Author's Accepted Manuscript

Validation of airway resistance Models for predicting pressure loss through anatomically realistic conducting airway Replicas of Adults and children Azadeh A.T. Borojeni, Michelle L. Noga, Andrew R. Martin, Warren H. Finlay

www.elsevier.com/locate/jbiomech

PII: DOI: Reference:

S0021-9290(15)00207-9 http://dx.doi.org/10.1016/j.jbiomech.2015.03.035 BM7117

To appear in:

Journal of Biomechanics

Accepted date: 27 March 2015 Cite this article as: Azadeh A.T. Borojeni, Michelle L. Noga, Andrew R. Martin, Warren H. Finlay, Validation of airway resistance Models for predicting pressure loss through anatomically realistic conducting airway Replicas of Adults and children, Journal of Biomechanics, http://dx.doi.org/10.1016/j.jbiomech.2015.03.035 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting galley proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1 2

Validation of Airway Resistance Models for Predicting Pressure Loss through

3

Anatomically Realistic Conducting Airway Replicas of Adults and Children

4

Azadeh A. T. Borojeni1,*, Michelle L. Noga2, Andrew R. Martin1’**, Warren H. Finlay1’**

5 6 7 8

9

10

11

1.

Department of Mechanical Engineering, University of Alberta, Edmonton, Alberta, Canada T6G 2G8

12

2.

Department of Radiology and Diagnostic Imaging, University of Alberta, Edmonton, Alberta, Canada

13 14

*

15

+1 780 492 2200

16

**

17

+1 780 492 4707; Fax: +1 780 492 2200

Corresponding author (during review and publication), Email: [email protected]; Tel: +1 780 492 2161; Fax:

Principal corresponding authors (post-publication), Email: [email protected], [email protected]; Tel:

18

19 20

Keywords: Airway resistance; conducting airways; pressure drop; pressure loss; ventilation

21

distribution

1

22

Abstract

23 24

This work describes in vitro measurement of the total pressure loss at varying flow rate through

25

anatomically realistic conducting airway replicas of ten children, 4 to 8 years old, and five adults.

26

Experimental results were compared with analytical predictions made using published airway

27

resistance models. For the adult replicas, the model proposed by van Ertbruggen et al. (J. Appl.

28

Physiol. 98:970-980,2005) most accurately predicted central conducting airway resistance for

29

inspiratory flow rates ranging from 15 to 90 L/min. Models proposed by Pedley et al. (J. Respir.

30

Physiol. 9:371-386,1970) and by Katz et al. (J. Biomechanics 44:1137-1143,2011) also provided

31

reasonable estimates, but with a tendency to over predict measured pressure loss for both models.

32

For child replicas, the Pedley and Katz models both provided good estimation of measured

33

pressure loss at flow rates representative of resting tidal breathing, but under predicted measured

34

values at high inspiratory flow rate (60 L/min). The van Ertbruggen model, developed based on

35

flow simulations performed in an adult airway model, tended to under predict measured pressure

36

loss through the child replicas across the range of flow rates studied (2 to 60 L/min). These

37

results are intended to provide guidance for selection of analytical pressure loss models for use in

38

predicting airway resistance and ventilation distribution in adults and children.

39 40 41

2

42

1. Introduction

43

Airway resistance is an important parameter used in describing the mechanics of breathing. The

44

relationship between pressure loss and flow rate through the conducting airways influences work

45

of breathing, and is widely used for diagnostic purposes in detecting and classifying respiratory

46

diseases. In addition, variation in airway resistance between individual airways or groups of

47

airways along different paths through the lung plays a role in determining the distribution of

48

ventilation between different lung regions. Ventilation distribution in turn influences regional

49

gas transport, mixing, and uptake, as well as transport and deposition of inhaled aerosols.

50

Analytical models of airway resistance have been employed recently, for example, in modeling

51

topographic distribution of ventilation in healthy lungs (Swan et al., 2012), in estimating

52

contributions of peripheral airway resistance to ventilation distribution measured by PET

53

imaging (Wongviriyawong et al., 2013), in predicting distribution of pressure through the airway

54

tree during xenon anesthesia (Katz et al., 2012), and in optimizing algorithms used to predict

55

alveolar pressure during mechanical ventilation (Damanhuri et al., 2014).

56 57

Study of pressure loss over the conducting airways dates to the foundational work of Rohrer

58

(1915). For a given flow rate, the difference in pressure between the airway opening (nose or

59

mouth) and the alveoli was expressed as a summation of linear and quadratic terms, each term

60

weighted by a dimensional constant.

61

depending on airway dimensions and geometry. The utility of such an approach was clearly

62

demonstrated by Otis and colleagues, who used the Rohrer airway resistance model in

63

combination with experimental measurements to describe work of breathing (Otis et al., 1950),

64

ventilation distribution (Otis et al., 1956), and effects of breathing different gas mixtures on

Values of these constants varied between subjects,

3

65

airway resistance (Otis et al., 1949). These studies provided early insight into the functional

66

relationship between pressure loss, flow rate, gas properties, and airway geometry. However,

67

significant further understanding was gained through fundamental experiments and fluid

68

dynamics analysis performed by Pedley and colleagues (1970). These authors presented an

69

analytical model of pressure loss through conducting airways of the human bronchial tree based

70

on measurements performed in idealized, symmetric airway bifurcations (Pedley et al., 1970).

71

Pressure loss was found to exceed that predicted by assuming fully developed Hagen-Poiseuille

72

flow, and a correction factor was proposed based on airway length and diameter, and the gas

73

Reynolds number (Pedley et al., 1970).

74 75

More recently, computational fluid dynamics (CFD) simulations have been performed to

76

investigate pressure loss through idealized (e.g. Wilquem and Degrez, 1997; Comer et al., 2001;

77

Kang et al., 2011; Katz et al., 2011) and anatomically realistic (e.g. van Ertbruggen et al., 2005;

78

Gemci et al., 2008; Ismail et al., 2013) adult airway bifurcations. Previous authors have noted

79

that, for adult airways, the Pedley model overestimates simulated pressure losses (Comer et al.,

80

2001; van Ertbruggen et al., 2005; Katz et al., 2011; Ismail et al., 2013).

81

alternative analytical models have been proposed (van Ertbruggen et al., 2005; Katz et al., 2011).

82

In contrast, considerably less work has been done to model airway pressure losses for the

83

conducting airways of children. The accuracy of using analytical models developed based on

84

measurements or simulations performed in adult airway models in predicting pressure loss in

85

children’s airways is presently unknown.

86

4

Accordingly,

87

In this paper, we describe in vitro measurements of the total pressure loss at varying flow rate

88

through anatomically realistic conducting airway replicas of adults and children. The in vitro

89

approach employed presents a unique opportunity to reduce the pressure-flow relationship

90

through the three-dimensional, multi-branching, conducting airway geometry to the classic fluid

91

mechanics problem of flow through a piping network comprising nonlinear resistances (White

92

2010). Accordingly, analytical predictions of the total pressure loss through the airway replicas

93

made using airway resistance models proposed by Pedley et al. (1970), van Ertbruggen et al.

94

(2005), and Katz et al. (2011) are compared to measured experimental values. Our results are

95

intended to provide guidance for selection of analytical models for use in predicting airway

96

resistance in adults and children.

97 98

2.

Methods

99 100

2.1 Airway pressure drop measurement

101 102

Fabrication of realistic adult and child conducting airway replicas has been explained in detail in

103

previous work (Borojeni et al., 2014). In brief, conducting airways were identified using Mimics

104

software (MIMICS 3D, Materialise, Ann Arbor, MI, USA) as regions of interest in CT-images of

105

5 adults and 10 children. Subject information and average airway dimensions are shown in

106

Table 1. A rapid prototyping technique (Invision® SR 3D printer, 3D systems, Rock Hill, SC,

107

USA) was used to fabricate the conducting airway replicas from acrylonitrile butadiene styrene

108

(ABS) plastic (P430, Stratasys, Eden Prairie, MN).

109

5

110

In the present work, pressure drop across the airway replicas was measured in vitro using a

111

digital pressure gauge (HHP-103, Omega Engineering, Inc., Stamford, CT) with low range (0-

112

498 Pa), high range (0-2590 Pa) and accuracy of ±0.2% full scale. The upstream pressure tap

113

was open to atmosphere and the downstream pressure tap was positioned in the tubing that

114

connected the outlet of an artificial thoracic chamber to a vacuum pump (Figure 1). Preliminary

115

measurements were made with the chamber open to characterize the pressure drop across only

116

the chamber outlet at flows ranging from 2 to 90 L/min. Constant flow rates were then drawn

117

through the chamber with the replicas in place by operating a vacuum pump, with flow rates

118

chosen in the range of 15 to 90 L/min for adult replicas and 2 to 60 L/min for child replicas. The

119

desired flow rate was set using a needle valve and monitored by a digital mass flowmeter (TSI

120

Model 4000 series, TSI Inc., Shoreview, MN, USA) downstream of the thoracic cylindrical

121

chamber. The pressure drop across the airway replicas is reported below as the difference

122

between the pressure drop measured with and without the airway replicas in place. Each

123

experimental measurement was repeated three times.

124 125

2.2 Analytical pressure loss model

126 127

Measured values of pressure drop across the airway replicas were compared with predicted

128

values made using three previously published models of airway resistance (Pedley et al., 1970;

129

Katz et al., 2011; van Ertbruggen et al., 2005). For each replica, subject-specific geometry of the

130

conducting airways was required as input in the analytical models of flow distribution and

131

resistance.

132

6

133 134

2.2.1 Determination of Airway Dimensions

135

Diameters and lengths for all airway segments in each replica were measured from segmented

136

CT images using a - post processing software package (MAGICS, Materialise, MI, USA). The

137

length of a segment was defined as the centerline path length from one bifurcation to the next.

138

The diameter was defined as the average of equivalent cross-sectional area diameters evaluated

139

every 2 mm along the length of the segment.

140 141

2.2.2 General Solution Procedure

142 143

For convenience of calculation, a hydraulic resistance, R, is defined for a segment as R =

144

∆h



(1)

145

146

where Q is the flow rate through the airway and ∆h is a form of the head loss, expressed in

147

dimensions of energy per unit mass as the ratio between pressure drop across the airway (∆p) and

148

fluid density (ρ), i.e. Δh=Δp/ρ .

149

Expressions for the head loss as a function of airway geometry, fluid properties and flow rate

150

result from the application of specific airway resistance models. These will be provided in the

151

following section. At present, Equation 1 is used to outline the solution procedure in general.

152

For example, for the simple network of segments shown in Fig. 2 we wish to determine the

153

fraction of the total flow entering airway 0 that flows through each branch, 1a and 1b. To do so,

154

the resistances of branches 1a and 1b must be taken into account, as must be the resistances of all 7

155

subsequent segments downstream. The equivalent hydraulic resistance of either branch can be

156

determined as follows (White, 2010): eq, 1a =  +

/

1

/

[ 2aa + 2ab ] (2)

157

158

eq, 1b =  +

/ [ 2ba

1

/

+ 2bb ] (3)

159

160

In the tracheobronchial tree, the distal airways shown in Figure 2 will proceed to bifurcate into

161

further downstream airway generations. It is noted that hydraulic resistances of these

162

downstream segments (i.e. R2aa, R2ab, R2ba, or R2bb) may themselves be expressed as equivalent

163

hydraulic resistance defined using these same principles.

164 165

The head loss through any branch can then be expressed as

166 167

∆hn = R eq,n Q2n (4)

168

169

where the equivalent hydraulic resistance Req,n of that branch is formulated by generalization of

170

equations 2 and 3 and is itself a function of Qn.

171

8

172

Applying this procedure to a branching airway network results in a system of nonlinear

173

equations. For a multiple-path network with n paths, a system of n equations is obtained. This

174

system of equations can be solved iteratively provided that the head loss is the same across each

175

path. This condition is met in the experiments described above given that all terminal branches

176

supply the artificial thorax chamber. The head loss calculation for each path included a minor

177

loss term accounting for dissipation of kinetic energy due to discharge of flow from terminal

178

branches into the artificial thorax. This term was in all cases small compared with the total head

179

loss.

180 181

The iterative solution procedure starts with an initial guess for the flow rate in each branch. This

182

was made assuming evenly distributed flow between daughter branches at all bifurcations. The

183

flow rates through each branch at subsequent iterations may then be calculated according to:

184

a 

=

/

( eq,b )# (Qinlet ) /



/

[( eq,a )# + ( eq,b )# ] (5)

185

186

where i+1 and i denote the current and previous iteration, respectively, and subscripts a and b

187

indicate daughter airways. Equivalent hydraulic resistances are flow rate dependent and thus

188

update on each iteration.

189 190

After each iteration, flow rates along each path were compared with corresponding values from

191

the previous iteration. The root mean square (RMS) difference in flow rates between successive

192

iterations was used to determine convergence. The solution was deemed converged when this 9

193

RMS difference fell below 0.006 L/min. The sensitivity of the total head loss to this criterion

194

was trivial for values less than 0.006 L/min. For the converged solution, the total head loss

195

across the airway tree was then calculated using the branch flow rates and resistances.

196 197 198

2.2.3

Airway Resistance Models

199

2.2.3.1 Pedley Model

200

Pedley and colleagues (1970) determined the pressure drop through idealized symmetric models

201

of bronchial airways based on measurement of changes to inspiratory flow profiles. They

202

derived an expression for the ratio relating the actual pressure drop through an airway section to

203

that which would be predicted assuming fully-developed Hagen-Poiseuille flow. This ratio was

204

given by:

205 / 

1.85

, $= * + . 4√2

(6)

206

207 208

where D and L are airway diameter and length, and Re is the flow Reynolds number:

209

+ =

4 0,1 (7)

210

211

where 1 is the kinematic viscosity of the fluid (air). 10

212 213

The pressure drop is then calculated as: ∆2pedley = $∆2Hagen-Poiseuille (8)

214

215

216

where, ∆2Hagen-Poiseuille =

128µ- 0,? (9)

217

218

where µ is the fluid dynamic viscosity.

219

Accordingly, the expression for hydraulic resistance, R in a given airway branch for the Pedley

220

model is as follows:

pedley =

128 ∙ 1.85 - A.B A.B * . + √20  ,? , (10)

221

222

where ρ is fluid density.

223

Note that in analyzing the realistic airway replicas, the value of Z was less than 1 at some distal

224

airways where airway diameter was considerably smaller than the airway length and Reynolds

225

number was small. This is not a physically realistic result as the dissipation of energy in realistic

11

226

airways cannot be less than dissipation of energy in the ideal Hagen-Poiseuille flow. Therefore,

227

the value of $ was replaced by 1 in these cases.

228

229

2.2.3.2 Modified Pedley Model (van Ertbruggen model)

230

van Ertbruggen and colleagues ( 2005) used CFD to simulate flow through an anatomically-

231

based model of the airways extending from the trachea to the segmental bronchi. This model

232

combined morphometric data data from Horsfield et al. (1971) with orientation of branching

233

planes based on bronchoscope and CT images. Simulated pressure drop through each airway

234

generation in the model was compared by van Ertbruggen and colleagues ((2005) to values

235

predicted using the Pedley model described above. Pressure drop predicted using the Pedley

236

model was consistently higher than simulated values. The authors therefore proposed a modified

237

form of equation 8, as follows:

∆2ModiDied-Pedley

/  32µ-G

, = E * + . -

,

238

(11)

239

where values of γ for generations 0-3 are 0.162, 0.239, 0.244 and 0.295 respectively (van

240

Ertbruggen et al., 2005).

241

Using these values for γ, the expression for hydraulic resistance, R, in a given airway branch for

242

the modified Pedley model is: modiDied Pedley =

512 - A.B E ( ) 0  ,? , + A.B

12

(12)

243

244

2.2.3.3 Katz Model

245

Katz and colleagues (2011) presented a model for calculating pressure drop through the

246

conducting airways based on concepts of major and minor head losses.

247

The expression for hydraulic resistance, R, in a given airway branch for the Katz model is:

248

Katz =

8

0  ,? L

[M

+ N] , (13)

249

250

where M is the friction factor, assumed to be 64/Re for Re < 2000 (laminar flow). Above this

251

value, flow is assumed turbulent and the friction factor is calculated using the Blasius correlation

252

for smooth tubes: M=

0.316 + A.B

253

(14)

254

The minor loss coefficient (N) in Equation 13, is calculated using the following correlation based

255

on CFD simulations of flow through idealized symmetric airway bifurcations (Katz et al., 2011): N=

Q +S(log +) + ,(log +) + T + R (15)

256

13

257

Values for the constants A through E in Equation 15 are provided for each tracheal-bronchial

258

lung generation by Katz and colleagues (2011).

259

For each of the three airway resistance models, prediction of measured pressure drop through the

260

adult and child airway replicas was assessed using Lin’s concordance correlation coefficient, ρc

261

(Lin 1989). In the present work, reported values of the concordance correlation coefficient

262

indicate goodness of fit of plotted predicted vs. measured data to the identity line, with higher

263

values indicating better fit and ρc = 1.0 indicating a perfect fit.

264

265

3. Results

266 267

Tracheal Reynolds numbers for adult and child subjects are presented in Table 2. At an

268

equivalent volumetric flow rate, Reynolds numbers were higher in child than in adult replicas,

269

owing to the smaller tracheal diameters (and thus higher flow speeds) of the child replicas.

270

Comparison of measured versus calculated pressure drop through the adult replicas is made in

271

Figures 3 for the three airway resistance models employed. These comparisons indicate there is,

272

in general, reasonable agreement between measured and predicted values, for all three resistance

273

models. It can be further observed that the experimental data are in best agreement with the

274

predicted values made using the modified Pedley model.

275 276

Comparison of measured versus calculated pressure drop through the child replicas is

277

demonstrated in Figures 4-6. Here the Katz model and original Pedley model displayed better

14

278

agreement with measured values than did the modified Pedley model, though all three models

279

had a tendency to under predict measured data for the highest measured values of pressure drop,

280

corresponding to higher flow rates through the replicas.

281 282

In calculating the pressure drop across the replicas, the distribution of flow through the

283

branching airways is also calculated. Calculated values of the fraction of flow through the right

284

lung are presented in Table 3. For both adults and child replicas, a greater fraction of the

285

inhalation flow passes through airway of the right lung. The standard deviation between replicas

286

was higher in the child group than in the adult group (0.15 versus 0.03 on average, respectively).

287 288 289

4. Discussion

290 291

In the present work, we measured pressure loss at varying flow rate through conducting airway

292

replicas of adults and children. Measurements were compared with values predicted using

293

analytical airway resistance models previously developed based on measurements or simulations

294

performed in adult conducting airway geometries. For the present five adult airways, the airway

295

resistance model proposed by Pedley and colleagues (1970) tended to over-predict measured

296

pressure loss. This is consistent with the results of previous CFD studies comparing predictions

297

of conducting airway pressure loss made using the Pedley model to simulated values (Comer et

298

al., 2001; van Ertbruggen et al., 2005; Katz et al., 2011; Ismail et al., 2013). Of the three

299

resistance models evaluated, the modified Pedley model proposed by van Ertbruggen et al.

300

(2005) demonstrated closest agreement with our measured adult data. These authors modified

301

the original Pedley model by proposing alternative coefficients for predicting the pressure loss 15

302

through each proximal lung generation. The modified coefficients were based on pressure losses

303

determined by CFD simulation of laminar flow through a single, representative, anatomically-

304

based adult airway geometry. The good agreement of predicted pressure losses made using their

305

model with our data in five different adult airway replicas, at tracheal Reynolds numbers

306

exceeding 10,000 in the highest case (Table 2), was therefore not expected a priori. However,

307

van Ertbruggen et al. (2005) argue that below Reynolds number averaging 8,000 the flow regime

308

through conducting airways can be considered laminar or ‘lightly’ transitional, but not turbulent.

309

Again with reference to Table 2, for a strong majority of cases we studied tracheal Reynolds

310

numbers in the adult replicas that fell below 8,000.

311

angle (and branching asymmetry) are not explicitly accounted for in the modified Pedley model,

312

and surely existed between our five adult replicas, Kang et al. (2011) found only a weak

313

influence of branching angle on simulated pressure loss through a model airway bifurcation.

Furthermore, while variation in branching

314 315

For the child airway replicas, predicted and measured pressure losses were in closer agreement

316

for the Katz model and the original Pedley model than for the modified Pedley model. Poorer

317

agreement of the modified Pedley model with the child versus adult data may be due to

318

differences in central conducting airway geometry between children and the single adult

319

geometry used in creating this model, that are not captured explicitly in Equation 12 (recalling

320

that model coefficients were fit to simulation data obtained in a representative single adult

321

airway geometry).

322 323

All three models tended to under predict measured pressure loss at the highest flow rate studied

324

in children (60 L/min), where tracheal Reynolds numbers averaged over 10,000. Indeed, for

16

325

replicas from the youngest children studied (ages 4 and 5), the pressure loss at 60 L/min was

326

sufficiently high that such an inspiratory flow rate is likely to be achieved in vivo by the

327

corresponding child subjects only near maximum inspiratory effort (Vilozni et al., 2009). At

328

lower flow rates, corresponding to lower pressure loss, the original Pedley model was the most

329

accurate of the three models studied. Pedley et al. (1970) based their model on measurements

330

made in idealized symmetric models of branching airways, where Reynolds number ranged from

331

~100 to ~1,000. Tracheal Reynolds numbers in our child replicas were on average within this

332

range for the two lowest flow rates studied (2 and 5 L/min).

333

relatively close agreement between our measured data and predictions made using the Pedley

334

model for low pressure loss (where flow rate and Reynolds number were also low) and lack of

335

such agreement for higher pressure loss.

This very likely explains the

336 337

In calculating the total pressure loss through the airway replicas using each of the three airway

338

resistance models, the distribution of flow through the replicas is also obtained. The fraction of

339

flow directed to the right lung side of the replicas was found to average 0.58 in the adult replicas,

340

with little variation with flow rate or between replicas.

341

distribution of ventilation to the right lung observed in vivo (Lumb 2010), especially given that

342

for the intact, healthy, in vivo lung, ventilation distribution is determined more by downstream

343

tissue compliance than by airway resistance (Swan et al., 2012). Ostensibly, the dimensions of

344

the central conducting airways act to reinforce, rather than oppose, this ventilation distribution.

345

For the child replicas, the fraction of flow directed to the right lung side of the replicas

346

(averaging 0.63) did not vary significantly from that for the adult replicas. However, variation

347

between replicas was considerably greater for child than for adult replicas. This variation is

17

This value is consistent with the

348

potentially attributable to a greater variation in airway geometry for the child replicas, given that

349

these geometries were obtained from images of children aged 4-8 years, and therefore at

350

different stages of physical development.

351 352

A limitation in the present work is the absence of any upper, extrathoracic airway upstream from

353

our conducting airway replicas. In this regard, our replicas are similar to the original single and

354

multiple bifurcation geometries used to develop the three analytical resistance models studied

355

herein. None of these models was based on measurements or simulations done in airway

356

geometries that included the upper airway.

357

performed by Xi et al. (2008), the upper airway most significantly influences flow features

358

within the trachea, whereas flow in subsequent downstream airway branches is much less

359

affected. The accuracy of the resistance models in predicting in vivo pressure loss through the

360

trachea, directly downstream from the upper airway, was not assessed in the present study.

As observed in numerical flow simulations

361 362

5. Conclusion

363

Pressure loss was experimentally measured at varying flow rate through anatomically realistic

364

conducting airway replicas of adults and children. Measured values were compared with values

365

predicted using published analytical airway resistance models. The modified Pedley model

366

proposed by van Ertbruggen et al. (2005) provided close agreement with measured central

367

conducting airway resistance in adult replicas across a wide range of inspiratory flow rates, from

368

15 to 90 L/min. In contrast, in child airway replicas, the Pedley et al. (1970) model most closely

369

matched the experimental data.

370

18

371 372

6. Acknowledgements

373

Andrew Grosvenor and Jamie Schmitt from the Institute for Reconstructive Sciences in Medicine

374

(iRSM) are acknowledged for their technical help in fabricating the replicas. Greg Wandzilak

375

from PET/CT center, University of Alberta, Stollery Children’s Hospital is gratefully

376

acknowledged for providing CT scans. Travis MacLean and John Chen are appreciated for their

377

help in measuring pressure drop in the child replicas.

378 379

7. Conflict of Interest

380

The authors report no conflict of interest.

381

19

382 383 384 385 386 387 388 389 390 391 392 393 394 395

8. References

396 397 398 399 400 401 402

van Ertbruggen, C., Hirsch, C., Paiva, M., 2005. Anatomically based three-dimensional model of airways to simulate ow and particle transport using computational fluid ynamics. Journal of Applied Physiology. 98, 970-980.

403 404 405 406 407 408 409 410 411 412

Horsfield, K., Dart, G., Olson, D.E., Filley, G.F., Cumming, G., 1971. Models of the human bronchial tree. Journal of Applied Physiology. 31, 207–217.

413 414 415 416 417 418 419 420 421 422 423 424 425

Katz, I.M., Martin, A.R., Muller, P.-A., Terzibachi, K., Feng, C.-H., Caillibotte, G., Sandeau, J., Texereau, J., 2011. The ventilation distribution of helium-oxygen mixtures and the role of inertial losses in the presence of heterogeneous airway obstructions. Journal of Biomechanics. 44, 1137-1143.

Borojeni, A.A., Noga, M.L., Vehring, R., Finlay, W.H., 2014. Measurements of total aerosol deposition in intrathoracic conducting airway replicas of children. Journal of Aerosol Science. 73, 39-47. Comer, J.K., Kleinstreuer C, Zhang Z., 2001. Flow structures and particle deposition patterns in double bifurcation airway models. 1. Airflow fields. Journal of Fluid Mechanics. 435, 25–54. Damanhuri, N.S., Docherty, P.D., Chiew, Y.S., van Drunen, E.J., Desaive, T., and Chase, J.G., 2014. A Patient-specific airway branching model for mechanically ventilated patients. Computational and Mathematical Methods in Medicine, Article ID 645732.

Gemci, T., Ponyavin, V., Chen, Y., Chen, H., Collins, R., 2008. Computational model of airflow in upper 17 generations of human respiratory tract. Journal of Biomechanics. 41(9), 2047-2054.

Ismail, M., Comerford, A., Wall, W.A., 2013. Coupled and reduced dimensional modeling of respiratory mechanics during spontaneous breathing. International Journal For Numerical Methods In Biomedical Engineering. 29, 1285-1305. Kang, M-Y, Hwang, J., and Lee, J-W., 2011. Effect of geometric variations on pressure loss for a model bifurcation of the human lung airways. Journal of Biomechanics. 44, 1196–1199.

Katz, I. M., Martin, A. R., Feng, C. H., Majoral, C., Caillibotte, G., Marx, T., Bazin, J.-E., and Daviet, C. 2012. Airway pressure distribution during xenon anesthesia: the insufflation phase at constant flow (volume controlled mode). Applied Cardiopulmonary Pathophysiology, 16, 5-16. Lin, IL. 1989. A concordance correlation coefficient to evaluate reproducibility. Biometrics, 45, 255-268.

20

426 427 428

Lumb, A.B., 2010. Nunn’s Applied Respiratory Physiology, Seventh Edition. Elsevier Ltd.

429

Otis A.B., Bembower W.C., 1949. Effect of gas density on resistance to respiratory gas

430 431

flow in man. Journal of Applied Physiology. 2, 300-306.

432

Otis A.B., Fenn W.O., Rahn H., 1950. Mechanics of Breathing in Man. Journal of

433

Applied Physiology. 2, 592-607.

434 435

Otis A.B., McKerrow C.B., Bartlett R.A., Mead J., McIlroy M.B., Selverstone N.J.,

436

Radford E.P., 1956. Mechanical factors in distribution of pulmonary ventilation. Journal

437 438 439 440 441 442 443 444 445

of Applied Physiology. 8, 427-443.

446 447 448 449 450 451 452 453 454 455 456 457 458

Swan, A.J., Clark, A.R., Tawhai, M.H., 2012. A computational model of the topographic distribution of ventilation in healthy human lungs. Journal of Theoretical Biology. 300, 222-231.

459 460 461

Wongviriyawong, C., Harris, R. S., Greenblatt, E., Winkler, T., and Venegas, J. G., 2013. Peripheral resistance: a link between global airflow obstruction and regional ventilation distribution. Journal of Applied Physiology, 114(4), 504-514.

462 463 464

Xi, J., Longest, P.W., and Martonen, T.B., 2008. Effects of the laryngeal jet on nanoand microparticle transport and deposition in an approximate model of the upper tracheobronchial airways. Journal of Applied Physiology, 104(6), 1761-1777.

Pedley, T.J., Schroter, R.C., Sudlow, M.F., 1970. Energy losses and pressure drop in models of human airways. Journal of Respiration Physiology. 9, 371-386. Rohrer, F., 1915. Der Strömungswiderstand in den menschlichen Atemwegen und der Einfluss der unregelmässigen Verzweigungen des Bronchialsystems auf den Atmungsverlauf in verschiedenen Lungenbezirken. Pflügers Arch ges Physiol, 162, 225–299.

Vilozni, D., Efrati, O., Barak, A., Yahav, Y., Augarten, A., Bentur, L., 2009. Forced inspiratory flow volume curve in healthy young children. Pediatric Pulmonology. 44(2), 105-111. White, F. M., 2010. Fluid Mechanics, 7th edition. Boston, USA: McGraw-Hill. Wilquem, F., Degrez. G., 1997. Numerical modeling of steady inspiratory airflow through a three-generation model of the human central airways. Journal of Biomechanical Engineering. 119(1), 59-65.

465

21

466 467 468

9. Tables

469

Table 1: Summary of child and adult subjects information and airway diameters of child and

470

adult replicas. Age

Subject

(year)

Number

4

Sex

Height

Weight

Diameter

Diameter

Diameter

Diameter

(cm)

(kg)

(stdev)

(stdev)

(stdev)

(stdev)

(Gen.0)

(Gen.1)

(Gen.2)

(Gen.3)

(mm)

(mm)

(mm)

(mm)

10c

F

99

16

7.15(0.04)

4.7(1.19)

3.48(0.27) 2.2(0.23)

14c

F

100

16

7.16(0.15)

6.47(2.77)

4.46(0.69) 2.23(0.76)

2c

M

117

22.9

7.05(0.09)

6.03(0.7)

6.00(1.4)

3c

M

112

20

7.99(0.51)

5.39(0.51)

4.93(0.49) 2.33(0.66)

9c

M

113

20

7.56(0.01)

6.44(0.87)

5.26(0.01) 3.38(1.05)

5c

F

112

18

7.99(0.42)

5.36(1.14)

5.35(1.41) 3.55(1.00)

6c

F

118

21.5

8.5(0.3)

6.75(1.32)

5.94(1.32) 2.9(1.47)

12c

F

124

24

7.41(0.46)

6.45(2.72)

3.28(0.26) 3.23(0.9)

7

13c

F

121

20

9.78(1.32)

7.64(0.09)

6.48(1.23) 3.58(0.01)

8

11c

M

124.5

24.5

10.49(0.95) 7.43(2.26)

6.23(1.42) 3.02(0.35)

50

7a

M

178

113

14.57(2.04) 12.34(1.08) 6.79(1.62) 3.6(0.15)

8a

F

155

99

12.4(0.01)

55

3a

F

159

68

14.94(2.67) 14(2.15)

9.45(1.65) 7.46(0.8)

62

4a

M

168

91

14.47(2.09) 13.63(2.5)

7.8(1.04)

80

5a

M

173

76

16.13(2.19) 14.27(2.3)

6.89(2.27) 4.47(1.87)

5

6

4.35(1.91)

10.83(2.68) 6.53(2.27) 5.18(0.23)

4.69(1.04)

471 472 473

Subject numbers ending with “c” indicate child subjects, while those ending with “a” indicate adult subjects.

474 475

Generation 0 (Gen.0) is the trachea, Gen.1 are the main bronchi, Gen.2 are the bronchi and Gen.3 are the segmental bronchi.

22

476 477

Stdev is the standard deviation of the range of diameters in each generation (Borojeni et al., 2014).

23

478 479

Table 2: Values of tracheal Reynolds number at different flow rate in different subjects

Subject 3a

Q (L/min) =2 ---

Q Q (L/min) (L/min) = = 10 5 -----

4U 0,trachea X Q Q (L/min) (L/min) = 30 = 15 1436 2872

Subject 4a

---

---

---

1483

2965

5931

8896

Subject 5a

---

---

---

1331

2661

5322

7984

Subject 7a

---

---

---

1473

2946

5893

8839

Subject 8a

---

---

---

1730

3460

6920

10380

Subject 9c

378

946

1892

2838

5676

11353

---

Subject 6c

337

842

1684

2526

5052

10104

---

Subject 5c

358

895

1790

2685

5371

10742

---

Subject 13c

292

731

1462

2193

4386

8773

---

Subject 2c

406

1014

2029

3043

6087

12174

---

Subject 3c

358

895

1790

2685

5370

10739

---

Subject 11c

273

682

1364

2046

4093

8185

---

Subject 10c

400

1000

2000

3000

6000

11999

---

Subject 12c

386

965

1930

2895

5790

11580

---

Subject 14c

400

999

1998

2997

5993

11987

---

Subject ID#

+ =

480 24

Q (L/min) = Q 60 (L/min) = 90 5744 8616

481

Table 3: Fraction of flow to the right lung. Data are presented as average (standard deviation).

482

N=5 for the adult replicas, and =10 for the child replicas.

483

Flow rate Katz (L/min) model

Adult Pedley model

Modified Pedley model -

Katz model

Child Pedley model

0.64(0.15)

0.63(0.15)

Modified Pedley model 0.63(0.15)

2

-

-

5

-

-

-

0.64(0.15)

0.63(0.15)

0.63(0.15)

10

-

-

-

0.64(0.15)

0.63(0.15)

0.63(0.15)

15

0.58(0.02)

0.58(0.03)

0.58(0.04)

0.63(0.15)

0.63(0.15)

0.63(0.15)

30

0.58(0.02)

0.58(0.03)

0.58(0.04)

0.62(0.13)

0.63(0.15)

0.62(0.15)

60

0.57(0.02)

0.58(0.03)

0.58(0.04)

0.61(0.12)

0.62(0.15)

0.62(0.15)

90

0.56(0.02)

0.58(0.04)

0.58(0.04)

484 485 486 487 488 489 490 491 492

25

-

-

-

493

9. Figure Captions

494 495 496

Figure 1: Experimental set-up used to measure pressure drop in adult and child CT-based

497

airway replicas.

498 499

Figure 2: Schematic of a simple branching airway network with numbering system used in the

500

analytical model.

501 502

Figure 3: Comparison of pressure drop data predicted by (a) the Katz et al. (2011) model, (b) the

503

Pedley et al. (1970) model, and (c) the modified Pedley model (van Ertbruggen et al., 2005) with

504

the measured data in adult replicas. The concordance correlation coefficient, ρc,between

505

modeled and experimental pressure drops is 0.84, 0.88, and 0.93 in panels (a), (b), and (c),

506

respectively.

507

508

Figure 4: (a) Comparison of pressure drop predicted by the Katz et al. (2010) model with the

509

measured data in child replicas. (b) Magnified view in the low range of pressure (0-100 Pa).

510

The concordance correlation coefficient, ρc, in panel (a) and (b) is 0.75 and 0.77, respectively.

511

512

Figure 5: (a) Comparison of pressure drop predicted by the Pedley et al. (1970) model with the

513

measured data in child replicas. (b) Magnified view in the low range of pressure (0-100 Pa). The

514

concordance correlation coefficient, ρc, in panel (a) and (b) is 0.61 and 0.84, respectively.

26

515

516

Figure 6: (a) Comparison of pressure drop predicted by the modified Pedley model (van

517

Ertbruggen et al., 2005) with the measured data in child replicas. (b) Magnified view in the low

518

range of pressure (0-100 Pa). The concordance correlation coefficient, ρc, in panel (a) and (b) is

519

0.43 and 0.65, respectively.

520

27

521 522 523 524 525

10. Figures 1

526 527 528 529

28

530 531 532 533

2

534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558

29

559 560

3 (a)

561 562 563

(b)

564 565 30

566

(c)

567

568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591

31

592 593

4 (a)

594 595

(b)

596 597 598 32

599 600 601 602 603

5 (a)

604 605

(b)

606 33

607 608 609 610 611 612

613 614

6 (a)

(b)

615

34

Validation of airway resistance models for predicting pressure loss through anatomically realistic conducting airway replicas of adults and children.

This work describes in vitro measurement of the total pressure loss at varying flow rate through anatomically realistic conducting airway replicas of ...
6MB Sizes 1 Downloads 7 Views