, and generic

200

hazards) hierarchical Bayesian model based on workers’ responses to proximity warnings of safety hazards.

201

Therefore, control measures can be hierarchically reviewed based on the prior knowledge on workers’

202

responses to warnings of similar hazards as well as the observational warnings and responses. The posterior

203

estimates of RRs at different levels are also referred to as estimated response rates (ERRs) in this study,

204

which are the primary indicator to review the effectiveness of a given control measure as well as the validity

205

of a previously identified safety hazard.

206 207

This study differentiates between two uses of proximity warnings (see Fig. 2), namely as the secondary risk

208

control measure and as the primary control measure. In contrast with those more effective control

209

measures such as elimination, substitution, and safety devices (Manuele, 2005), providing warning systems

210

is a second or supplementary choice. When those more effective control measures are adopted, warnings

211

can be a complementary prevention. In this case, warnings are referred to as a secondary control measure.

212

In practice, however, elimination, substitution, and safety devices are not employed to cope with all risk

213

situations, due to tight time schedules or technical reasons. For example, on those congested construction

214

sites, workers-on-foot can frequently be observed working around large construction equipment such as

215

excavators and trucks, and moving near lifting areas, openings that are not defended for material convey.

216

In these situations, proximity warnings are used as the primary control measure. As a result, the decision

217

tree for reviewing control measures is formed as follows:

218

(1) When proximity warnings are used as the primary risk control measure of a hazard, the higher the ERR

219

is, the more effective the control would be. In this situation, when the ERR arrives at a level, more

220

effective control strategies such as elimination, substitution, or safety devices should be considered.

221

On the contrary, a small ERR denotes that proximity warnings are little meaningful to workers, and the

222

virtual hazard can be removed in the PCMS.

223

(2) When warnings are used as the secondary risk control, the higher the ERR is, the less effective the

224

primary control would be. In this situation, the primary control should be adjusted in response to the

225

changed workplace conditions. However, if the ERR is small, safety officers are suggested to involve the

226

review decision in that the fact can be either the primary control is effective, or the hazard is invalid. In

227

the former situation, it is adequate to maintain the status quo. In the latter case, it is suggested that

228

the virtual hazard is removed in the PCMS to reduce false warnings.

229

230 231

Fig. 2. Decision tree for reviewing control measures

232 233

Hierarchical Bayesian Modelling

234

Fig. 3 shows the hierarchical Bayesian model, which was proposed to implement the control measure

235

review logic, and defined by latent data intermediate between the observational data and underlying

236

parameters. In this model, the observational data precisely consists of (1) the number of warnings 𝑛𝑖

237

occurred in hazard area i, (2) the number of responses 𝑥𝑖 observed in hazard area i, (3) the number of

238

hazard areas 𝑁ℎ , and (4) the number of hazard types 𝑁𝑡 . The underlying parameters were organized into

239

a three-level structure: individual hazards, hazard types, and generic hazards, and thus the estimate of the

240

parameter of any particular hazard depends on the estimates of the parameters of its hazard type, which

241

in turn depend on the estimates of the parameters of generic hazards.

242 243

Fig. 3. Bayesian hierarchical model

244 245

In the hierarchical model, the level of individual hazards was designed to describe workers’ responses to

246

proximity warnings in a particular hazard area i, which belongs to a specific hazard type j. As a result, the

247

number of responses 𝑥𝑖 depends on the number of warnings 𝑛𝑖 and the RR 𝜃𝑖 in the form of a binomial

248

distribution, expressed as Equation 1. This modeling section recognized that site workers tend to have

249

different RRs to various hazards because of the distinct and complicated nature of the workplace conditions

250

that form individual hazards, and other complex factors such as task urgency and complexity, workers’

251

safety knowledge and experience (Luo, et al., 2016a).

252 𝑥𝑖 ~𝐵(𝑛𝑖 , 𝜃𝑖 )

(1)

253 254

The RR 𝜃𝑖 obviously is not a constant and was supposed to be distributed as a beta density with shape

255

constants 𝑎𝑗 and 𝑏𝑗 , expressed as Equation 2, where the subscript j indicates that hazard i falls into hazard

256

type j. This treatment assumes that workers’ responses to a type of hazards are similar and thus can be

257

clustered. It enables that ERRs of various hazards of an identical hazard type can be meaningfully compared

258

and evaluated.

259 𝜃𝑖 ~𝐵𝑒𝑡𝑎(𝑎𝑗 , 𝑏𝑗 ) 260

(2)

261

However, to convert a prior belief expressed in terms of central tendency and sample size into equivalent,

262

the values of 𝑎𝑗 and 𝑏𝑗 in the beta distribution were re-expressed with the mode ω𝑗 , and concentration

263

κ𝑗 of the beta distribution: 𝑎𝑗 = 𝜔𝑗 (𝜅𝑗 − 2) + 1 and 𝑏𝑗 = (1 − 𝜔𝑗 )(κ𝑗 − 2) + 1 for κ𝑗 > 2 , where

264

𝜔𝑗 = (𝑎𝑗 − 1)/(𝑎𝑗 + 𝑏𝑗 − 2) and κ𝑗 = 𝑎𝑗 + 𝑏𝑗 . Therefore, concerning hazard type j, 𝜔𝑗 represents a

265

prior belief about 𝜃𝑖 , its posterior estimate is the ERR of hazard type j, and the value of κ𝑗 governs how

266

near 𝜃𝑖 is to 𝜔𝑗 , with larger values of κ𝑗 generating values of 𝜃𝑖 more concentrated near 𝜔𝑗 (Kruschke,

267

2014). Consequently, Equation 2 was re-written as Equation 3.

268 𝜃𝑖 ~𝐵𝑒𝑡𝑎(𝜔𝑗 (𝜅𝑗 − 2) + 1, (1 − 𝜔𝑗 )(𝜅𝑗 − 2) + 1)

(3)

269 270

Besides, the study supposed the distribution over 𝜔𝑗 as another beta distribution expressed as Equation

271

4. Similarly, the parameters 𝜔 and 𝜅 denote the mode and the concentration of the distribution over 𝜔𝑗 .

272

Most importantly, 𝜔 and 𝜅 are the parameters at the generic hazard level constraining the estimates of

273

the parameters at the hazard type level, which in turn limit the estimates at the individual hazard level. The

274

posterior estimate of 𝜔 is the ERR of the generic hazards.

275 𝜔𝑗 ~𝐵𝑒𝑡𝑎(𝜔(𝜅 − 2) + 1, (1 − 𝜔)(𝜅 − 2) + 1)

(4)

276 277

In real situations, the value of 𝜅𝑗 is not well known in advance, and instead, this paper used the previously

278

observational data to inform its credible values. Intuitively, when the data from different hazards of hazard

279

type j shows very similar RRs, 𝜅𝑗 is large. But when the data shows diverse RRs, then 𝜅𝑗 is small. To

280

express prior uncertainty about 𝜅𝑗 , the hierarchical model includes a prior distribution on 𝜅𝑗 . Because the

281

value of 𝜅𝑗 2 must be non-negative, the prior distribution on 𝜅𝑗 2 must not allow negative values. A

282

gamma distribution was adopted to represent its density, as in Equation 5, where S𝜅 and 𝑅𝜅 are its shape

283

and rate parameters. Similarly, at the generic hazard level, 𝜅 was treated in the same way with 𝜅𝑗 and

284

supposed to be distributed as another gamma distribution as in Equation 6, which shares the shape and

285

rate parameters with 𝜅𝑗 .

286 (𝜅𝑗 − 2)~𝐺𝑎𝑚𝑚𝑎(S𝜅 , 𝑅𝜅 )

(5)

(𝜅 − 2)~𝐺𝑎𝑚𝑚𝑎(S𝜅 , 𝑅𝜅 )

(6)

287 288

The prior distribution over 𝜔 was selected as a beta distribution as in Equation 7, where 𝐴𝜔 and 𝐵𝜔 are

289

constants determined by prior knowledge on the RR of generic hazards. In this case, it is believed that ω is

290

typically near the mode of the distribution, (𝐴𝜔 − 1)/(𝐴𝜔 + 𝐵𝜔 − 2).

291 𝜔~𝐵𝑒𝑡𝑎(𝐴𝜔 , 𝐵𝜔 )

(7)

292 293

In summary, the proposed hierarchical Bayesian model consists of three levels. At the individual hazard

294

level, the parameters to be estimated are their RRs < 𝜃1 , 𝜃2 , … , 𝜃𝑁ℎ > . At the hazard type level, the

295

parameters to be estimated are the modes < 𝜔1 , 𝜔2 , … , 𝜔𝑁𝑡 >, which represent the RRs of hazard types,

296

and the concentrations < 𝜅1 , 𝜅2 , … , 𝜅𝑁𝑡 >. At the generic hazard level, the parameters to be estimated are

297

the mode 𝜔 representing the RR of generic hazards, and the concentration 𝜅. Therefore, the continuous

298

review of safety risk control measures is transformed into estimating these model parameters according to

299

the observational data and the prior knowledge on the prior distributions of 𝜅𝑗 , 𝜔, and 𝜅.

300 301

Experiment

302

The specific technologies and techniques of the PCMS have been introduced in the authors' previous

303

research. For example, Li, et al. (2015) focused on evaluating various real-time location technologies,

304

including chirp spread spectrum, global positioning system, radio-frequency-identification, and ultra-wide

305

band. Luo, et al. (2016b) further concentrated on risk assessment by dynamically quantifying workers’

306

hazard exposure based on the real-time trajectories of workforce and equipment. Workers’ responses to

307

proximity warnings were also investigated (Luo, et al., 2016a) and the findings were used as the prior

308

knowledge in establishing the prior parameters of the proposed model.

309 310

Experimental Settings and Procedure

311

The experiment was carried out on a viaduct construction project in Shanghai, China. A total of 72 male

312

worker volunteers aged between 21 and 55 participated and the trial lasted 17 workdays between 12

313

January 2015 and 21 April 2015. To initialize the PCMS, a latest site layout drawing was uploaded to its

314

backend, and location anchors were marked on the layout and deployed physically on-site, location tags

315

were mounted on workers’ safety helmets, and most importantly, hazard zones were defined in the form

316

of virtual fences at the system’s backend, as shown in Fig. 4. In running system, when workers were detected

317

entering an unauthorized hazard zone, a signal was sent to their tags, and a personal hazard-specific

318

warning (e.g., “ground uneven,” “vehicle approaching,” and “lifting hook approaching”) would sound.

319

Finally, warnings and workers’ reactions regarding the response time were recorded for further analysis.

320

Readers are referred to the articles by Luo, et al. (2016a) and Luo, et al. (2016b) for more details about the

321

application of the PCMS. Additionally, to ensure the protection of participants’ rights and welfare, the

322

authors had clearly explained the use of location devices and the functioning mechanism of the warnings

323

to them. Most importantly, they were allowed to switch off the location tags to quit the experiment if they

324

feel uncomfortable.

325 326

Fig. 4. System deployment in the experiment

327 328

In contrast with (Luo, et al., 2016a), this study focused on demonstrating the hierarchical Bayesian model’s

329

capability of continuously reviewing safety risk control measures. To this end, the review time interval was

330

selected as one workday and the hazard zones that lasted more than one workday were counted as separate

331

ones. As a result, a total of 50 nominal static hazard zones (namely 30 independent hazard zones), which

332

fell into nine hazard types, were identified and defined manually in the 17-workday experiment and a total

333

of 5391 warnings were recorded. Table 1 summarizes the experimental settings and procedure as well as

334

the warnings and responses observed, which were fed into the proposed model for estimation.

335 336

Table 1. Experimental settings and procedure, and recorded warnings and responses

337 338

In the experiment, the values of 𝐴𝜔 and 𝐵𝜔 were set at 30 and 30 respectively, which denoted there were

339

30 responses and 30 neglects in a genetic hazard area in a unit time (a work day) and the concentration

340

(𝐴𝜔 + 𝐵𝜔 ) was 60, on account of the median of responses in all hazard zones was 35 and that of neglects

341

was 36.5. The two numbers (35 and 36.5) were established according to the prior knowledge on workers’

342

responses obtained in the last two years of the development and trial of the PCMS. The adopted values

343

were slightly smaller than actually possible numbers of responses and neglects with an aim to reflect more

344

influence of the observational data on the estimates of model parameters.

345 346

To establish the values of S𝜅 and 𝑅𝜅 , the mode 𝜔𝜅 was selected as 58, which was also the concentration

347

(𝐴𝜔 + 𝐵𝜔 ) minus 2, and the standard deviation 𝜎𝜅 was 58, denoting the significant variation of the

348

warning numbers in various hazard areas. Finally, S𝜅 , 2.6180, and 𝑅𝜅 , 0.0279, were derived according to

349

𝑆𝜅 = 1 + 𝜔𝜅 𝑅𝜅 , where 𝑅𝜅 = (𝜔𝜅 + √𝜔𝜅 2 + 4𝜎𝜅 2)/2𝜎𝜅 2 (Kruschke, 2014).

350 351

Computational Algorithm

352

Ideally, the posterior estimates can be calculated analytically, but for the experimental model with 70

353

parameters < 𝜃1 , 𝜃2 , … , 𝜃50 , 𝜔1 , 𝜔2 , … , 𝜔9 , 𝜅1 , 𝜅2 , … , 𝜅9 , 𝜔, 𝜅 > , an analytical calculation is almost

354

infeasible. In this situation, two approaches are available (Salvatier, et al., 2016): (1) finding the maximum

355

a posteriori (MAP) point using optimization methods, and (2) computing summaries based on samples

356

drawn from the posterior distribution using Markov Chain Monte Carlo (MCMC) sampling methods. Since

357

the MAP approach gives point estimates of the parameters and can be biased if the model isn’t

358

representative of the distribution (Salvatier, et al., 2016), this paper adopted one of the MCMC methods —

359

the Metropolis algorithm, which was placed among the ten algorithms that have had the greatest influence

360

on the development and practice of science and engineering in the 20th century (Beichl and Sullivan, 2000).

361

PyMC3, which is a Python module for Bayesian statistical modeling and model fitting which focuses on

362

advanced MCMC fitting algorithms (Salvatier, et al., 2016), was adopted to compute the posterior estimates.

363 364

Results and Analysis

365

The total number of simulation steps was 51,000 and the first 1,000 steps in the burn-in period were

366

removed in the final results due to their unrepresentativeness. Fig. 5 shows the traces of several parameters,

367

which are selected to illustrate the simulation process. Table 2 summarizes the estimates of the posterior

368

parameters in the form of quantiles, and Fig. 6 is a box-and-whisker plot showing all omegas and thetas,

369

which are grouped by hazard types. In this section, for the sake of succinctness, the estimates of the

370

parameters implicitly refer to their medians, unless they are explicitly specified.

371 372

Fig. 5. Trace plot of several selected parameters, where the burn-in step number is 1,000 and the thin

373

value 20 resulting in there are only (51,000 – 1,000)/20 = 2,500 steps shown.

374 375

Table 2. Quantiles of parameter estimates

376 377

Fig. 6. Box-and-whisker plot of ERRs (omegas and thetas), where ‘*’ indicates that proximity warnings

378

are used as the secondary safety control measure in those hazard areas; those without ‘*’ use proximity

379

warnings as the primary control measure. The red dashed line indicates the ERR of the generic hazards,

380

̂ (0.528), and the red horizontal span represents its interquartile range, which is the namely 𝝎

381

difference between the lower quartile (0.503) and the upper quartile (0.552). Similarly, the yellow

382

dash-dot line in each column shows each hazard type ERR and the yellow horizontal span represents

383

the interquartile range between its lower and upper quartiles. For example, in the first column

384

̂ 𝟏 ) was about 0.487 with its interquartile range representing the welding areas, their hazard type ERR (𝝎

385

between the lower quartile (0.466) and the upper quartile (0.510), which is represented by the yellow

386

horizontal span. The estimates of individual hazards are represented in the form of boxes with

387

whiskers.

388 389

Use 1: Proximity warnings as the primary risk control measure

390

In this situation, the higher the ERR is, the more effective the primary control would be. As shown in the

391

second column of Fig. 6, proximity warnings in laydown areas, as the primary control measure to reduce

392

the risk of being struck by moving or dropping objects, had a relatively high hazard type ERR (𝜔 ̂2 was about

393

0.648). However, the ERRs varied in this type of hazards (𝜅̂ 2 was about 63.9). For example, Hazard 24 had

394 395

62 responses out of 77 warnings, which led its ERR to the highest in this type (𝜃̂24 was about 0.733). By contrast, Hazard 28 had the lowest ERR in this hazard type (𝜃̂28 was about 0.572 originated from 104

396

responses of 191 warnings). Although the higher ERRs justified the effectiveness of the control measure for

397

these hazards, more effective engineering control measures (e.g., separating average workers from the

398

laydown areas using caution tape) were supposed to be adopted.

399 400

On the contrary, the lower the ERR is, the less valid the hazard would be. In this situation, the virtual hazard

401

defined at the backend of the PCMS can be removed to reduce false warnings. Take the easy-to-stumble

402

areas as an example. They had the lowest ERRs, which informed the estimate of the hazard type’s ERR (𝜔 ̂5

403

was about 0.412), which was lower than the overall ERR (0.528). Thus, as the primary control measure,

404

proximity warnings of the easy-to-stumble risk were ineffective, probably due to there were full of this kind

405

of hazards on construction sites; warnings of the risks were little meaningful to workers. Specifically, Hazard

406

30 received 24 responses out of 80 warnings, and this led to the lowest ERR (𝜃̂30 was about 0.356) during

407

the whole experiment, which indicated that workers just ignore the majority of the warnings. Therefore,

408

these easy-to-stumble areas should be removed from the PCMS.

409 410

Use 2: Proximity warnings as the secondary risk control measure

411

In this situation, the higher the ERR is, the less effective the primary control would be. Workers believe that

412

the warnings are valid and perceive the high risks even with the main control measure engaged. For

413

example, the warnings in the hazard areas related to floor openings had the highest ERR (𝜔 ̂7 was about

414 416

0.795). Among this type of hazards, Hazard 47 had the lowest ERR (𝜃̂47 was about 0.763 derived according to 23 responses out of 33 warnings). Hazard 41 had the highest ERR (𝜃̂41 was about 0.829 derived according to 26 responses out of 28 warnings). In comparison with Hazard 41, the ERR of Hazard 42 (𝜃̂42 was about

417

0.807) was not the highest one in spite of all its six warnings got responses. The higher ERRs showed that

418

the primary control measure should be strengthened essentially or in a more visual way (e.g., fencing

419

openings with hand rails of high contrast colors showing the risk of fall from height relevant to the floor

420

openings).

415

421 422

On the contrary, when proximity warnings are adopted as the secondary control of a safety risk, a lower

423

ERR denotes that the primary control is effective, or the relevant hazard is not valid any longer. Take Hazard

424

3 as an example. In the experiment, its primary control was fencing the ground area that could be influenced

425

by dropping objects of the aerial work with yellow caution tape to prevent workers from entering

426

unwittingly. The virtual fence in the PCMS was defined by expanding the virtual ground area by 2 meters.

427

As shown in the sixth column of Fig. 6, Hazard 3 had a relatively low ERR (𝜃̂3 was about 0.435), which was

428

lower than that of all aerial work areas (𝜔 ̂6 was about 0.455). In turn, the type’s ERR (0.454) was lower

429

than the overall ERR (0.528). The result indicated that the primary control measure of the aerial work area

430

was effective, the risk of being struck by falling objects had been reduced to an acceptable level, and the

431

workers would not like to respond to the false warnings.

432 433

Discussions

434

This section first discusses the knowledge contribution of this study regarding its innovative points, which

435

are followed by its practical implication. Finally, the research limitations are discussed.

436 437

Contribution to Knowledge

438

This study primarily contributes to knowledge from two perspectives. First, workers’ responses to proximity

439

warnings of safety hazards are innovatively utilized to review safety control measures in the context of

440

dynamic and complicated construction sites. The proposed approach possesses the potential to make the

441

reviewing process continuous and meaningful because proximity warnings and workers’ responses are

442

hazard-specific, real-time, and objective. In the long term, a continuous and automatic method for

443

reviewing construction safety risk control measures is critical for implementing the roadmap of digital sites

444

(Bowden, et al., 2006).

445 446

Second, the proposed approach towards constant review of risk control measures takes into account both

447

the prior knowledge on safety risk management and observational process information. Specifically, the

448

hierarchical Bayesian modeling of workers’ responses to proximity warnings of safety hazards uses the

449

safety knowledge to establish prior parameters and the real-time ORRs to estimate posterior parameters.

450

The three-level structure of the model recognizes RR differences among various hazard types and

451

characteristics of individual hazards. Moreover, the parameter estimates of a specific hazard are

452

simultaneously informed by data from all the other hazards of an identical hazard type, since all these

453

parameters inform the parameters of higher levels (i.e., the hazard type level and, in turn, the generic

454

hazard level), which in reverse constrain the parameter estimates of all individual hazards.

455 456

Practical Implication

457

The proposed model takes into account two uses of proximity warnings and can be embraced to improve

458

the PCMS and other location-based PWS’s that depends on predefined virtual hazards. When proximity

459

warnings are used as the primary risk control measure, the higher ERR is, the more effective the control

460

would be; on the contrary, the lower ERR, the less valid the hazard. When they are used as the secondary

461

risk control, the higher ERR is, the less effective the primary control would be; in contrast, the lower ERR,

462

the more effective the primary control, or the less valid the hazard. Therefore, the simple logic can be coded

463

to reinforce the functions of the PCMS to make it more viable in practical application. For example, the

464

PCMS can be extended by adding the automatic hazard recognition and definition algorithms (Kiviniemi, et

465

al., 2011) and implementing the proposed hierarchical Bayesian model towards an automatic safety risk

466

management cycle.

467 468

Research Limitations

469

Despite the value of this research mentioned above, several research limitations need to be acknowledged.

470

First, the response differences between individual workers were ignored, and only the number of effective

471

responses and the number of warnings as an overarching group were collected. The authors believe that

472

this treatment has limited impact on the accomplishment of the research objective, which is to review risk

473

control measures continuously, rather than the RRs of individual workers.

474

475

Second, the selection of prior distributions and the value establishment of prior parameters seemed

476

subjective. In fact, these problems are inherent and typical of Bayesian statistics (Kass and Wasserman,

477

1996), which have been debated since Bayes (1763). Furthermore, there is an extensive literature on the

478

construction of priors, and the subject seems far from a definitive solution (Kass and Wasserman, 1993).

479

Priors can be selected using some methods (Carlin and Louis, 2008), including determination from past

480

information, elicitation from the purely subjective assessment of experts, symmetry or maximizing entropy

481

given constraints, etc. In this study, prior distributions were selected on account of if there were conjugate

482

priors with the aim of speeding up the fitting process of the model, and prior values were established

483

primarily based on past information accumulated in the development and trial of the PCMS.

484 485

Third, although workers’ responses to proximity warnings are determined by a combination of various

486

factors, including task urgency and complexity, safety knowledge and experience, existing risk control

487

measures, and hazard validity (Luo, et al., 2016a), this study primarily focused on the last two factors. Task

488

urgency and complexity were fused into the three-level structure of the model, namely individual hazards,

489

hazard types, and generic hazards. Safety knowledge and experience are closely related to the first

490

limitation and were treated indistinctly among workers. From a practitioner’s perspective, it is worth to

491

sacrifice somewhat pedantic integrity by fusing and simplifying some factors for the understandability of

492

the model and its computing efficiency.

493 494

Concluding Remarks

495

This paper introduced a three-level (i.e., individual hazards, hazard types, and generic hazards) hierarchical

496

Bayesian model based on workers’ responses to proximity warnings of specific hazards to review safety risk

497

control measures with the consideration of quickly changing and complicated workplace conditions. The

498

proposed model distinguishes between two uses of proximity warnings (i.e., as the primary control measure

499

and as the secondary measure), takes as input observational response rates, takes into account the prior

500

knowledge on workers’ responses to similar hazards, and finally produces as output ERRs of individual

501

hazards, hazard types, and generic hazards.

502 503

These ERRs are used as the primary indicator of the effectiveness of a given control measure as well as the

504

validity of a previously identified safety hazard. As a result, the control measure review logic consists of the

505

following rules: (1) when proximity warnings are used as the primary risk control measure, the higher the

506

ERR is, the more effective the control measure would be; on the contrary, the lower the ERR, the less valid

507

the hazard; and (2) when they are used as the secondary risk control, the higher the RR is, the less effective

508

the primary control would be; on the contrary, the lower the ERR, the more effective the primary control,

509

or the less valid the hazard.

510 511

To preliminarily test the proposed approach, a 17-day field experiment using the PCMS was conducted and

512

a total of 5391 warnings and 2712 responses in 50 hazards, which fell into nine types, were collected. The

513

observational data was fed into the model to estimate the model parameters. The analysis of experiment

514

results demonstrated how to utilize the estimates in practice as well as the potential of the proposed

515

approach to continuously reviewing safety risk control measures with the consideration of quickly changing

516

and complicated workplace conditions. Given more data collected in practical applications of the proposed

517

approach, the posterior estimates of the parameters, especially those of hazard types and generic hazards,

518

can be meaningfully referenced to establish their prior values, which in turn can significantly improve the

519

credibility of the estimates.

520 521

Automatic safety risk management cycle is an essential part of “the construction sites of the future”

522

(Bowden, et al., 2006), where context-aware and self-adapting sensor networks and decision support

523

systems monitor and complement workers' activities, at every level and in real time. The proposed review

524

approach could represent one step towards the vision. However, as indicated in the research limitations, to

525

ensure the short- and mid-term viability of the model, the Bayesian model fused and simplified some

526

factors, e.g., task urgency and complexity, and safety knowledge and experience. As the development of

527

data mining technologies and wireless sensor technologies, the indirect quantification of these factors

528

becomes possible, for example, by finding correlations in large data (Speed, 2011), which has the potential

529

to extend the proposed approach regarding comprehensiveness without undermining its viability.

530 531

Acknowledgements

532

The authors are in great debt to the site manager and the workers of the construction project for their

533

support during the experiment. Many thanks also go to Bo Ye, Wei Lu and Xintao Yang for their

534

indispensable efforts in the system development and experimental study. The work was supported by the

535

Innovation and Technology Commission of Hong Kong, under the public sector trial scheme of a project

536

"Location-based Technologies for Asset Tracking and Risk Management" (ITT/004/15LP), and the Research

537

Grants Council of Hong Kong with the grant titled “Proactively Monitoring Construction Progress by

538

Integrating 3D Laser-scanning and BIM” (PolyU 152093/14E).

539 540

541

References

542

Allenby, G. M., Rossi, P. E., and McCulloch, R. E. (2005). "Hierarchical Bayes models: a practitioners guide."

543

. (May 11, 2015).

544

Bayes, T. (1763). "An Essay towards solving a Problem in the Doctrine of Chances." Philosophical Transactions of

545

the Royal Society of London, 53(0), 370-418.

546

Behzadan, A. H., Aziz, Z., Anumba, C. J., and Kamat, V. R. (2008). "Ubiquitous location tracking for context-specific

547

information delivery on construction sites." Automation in Construction, 17(6), 737-748.

548

Beichl, I., and Sullivan, F. (2000). "The Metropolis Algorithm." Computing in Science & Engineering, 2(1), 65-69.

549

Blackmon, R., and Gramopadhye, A. (1995). "Improving Construction Safety by Providing Positive Feedback on

550

Backup Alarms." Journal of Construction Engineering and Management, 121(2), 166-171.

551

Bowden, S., Dorr, A., Thorpe, T., and Anumba, C. (2006). "Mobile ICT support for construction process

552

improvement." Automation in Construction, 15(5), 664-676.

553

Carbonari, A., Giretti, A., and Naticchia, B. (2011). "A proactive system for real-time safety management in

554

construction sites." Automation in Construction, 20(6), 686-698.

555

Carlin, B. P., and Louis, T. A. (2008). Bayesian Methods for Data Analysis, CRC Press, Boca Raton & London & New

556

York.

557

Carter, G., and Smith, S. D. (2006). "Safety hazard identification on construction projects." Journal of Construction

558

Engineering and Management, 132(2), 197-205.

559

Esmaeili, B., and Hallowell, M. (2013). "Integration of safety risk data with highway construction schedules."

560

Construction Management and Economics, 31(6), 528-541.

561

Gelman, A., Carlin, J. B., Stern, H. S., and Rubin, D. B. (2014). Bayesian Data Analysis, CRC Press, Boca Raton &

562

London & New York.

563

Green, M. (2000). ""How Long Does It Take to Stop?" Methodological Analysis of Driver Perception-Brake Times."

564

Transportation Human Factors, 2(3), 195-216.

565

Hallowell, M., Esmaeili, B., and Chinowsky, P. (2011). "Safety risk interactions among highway construction work

566

tasks." Construction Management and Economics, 29(4), 417-429.

567

Hammad, A., Zhang, C., Setayeshgar, S., and Asen, Y. (2012). "Automatic generation of dynamic virtual fences as

568

part of BIM-based prevention program for construction safety." Simulation Conference (WSC 2012), IEEE, Berlin,

569

Germany, 1-10.

570

Haslam, R. A., Hide, S. A., Gibb, A. G., Gyi, D. E., Pavitt, T., Atkinson, S., and Duff, A. (2005). "Contributing factors

571

in construction accidents." Applied ergonomics, 36(4), 401-415.

572

Heinrich, H. W. (1939). Industrial Accident Prevention, McGraw-Hill, New York.

573

Hide, S., Atkinson, S., Pavitt, T. C., Haslam, R., Gibb, A. G., and Gyi, D. E. (2003). "Causal Factors in Construction

574

Accidents." . (May 11, 2015).

575

Holt, S. J. A. (2001). "Principles of construction safety." Construction safety, Blackwell Science, Oxford OX4 2DQ,

576

UK.

577

Kass, R. E., and Wasserman, L. (1993). "Formal rules for selecting prior distributions: A review and annotated

578

bibliography." Rapport interne(583).

579

Kass, R. E., and Wasserman, L. (1996). "The selection of prior distributions by formal rules." Journal of the

580

American Statistical Association, 91(435), 1343-1370.

581

Khanzode, V. V., Maiti, J., and Ray, P. K. (2012). "Occupational injury and accident research: A comprehensive

582

review." Safety Science, 50(5), 1355-1367.

583

Kim, H., Lee, H.-S., Park, M., Chung, B., and Hwang, S. (2016). "Automated hazardous area identification using

584

laborers' actual and optimal routes." Automation in Construction, 65, 21-32.

585

Kiviniemi, M., Sulankivi, K., Kähkönen, K., Mäkelä, T., and Merivirta, M.-L. (2011). "BIM-based Safety Management

586

and Communication for Building Construction." . (May

587

11, 2015).

588

Kruschke, J. (2014). Doing Bayesian Data Analysis: A Tutorial with R, JAGS, and Stan, Academic Press, 32

589

Jamestown Road, London NWI 7BY, UK.

590

Kumamoto, H., and Henley, E. (1996). Probabilistic Risk Assessment and Management for Engineers and Scientists,

591

IEEE Press.

592

Li, H., Chan, G., Huang, T., Skitmore, M., Tao, T. Y. E., Luo, E., Chung, J., Chan, X. S., and Li, Y. F. (2015). "Chirp-

593

spread-spectrum-based real time location system for construction safety management: A case study."

594

Automation in Construction, 55(2015), 58-65.

595

Luo, X., Li, H., Huang, T., and Rose, T. (2016a). "A field experiment of workers’ responses to proximity warnings

596

of static safety hazards on construction sites." Safety Science, 84, 216-224.

597

Luo, X., Li, H., Huang, T., and Skitmore, M. (2016b). "Quantifying Hazard Exposure Using Real-Time Location Data

598

of Construction Workforce and Equipment." Journal of Construction Engineering and Management, 04016031.

599

Manuele, F. A. (2005). "Risk Assessment & Hierarchies of Control." Professional Safety, 50(5), 33-39.

600

Mitropoulos, P., and Namboodiri, M. (2010). "New Method for Measuring the Safety Risk of Construction

601

Activities: Task Demand Assessment." Journal of Construction Engineering and Management, 137(1), 30-38.

602

Navon, R., and Kolton, O. (2006). "Model for Automated Monitoring of Fall Hazards in Building Construction."

603

Journal of Construction Engineering and Management, 132(7), 733-740.

604

Occupational Safety and Health Branch (2015). "Occupational Safety and Health Statistics 2014."

605

. (May 5, 2016).

606

Occupational Safety and Health Branch (2016). "Summary of Occupational Safety and Health Statistics of 2015."

607

. (May 5, 2016).

608

Safe Work Australia (2011). "How to Manage Work Health and Safety Risks: Code of Practice."

609

. (May 5, 2016).

610

Salvatier, J., Wiecki, T., and Fonnesbeck, C. (2016). "Probabilistic Programming in Python using PyMC3." PeerJ

611

Computer Science, 2(e55), 1-20.

612

Speed, T. (2011). "A Correlation for the 21st Century." Science, 334(6062), 1502-1503.

613

Summala, H. (2000). "Brake Reaction Times and Driver Behavior Analysis." Transportation Human Factors, 2(3),

614

217-226.

615

Teizer, J., and Cheng, T. (2015). "Proximity hazard indicator for workers-on-foot near miss interactions with

616

construction equipment and geo-referenced hazard areas." Automation in Construction, 60, 58-73.

617

Zhang, S., Lee, J.-K., Venugopal, M., Teizer, J., and Eastman, C. (2011). "Integrating BIM and safety: an automated

618

rule-based checking system for safety planning and simulation." Proceedings of CIB W099, CIB, Washington, DC,

619

24-26.

620

Zhang, S., Sulankivi, K., Kiviniemi, M., Romo, I., Eastman, C. M., and Teizer, J. (2015). "BIM-based fall hazard

621

identification and prevention in construction safety planning." Safety Science, 72, 31-45.

622

Zhang, S., Teizer, J., Lee, J.-K., Eastman, C. M., and Venugopal, M. (2013). "Building Information Modeling (BIM)

623

and Safety: Automatic Safety Checking of Construction Models and Schedules." Automation in Construction, 29,

624

183-195.

625

Table 1 Experimental settings and procedure, and recorded warnings and responses Zone Zone No.

Date

Num. of

Num. of

Zone Name

Zone No. Type

Warnings

Responses

Date

Zone

Num. of

Num. of

Type

Warnings

Responses

Zone Name

1

12-Jan-15

Dangerous goods

8

3

1

26

20-Mar-15

Dangerous goods

8

64

31

2

12-Jan-15

Welding area

1

12

8

27

20-Mar-15

Power distribution box

3

209

91

3

12-Jan-15

*Aerial work

6

77

32

28

20-Mar-15

Laydown area

2

191

104

4

19-Jan-15

Unprotected edge A

4

211

94

29

23-Mar-15

Lifting area

9

4

2

5

19-Jan-15

Unprotected edge B

4

55

18

30

23-Mar-15

Easy-to-stumble area

5

80

24

6

19-Jan-15

Lifting area

9

62

25

31

23-Mar-15

Welding area

1

32

14

7

19-Jan-15

Welding area

1

93

42

32

24-Mar-15

Lifting area

9

71

36

8

20-Jan-15

Unprotected edge A

4

312

125

33

24-Mar-15

Easy-to-stumble area

5

51

25

9

20-Jan-15

Unprotected edge B

4

58

19

34

24-Mar-15

Laydown area

2

69

43

10

20-Jan-15

Lifting area

9

54

18

35

25-Mar-15

Lifting area

9

63

34

11

20-Jan-15

Welding area

1

81

44

36

25-Mar-15

Easy-to-stumble area

5

97

39

12

21-Jan-15

Unprotected edge A

4

494

207

37

25-Mar-15

Laydown area

2

254

175

13

21-Jan-15

Unprotected edge B

4

49

24

38

31-Mar-15

Laydown area

2

104

67

14

21-Jan-15

Lifting area

9

72

30

39

3-Apr-15

Dangerous goods

8

168

92

15

21-Jan-15

Welding area

1

184

87

40

3-Apr-15

*Floor opening A

7

185

151

16

6-Feb-15

Unprotected edge

4

307

146

41

3-Apr-15

*Floor opening B

7

28

26

17

6-Feb-15

Dangerous goods

8

220

89

42

3-Apr-15

*Floor opening C

7

6

6

18

6-Feb-15

Lifting area

9

360

157

43

8-Apr-15

Power distribution box

3

39

27

19

17-Mar-15

Unprotected edge

4

23

13

44

8-Apr-15

Dangerous goods

8

16

6

20

17-Mar-15

Dangerous goods

8

63

39

45

8-Apr-15

Power distribution box

3

118

89

21

18-Mar-15

Lifting area

9

74

37

46

15-Apr-15

Lifting area

9

146

71

22

18-Mar-15

Welding area

1

58

26

47

15-Apr-15

*Floor opening

7

33

23

23

19-Mar-15

Power distribution box

3

120

52

48

15-Apr-15

Power distribution box

3

14

10

24

19-Mar-15

Laydown area

2

77

62

49

17-Apr-15

*Floor opening

7

16

15

25

19-Mar-15

Dangerous goods

8

188

98

50

17-Apr-15

Power distribution box

3

26

18

Zone type: 1, Welding areas; 2, Laydown areas; 3, Power distribution boxes; 4, Unprotected edges; 5, Easy-to-stumble areas; 6, Aerial work areas; 7, Floor openings; 8, Dangerous goods; 9, Lifting areas. ‘*’ indicates that proximity warnings are used as the secondary safety control measure in those hazard areas; the hazard areas without ‘*’ use proximity warnings as the primary control measure.

Table 2 Quantiles of parameter estimates Lower Parameter

Upper Median

Quantile

Lower Parameter

Quantile

Upper Median

Quantile

Quantile

𝜔 ̂

0.503

0.528

0.552

𝜃̂16

0.447

0.463

0.480

𝜅̂

14.350

19.566

26.785

𝜃̂17

0.410

0.431

0.453

𝜔 ̂1

0.466

0.487

0.510

𝜃̂18

0.425

0.441

0.457

𝜅̂ 1

72.322

105.634

147.619

𝜃̂19

0.424

0.456

0.490

𝜔 ̂2

0.626

0.648

0.672

𝜃̂20

0.520

0.552

0.586

𝜅̂ 2

41.774

63.856

94.026

𝜃̂21

0.448

0.474

0.501

𝜔 ̂3

0.560

0.589

0.621

𝜃̂22

0.442

0.473

0.504

𝜅̂ 3

19.089

27.688

39.021

𝜃̂23

0.432

0.462

0.492

𝜔 ̂4

0.414

0.430

0.446

𝜃̂24

0.702

0.733

0.763

𝜅̂ 4

76.438

109.140

151.330

𝜃̂25

0.494

0.514

0.535

𝜔 ̂5

0.381

0.412

0.443

𝜃̂26

0.464

0.494

0.525

𝜅̂ 5

51.257

79.079

118.147

𝜃̂27

0.430

0.453

0.476

𝜔 ̂6

0.412

0.455

0.501

𝜃̂28

0.548

0.572

0.593

𝜅̂ 6

53.160

81.782

126.450

𝜃̂29

0.423

0.461

0.496

𝜔 ̂7

0.768

0.795

0.821

𝜃̂30

0.324

0.356

0.388

𝜅̂ 7

50.994

80.145

118.544

𝜃̂31

0.442

0.476

0.507

𝜔 ̂8

0.480

0.501

0.522

𝜃̂32

0.452

0.479

0.506

𝜅̂ 8

55.432

81.035

117.973

𝜃̂33

0.408

0.444

0.482

𝜔 ̂9

0.441

0.458

0.475

𝜃̂34

0.607

0.636

0.665

𝜅̂ 9

77.079

107.227

151.841

𝜃̂35

0.460

0.488

0.517

𝜃̂1

0.452

0.495

0.537

𝜃̂36

0.377

0.406

0.434

𝜃̂2

0.469

0.507

0.548

𝜃̂37

0.661

0.680

0.697

𝜃̂3

0.401

0.435

0.470

𝜃̂38

0.617

0.644

0.671

𝜃̂4

0.421

0.441

0.460

𝜃̂39

0.508

0.531

0.555

𝜃̂5

0.367

0.396

0.425

𝜃̂40

0.790

0.808

0.825

𝜃̂6

0.410

0.438

0.467

𝜃̂41

0.798

0.829

0.861

𝜃̂7

0.444

0.471

0.495

𝜃̂42

0.768

0.807

0.842

𝜃̂8

0.393

0.410

0.425

𝜃̂43

0.604

0.647

0.689

𝜃̂9

0.367

0.395

0.423

𝜃̂44

0.440

0.480

0.518

𝜃̂10

0.387

0.418

0.445

𝜃̂45

0.694

0.721

0.749

𝜃̂11

0.486

0.511

0.540

𝜃̂46

0.453

0.474

0.497

𝜃̂12

0.408

0.421

0.436

𝜃̂47

0.726

0.763

0.795

𝜃̂13

0.421

0.449

0.479

𝜃̂48

0.579

0.627

0.682

𝜃̂14

0.414

0.441

0.468

𝜃̂49

0.782

0.818

0.852

𝜃̂15

0.457

0.478

0.500

𝜃̂50

0.589

0.637

0.681

Figures

Fig. 1. Safety risk management cycle Fig. 2. Decision tree for reviewing control measures Fig. 3. Bayesian hierarchical model Fig. 4. System deployment in the experiment Fig. 5. Trace plot of several selected parameters, where the burn-in step number is 1,000 and the thin value 20 resulting in there are only (51,000 – 1,000)/20 = 2,500 steps shown. Fig. 6. Box-and-whisker plot of ERRs (omegas and thetas), where ‘*’ indicates that proximity warnings are used as the secondary safety control measure in those hazard areas; those without ‘*’ use proximity warnings as the primary control measure. The red dashed line indicates the ERR of the generic hazards, ̂ (0.528), and the red horizontal span represents its interquartile range, which is the difference namely 𝝎 between the lower quartile (0.503) and the upper quartile (0.552). Similarly, the yellow dash-dot line in each column shows each hazard type ERR and the yellow horizontal span represents the interquartile range between its lower and upper quartiles. For example, in the first column representing the welding ̂ 𝟏 ) was about 0.487 with its interquartile range between the lower quartile areas, their hazard type ERR (𝝎 (0.466) and the upper quartile (0.510), which is represented by the yellow horizontal span. The estimates of individual hazards are represented in the form of boxes with whiskers.