Sei sulla pagina 1di 105

University of Iowa

Iowa Research Online


Theses and Dissertations

Fall 2010

Computational fluid dynamics (CFD) study


investigating the effects of torso geometry
simplification on aspiration efficiency
Kimberly Rose Anderson
University of Iowa

Copyright 2010 Kimberly R. Anderson

This thesis is available at Iowa Research Online: http://ir.uiowa.edu/etd/774

Recommended Citation
Anderson, Kimberly Rose. "Computational fluid dynamics (CFD) study investigating the effects of torso geometry simplification on
aspiration efficiency." MS (Master of Science) thesis, University of Iowa, 2010.
http://ir.uiowa.edu/etd/774.

Follow this and additional works at: http://ir.uiowa.edu/etd

Part of the Occupational Health and Industrial Hygiene Commons


COMPUTATIONAL FLUID DYNAMICS (CFD) STUDY INVESTIGATING THE

EFFECT OF TORSO GEOMETRY SIMPLIFICATION ON ASPIRATION

EFFICIENCY

by
Kimberly Rose Anderson

A thesis submitted in partial fulfillment


of the requirements for the Master of
Science degree in Occupational and Environmental Health
in the Graduate College of
The University of Iowa

December 2010

Thesis Supervisor: Assistant Professor T. Renée Anthony


Copyright by

KIMBERLY ROSE ANDERSON

2010

All Rights Reserved


Graduate College
The University of Iowa
Iowa City, Iowa

CERTIFICATE OF APPROVAL

_______________________

MASTER'S THESIS

_______________

This is to certify that the Master's thesis of

Kimberly Rose Anderson

has been approved by the Examining Committee


for the thesis requirement for the Master of Science
degree in Occupational and Environmental Health at the December 2010
graduation.

Thesis Committee: ___________________________________


T. Renée Anthony, Thesis Supervisor

___________________________________
Thomas Peters

___________________________________
Patrick O’Shaughnessy
ACKNOWLEDGMENTS

First and foremost I would like to thank my advisor, Renée. I cannot possibly

thank you enough for providing me with the support and direction to complete this work.

You sparked my interest in research and set me on the path to pursuing my PhD, and I

will always be grateful to you for that. Thank you for indulging me on my various

interests and side projects and providing me with the knowledge and resources to pursue

those. And thank you for your guidance and focusing me on the tasks at hands when my

ambitions got the better of me. I will always be honored to call you my advisor and

mentor.

I want to thank my family for always being there and supporting me through

everything. To my parents, I can’t express my gratitude enough. I wouldn’t be where I

am today if it wasn’t for you instilling in me a love of learning and pushing me to always

do my best. I appreciate your unfaltering belief in my ability to succeed and your

unconditional support.

And lastly, to my friends who were there to support me, who shared in the

laughter and celebrations, and offered encouragement and support through the tears.

Thanks for being there and helping me through this.

ii
ABSTRACT

In previous studies truncated models were found to underestimate the air’s

upward velocity when compared to wind tunnel velocity studies, which may affect

particle aspiration estimates. This work compared aspiration efficiencies using three torso

geometries: 1) a simplified truncated cylinder; 2) a non-truncated cylinder; and 3) an

anthropometrically realistic humanoid body. The primary aim of this work was to (1)

quantify the errors introduced by using a simplified geometry and (2) determine the

required level of detail to adequately represent a human form in CFD studies of aspiration

efficiency. Fluid simulations used the standard k-epsilon turbulence models, with

freestream velocities at 0.2 and 0.4 m s-1 and breathing velocities at 1.81 and 12.11 m s-1
to represent at-rest and heavy breathing rates, respectively. Laminar particle trajectory

simulations were used to determine the upstream area where particles would be inhaled.

These areas were used to compute aspiration efficiencies for facing the wind. Significant

differences were found in vertical velocity and location of the critical area between the

three models. However, differences in aspiration efficiencies between the three forms

was less than 6% over all particle sizes, indicating that there is little difference in

aspiration efficiency between torso models.

iii
TABLE OF CONTENTS

LIST OF TABLES ............................................................................................................. vi

LIST OF FIGURES ......................................................................................................... viii


CHAPTER

I. LITERATURE REVIEW .................................................................................1

ACGIH Inhalable Particulate Sampling Criterion ............................................1


Flow around Simplified Geometries with No Suction .....................................2
Simplified Geometry Aspiration Studies ..........................................................4
Flow around Complicated Geometries with No Suction ..................................6
Aspiration Studies using Complicated Geometries ........................................10
Limitations of Wind Tunnel Studies .......................................................11
Computational Fluid Dynamics ......................................................................11
CFD Aspiration Studies ...........................................................................12
Objectives .......................................................................................................13

II. EFFECT OF TORSO GEOMETRY ON ASPIRATION EFFICIENCY .......15

Introduction.....................................................................................................15
Methods ..........................................................................................................17
Torso Mode Geometries ..........................................................................17
Mesh Generation and Refinement ...........................................................18
Computational Method ............................................................................19
Convergence Studies ...............................................................................20
Velocity Estimates ...................................................................................22
Particle Release and Tracking .................................................................22
Determination of Critical Area ................................................................24
Data Analysis...........................................................................................25
Results.............................................................................................................26
Convergence Studies ...............................................................................26
Velocity Estimates ...................................................................................27
Velocity Data Above the Neck.........................................................27
Velocity Data Below the Neck .........................................................28
Critical Area ............................................................................................28
Aspiration Efficiency...............................................................................29
No Bounce Simulations ....................................................................30
Bounce Simulations..........................................................................31
Disscusion .......................................................................................................31

III. CONCLUSION...............................................................................................59

APPENDIX A. MESH REFINEMENT DETAILS ...........................................................60

APPENDIX B. RAKE/DATA EXTRACTION POINTS .................................................62

APPENDIX C. RAKE BUILDING AND DATA EXTRACTION PROCEDURE ..........63

APPENDIX D. L2 AND R2 ERROR NORMS.................................................................65

iv
APPENDIX E. VELOCITY REGRESSION PLOTS .......................................................77

APPENDIX F. CRITICAL AREA REGRESSION PLOTS .............................................87

APPENDIX G. ASPIRATION EFFICIENCY REGRESSION PLOTS ...........................89

REFERENCES ..................................................................................................................93

v
LIST OF TABLES

Table 1: Comparison of dimensions between the anatomical model and the 50th
percentile female dimensions....................................................................................37

Table 2: Mesh refinement details .......................................................................................39

Table 3: Mesh refinement ratios ........................................................................................39

Table 4: Boundary conditions. ...........................................................................................40

Table 5: L2 error norms for global convergence tolerance limits 1e-4 to 1e-5. ...................41

Table 6: Between-mesh R2 error norms for 1e-5 GSE tolerance levels..............................43

Table 7: Summary of Differences between torso models…… ..........................................44

Table 8: Between-geometry comparisons ..........................................................................45

Table 9: Linear regression models for velocity data above the neck. ................................46

Table 10: Linear regression models for velocity data below the neck. .............................47

Table 11: Velocity minimum and maximum values ..........................................................48

Table 12: Location of critical area midpoint......................................................................52

Table 13: Aspiration efficiency for all torso geometries. ..................................................53

Table 14: Difference in aspiration efficiencies between models .......................................54

Table 15: Aspiration efficiency for all torso geometries with bounce...............................57

Table 16: Difference in aspiration efficiencies between bounce and no bounce


simulations ................................................................................................................58

vi
LIST OF FIGURES

Figure 1: Illustrations of truncated, full, and anatomical geometries ................................38

Figure 2: Critical area for 7 µm particle in 0.4 m/s freestream velocity with at-rest
breathing inhalation (1.81 m/s).. ...............................................................................49
Figure 3: Critical area for 82 m particle in 0.4 m/s freestream velocity with heavy
breathing inhalation (12.11 m/s). ..............................................................................50

Figure 4: Critical area for 7 m particle for 0.4 m/s freestream velocity and at-rest
breathing (1.81 m/s). Particles released 0.75 m upstream from mouth
opening......................................................................................................................51

Figure 5: Illustration of bounce for 116 µm particles. .......................................................55

Figure 6: Comparison of aspiration efficiency for truncated, full, and anatomical


torso models to experimental data from Kennedy and Hinds (2002). ......................56

vii
1

CHAPTER I

LITERATURE REVIEW

ACGIH Inhalable Particulate Sampling Criterion

Inhalability is defined as the fraction of particles that are capable of being

aspirated into the human respiratory system. Inhalable samplers are designed to match the

entry characteristics of the human head averaged over all orientations to wind direction.

The human head itself can be considered a blunt sampler with complex geometry. As air

moves towards a bluff body with suction, there is both a large scale divergence as the air

flows around the blunt object and a smaller scale convergence as air approaches the inlet.

Using this idea that the human head is a blunt sampler, Ogden and Birkett (1977)

investigated the aspiration efficiency for a human head in wind speeds ranging from 0.75

m/s to 2.75 m/s. Other researchers have also investigated the effects of airflow around

similar blunt bodies with aspiration (Ingham, 1981; Sreenath, Ramachandran and

Vincent, 1997).

The American Conference of Industrial Hygienists (ACGIH) Inhalable Particulate

Matter (IPM) sampling criterion is based on the early experimental work conducted by

Vincent and Armbruster (1981), Ogden and Birkett (1977), Vincent and Mark (1982),

and Armbruster and Breuer (1982), and summarized by Soderholm (1989). In the early

1990s the Internal Organization for Standardization (ISO), European Committee for

Standardization (CEN) and ACGIH unified the definition of inhalable aerosols and the

inhalable particulate sampling criterion (IPM) was the result. The ACGIH inhalable

particulate matter (IPM) sampling criterion defines the desired sampling performance of

a sampler for particles up to 100 m and is based on the aspiration efficiency of a human

head. The ACGIH inhalable fraction and the inhalable particulate mass criterion IPM is:

IPM 0.5 1 exp 0.06d


da = aerodynamic diameter in µm, < 100 m
2

The original human aspiration studies agreed with thin walled sampler studies, or

isokinetic samplers, which showed that aspiration efficiency is dependent on freestream

velocity, breathing rate and orientation to the wind. However, the experimental studies

used to develop the IPM criterion used wind velocities that were much higher than those

found in the workplace. In a study conducted by Baldwin and Maynard (1998) it was

found that 85% of the time indoor wind speeds were less than 0.3 m/s. Since the IPM

sampling criterion was defined in higher freestream velocities, the question of whether

the current sampling criterion accurately mimics human aspiration efficiencies in low
velocity conditions was raised. At very low velocities, large particles are more influenced

by gravitational effects than by velocity. Aspiration efficiency, as described by the IPM

sampling criterion, may overestimate actual aspiration of large particles at lower

velocities. If true, the IPM sampling criterion may need to be revised for lower velocities

typical of indoor environments.

Since most of the human aspiration studies rely on tests with torso simplifications,

either truncation in wind tunnels or geometrical simplifications in numerical studies, the

effects of flow around simplified geometries is worth discussion. The research

investigating airflow and aspiration efficiency starting from very simplified to

increasingly complex geometries is presented here.

Flow Around Simplified Geometries with No Suction

This section examines the research investigating airflow around simple

geometries, typically extremely simplified human models that lack suction. There has

been a large amount of research conducted investigating the thermal effects of bodies on

pollution transport and heat transfer. Murakami et al. (1999) investigated airflow (0.25

m/s) around a standing smoothed human body to simulate conductive heat transfer and

convection flow in a ventilated room. The model was placed in stagnant flow, downward

uniform flow and upward uniform flow. For the stagnant flow, warm rising air was found
3

to be generated around the model, while the rest of the flow field was stagnant and a very

weak downward flow was generated around the legs. The same model was used by

Murakami et al. (2000) to investigate airflow, radiation and moisture transport from a

human body. The results were compared to flow visualization patterns near the shoulders

and top of the head from an experimental mannequin, a real human body, and previous

experiments, and agreement was found between simulations and experiments. Murakami

et al. (1999) investigated the effect of wind on a human body for stagnant flow, low wind

velocity (0.25 m/s) and high wind flow (2.5 m/s). The same simplified, smoothed models
used in the previous studies were also used in this study. For the stagnant flow, rising air

was found to accelerate as it passes from the neck to the head region. The rising thermal

plume above the body was found to be consistent with results from an experimental

thermal mannequin and a human body. For the low uniform flow (0.25 m/s) the flow

field below the chest showed impinging air flowing downwards, then flowing past the

body.

Brohus and Nielsen (1996) examined airflow around three heated but simplified

torsos: a solid rectangular torso, a rectangular torso with legs and a rectangular torso with

a box head; and none of these models had facial features, arms or breathing simulation.

The effect of the inclusion of legs in the model resulted in a decrease in upward vertical

velocity in the region of the head. The researchers hypothesized that this was due to air

slipping between the legs, instead of being forced upwards toward the head region.
Brohus and Nielsen (1996) and Hayashi et al. (2002) also examined the heated

torso while investigating contaminant transport. Brohus and Nielsen (1996) modeled

three torso geometries: a flat rectangular model, a rectangular torso with legs, and a

rectangular torso with legs and a box head, facing 180° to the wind source. They found

that the model lacking legs caused air to be entrained and directed toward the person, but

the model with a head and legs caused the air to be moved away from the person.

Hayashi et al. (2002) used a smoothed simplified heated torso in three postures: standing,
4

sitting and sleeping. Inhalation was estimated by calculating the index of effectiveness of

contamination ventilation, which was derived from the scale of ventilation efficiency 5

(SVE5). They examined the region where airflow would have been inhaled. However,

those studies lacked convergence data and failed to examine mesh densities. Bjorn and

Nielsen (2002) examined the influence of human exhalation on flow fields using life-

sized, breathing, thermal mannequins for both experimental and numerical studies. Their

experiments looked at simple heated rectangular boxes, stacked rectangular boxes to

represent head, torso and legs, as well as simplified and realistic mannequins. They also
reported that the presence of legs in the model decreased the concentrations when the

contamination source is near to the ground.

Most of these studies were primarily focused on examining the effects of heating

a body on airflow and pollution contamination and as such, none of these models include

inhalation. Although these studies provide evidence that torso geometry effects air flow,

it is not evident whether aspiration efficiencies would be affected, since inhalation was

ignored.

Simplified Geometry Aspiration Studies

Simplified geometries with suction to represent breathing have been used to

investigate aspiration. Many previous numerical investigations of particle aspiration have

used simplified geometries for the human form and facial features, including a simplified

cylinder (Dunnett and Ingham, 1986; Chung and Dunn-Rankin, 1992; Dunnett and

Ingham, 1987; Ingham and Hildyard, 1991). Ingham (1981) conducted a two-dimensional

analysis on a horizontal cylindrical sampling head for two conditions: particles could

bounce on the surface of the sampler or adhere to the surface of the sampler. This study

derived aspiration coefficients for the semi-empirical equations obtained by Vincent et.

al. (1979), Vincent and Mark (1980), and Vincent, Hutson and Mark (1981). The

aspiration coefficient was found to be close to unity for small Stokes numbers, and a
5

function of the bluffness when Stokes number is large. For the condition where particle

adhere, the maximum value of the aspiration coefficient was found to be at Stokes

number equal to one, not at very large or very small Stokes numbers. This corresponds

with experimental results by Ogden and Birkett (1977). Dunnett and Ingham (1987)

examined the effect of particle Reynolds number on aspiration of particles into a

cylindrical sampler. Erdal and Esmen (1995) used a cylinder with a rounded top to

examine coarse particle aspiration. They investigated a cylinder with a round top to

represent a human head and examined aspiration for particles up to 200 µm at three
breathing rates and ambient windspeeds ranging from 0.5-9 m/s. They reported that

numerical calculations agreed with the experimental results showing a leveling off of the

aspiration efficiency. They reported no evidence of a “cut-off” or aspiration efficiencies

going to zero for particles up to 200 µm, however aspiration efficiencies were in the

range of 5-10%. They also reported that at high ambient windspeeds there was an

increase in aspiration efficiency by increasing particle size, which also agreed well with

experimental results. Dunnett (1997) characterized the flow field around a three-

dimensional spherical bluff body with aspiration. Dunnett (1999) further investigated

aspiration for a spherical bluff body with suction and extended previously derived

equations to small angles of orientation.

Several researchers have investigated the effects of airflow and aspiration around

blunt bodies (Ingham, 1981; Sreenath, Ramachandran and Vincent, 1997). However,
many of the numerical simulation studies have been conducted at wind velocities greater

than the 0.3 m/s found in indoor workplaces. Many of these studies also have no

validation studies to compare to, with some notable exceptions. Dunnett and Ingham

(1988) investigated aspiration in both experimental mannequin studies and numerical

simulation experiments. Their results found that facial features may affect particle

transport in low suction and high freesteam velocities. Sreenath et al. (1997) examined

airflow around simple two- and three-dimensional geometries that contained single
6

sampling orifices in wind tunnels. Using smoke for flow visualization they examined

orientations ranging from 0° to 180° with respect to the wind and used both cylindrical

bodies as well as spherical bodies. The experimental results compared well with the flow

models developed by Dunnett and Ingham (1988) and Ingham and Hildyard (1991).

Anthony, Flynn and Eisner (2005) compared a simple elliptical cylindrical

mannequin to a 2/3 scale anatomical mannequin in a wind tunnel study to determine if

the use of simplistic geometries in computational fluid dynamics (CFD) studies was

appropriate. They reported significant velocity differences in the mouth region for the
anatomical mannequin compared to the elliptical model, indicating limited applicability

of using simplified geometries as a surrogate for the human head. Previous numerical

investigations of aspiration using simplified geometries (Ingham, 1981, Dunnett and

Ingham, 1986, 1987, 1988, Ingham and Hildyard, 1991, Chung and Dunn-Rankin, 1992,

and Erdal and Esmen, 1995) may have limited applicability to understanding human

aspiration of particles.

These numerical studies were some of the first studies to examine and

characterize flow around simplified objects intending to represent human features.

However, more recent work conducted by Anthony, Flynn and Eisner (2005), indicated

that simplified geometries may not be appropriate substitutes for realistic facial features.

Flow Around Complicated Geometries with No Suction

The effect of torso geometry shapes has been investigated, however, not

extensively and not for the effects on human aspiration efficiencies. Vincent (1989)

showed that the presence of a body influences the airflow around the complex geometry

of the nose and mouth. Studies investigating sampler placement have found regions of

stagnation and lower velocities located in the center of the front of a mannequin (Rodes,

Kamens, and Wiener, 1995). Smith and Bird (2002) examined whether sampling

efficiency of personal samplers was influenced by the presence of a mannequin. The


7

study used a non-breathing full sized, but truncated, tailor’s mannequin at low wind

speeds (0.5 and 2 m/s). Velocity measurements were taken in the chest region as well as

in the face area. They reported that at the center of the mannequin wind speeds

decelerated and approached a low average velocity with a slight upwards component. At

locations slightly to the right and left (0.1 m) of the center, there was less of a decrease in

velocity. In the region of the face, airflow was found to decelerate, but not as much as in

the chest region. Those studies showed that the presence of a mannequin was necessary

when evaluating sampler performance.


Kennedy, Tatyan and Hinds (2001) investigated the differences in inhalable

sampler performance between a simplified and full-sized mannequin in an experimental

study at low velocity conditions: 0.4, 1.0 and 1.6 m/s. Although the sampler performance

agreed well with the ACGIH IPM sampling criterion, the simplified mannequin resulted

in an aspiration efficiency that was approximately 35% that of the full mannequin’s

aspiration efficiency at low wind velocity (0.4 m/s). The researchers hypothesized that

the deflected air is greater for the full sized mannequin due to its larger surface area,

which accounted for the differences. When air approaches a full sized mannequin, it is

deflected upwards as well as laterally to flow around the mannequin. Since the deflected

air is approaching the inlet at an angle, the aspiration efficiencies are reduced. The

truncated mannequin deflects less air due to a smaller surface area and smaller vertical

height and causes an increase in aspiration efficiencies.


Schmees, Wu and Vincent (2008) investigated the airflow around a life-sized but

truncated heated breathing mannequin at ultralow (0.1, 0.24, and 0.42 m/s) windspeeds.

Smoke was used to visualize the airflow around the mannequin that both inhaled and

exhaled. They reported that expiration though the mouth has a significant effect on the

airflow patterns at 0.1 m/s freestream velocity for both breathing rates, and at 0.24 m/s at

the high breathing rate. They also found that body temperature had a minimal effect on

the airflow, contrary to the results found by Murakami et al. (1999).


8

There have been several studies examining thermal mannequins (Bjorn and

Nielsen, 2002; Brohus and Nielsen, 1996; Xing, Hatton and Awbi, 2001; Chang and

Gonzalez, 1993; Myers, Hosni and Jones, 1998; Lewis et al., 1969, etc) in the literature

and the complexity of these mannequins ranges from simplified boxes or cylinders

(Murakami et al., 1999; Bjorn and Nielsen, 2002) to more realistic human like

mannequins (Sorensen and Voigt, 2003). Hyun and Kleinstreuer (2001) created a

breathing mannequin with divided legs, cylindrical neck and spherical head and

calculated the flow field. They compared the inhaled non-dimensional dose concentration
to concentrations measured by Brohus’ (1997) simple rectangular box model and found

that the concentrations agreed well, which would indicate that increasing the complexity

of torso geometry did not influence inhalability.

Sorensen and Voigt (2003) also examined the convective flow using

computational fluid dynamics (CFD), this time with a seated mannequin generated from

imaging of an experimental mannequin, and the results were compared to measurements

made with particle image velocimetry and other experimenter’s velocity data. Their

results indicated that the local heat transfer properties are significantly affected by the

presence of arms, armpits, and legs. These CFD results correlated well with published

whole-body data, with variations between the CFD and published velocities being

attributed to differing postures.

Zhu, Kato, Murakami, and Hayashi (2005), investigated the region of inhalation
under a steady inhalation rate in both experimental and CFD studies. The researchers

investigated the inhalation region, which was defined as the ratio of air at any point

indoors that the occupant inhales. A simplified torso geometry and a more complex CFD

model with feet, arms, jaw, and breast were generated and the fluid flow fields solved.

The inhalation region was shown to be different in concentration between the two

models. The results from this study indicate that simplifying the human torso geometry

can affect the estimates of concentration near an inhaling mouth. When compared to
9

experimental data, the more complex torso geometry estimates correlated better with the

experimental results than the simplified torso model.

A validation study for the computer simulated person used in Nielsen, Murakami,

Kato, Topp, and Yang was performed by Sideroff and Dang (2005). Their results showed

that CFD could be used to accurately estimate personal microclimate environment as long

as (1) the number of grid cells was sufficiently high (approximately a few million cells)

to achieve grid convergence, (2) that the boundary conditions were accurately quantified

to provide uniform velocity flow at the inlet, and (3) that the correct turbulence model
was used to predict flow. The standard k-ε turbulence model resulted in velocity

estimates that did not compare well to test data, so two other Reynolds Averaged Navier

Stokes (RANS) models were evaluated, and the v2-f model provided the best estimates of

flow.

Topp, Hesselholt, Trier, and Nielsen (2003) investigated the influence of torso

complexity in their experimental study using seated mannequins. They reported that the

complexity of the mannequin geometry had no effect on personal exposures, but did

influence the convective flow and concentration distribution at some distance away from

the mannequin. The presence of legs was found to be important. The velocity was found

to be significantly lower at the center of the torso when the air was forced to flow around

instead of between the legs. However at the mouth, the velocity profiles were found to be

similar, indicated that the presence of legs may not affect aspiration efficiencies greatly.

These thermal studies identified differences in vertical velocity depending on

torso complexity and identified the presence of legs important. They also found that the

velocity differences were greater in the torso region compared to the facial region;

however, these studies lacked inhalation in their mannequins which may be an important

factor.
10

Aspiration Studies using Complicated Geometries

Studies looking at inhalability and the aspiration efficiency of the human head

were conducted as early as 1977, when Ogden and Birkett recognized the need to define

inhalability and create a consistent sampling criterion that matched the biological effects

of exposure. Ogden and Birkett (1977) examined human head aspiration efficiencies and

characterized the intake efficiency measurements for the human head. The head was

found to consistently under-sample compared to a 10-mm sharp-edged probe (isokinetic

sampler) and that under-sampling increased with particle size. Despite the limitations of

this study, including a small wind tunnel with high blockage ratio, a lack of torso on the

test mannequin, and maximum particle size of 30 µm, these studies formed the basis of

human aspiration studies. Vincent and Armbruster (1981) compared data from the Essen

and Edinburgh experiments (Armbruster and Breuer, 1981) with data from Ogden and

Birkett (1977). In the Essen experiments, an isolated head was used to examine

inhalablity for particles up to 60 m averaged over all orientations. In the Edinburgh

experiments, a model head mounted on a full torso and used to examine inhalability of

particles up to 100 m over all orientations at wind speeds up to 4 m/s. Both studies

found that inhalability decreased from unity as particle aerodynamic diameter increased.

These broad trends in inhalability were not only seen in the Essen and Edinburgh

experiments, but in Armbruster and Breuer (1981) and Ogden and Birkett (1977) as well.

For particles larger than 40 m, aspiration efficiency was observed to remain around

50%. Those experiments, however, did not examine inhalablity for particles up to the 100

µm, as defined by the current criterion, nor were torsos included in all of those

experiments.

The results are conflicting for many of the experimental studies in low velocities.

Aitken et al. (1999) reported 50% aspiration efficiency for particles as large as 150 µm in

a low air environment, which was defined as effectively no net air movement and peak

velocities less than 0.1 m/s, whereas Hsu and Swift (1999) reported no aspiration for 150
11

µm particles in calm air. Hsu and Swift found that inhalability was approximately 20% at

particle diameters of 65 µm and for a particle diameter of 135 µm inhalablity was

approximately 2%. Kennedy and Hinds (2002) reported orientation averaged results that

showed for particles smaller than 35 µm, average inhalability was significantly lower

than the IPM sampling criterion, and that for particles larger than 70 µm inhalability

leveled off at approximately 30%. Each of these studies used an experimental full-torso

mannequin to examine nose and mouth breathing aspiration efficiencies, but reported

having particle suspension problems.


The results for aspiration studies using more realistic geometries are conflicting.

This may be due in part to difficulties associated with wind tunnel studies.

Limitations of Wind Tunnel Studies

Experimental studies are faced with several difficulties when examining

aspiration of larger particles at low velocity conditions. First, it is difficult to maintain a

uniform flow in a wind tunnel large enough for a full-scale mannequin. Secondly,

generating and maintaining uniform particle concentrations for larger particles sizes is

difficult. With recent advances in computer technology, the use of CFD has become an

effective method used to examine aspiration while overcoming the difficulties faced in

experimental wind tunnel studies. Within the past couple of years several more complex

CFD models have been developed to investigate the factors influencing aspiration
efficiency.

Computational Fluid Dynamics

Several CFD studies have been conducted to examine aspiration in human

models. Before discussion of those studies are possible, however, an understanding of

aspiration theory in numerical studies is necessary.

For uniform particle concentrations upstream of a suction source, aspiration

efficiency can be defined as the concentration of inhaled particles over the concentration
12

in the freestream. Anthony and Flynn (2006) derived the equation for aspiration

efficiency used in numerical studies as the following:

(1)

Where Ac = critical area

Uc = freestream critical velocity

Am = area of the mouth

Um = inhalation velocity

The critical area is defined as the upstream location within which all particles will be

inhaled. This definition relies on the assumption that there is a uniform distribution of

particles upstream. The freestream critical velocity, Uc, is the velocity in the freestream

where particles are released.

CFD Aspiration Studies

Models with more realistic facial features, but simplified cylindrical torsos

truncated at waist height, have been developed in recent years.

Anthony and Flynn (2006B) created a full scale CFD human form with detailed

facial features to examine aspiration efficiency. Aspiration efficiencies were compared to

facing-the-wind experimental data from Kennedy and Hinds (2002). They reported that

aspiration efficiency for particles smaller than 52 µm compared well with the

experimental results from Kennedy and Hinds (2002), but for particles larger than 52 µm

the CFD model underestimated aspiration efficiencies. Anthony and Flynn hypothesized

that the truncated, simplified torso in the CFD model may have contributed to velocity

differences which may have affected aspiration efficiency estimates.

More recently, King Se, Inthavong and Tu (2010) investigated nose and mouth

inhalability of a CFD model with a realistic human head. In that study two inhalation

rates were examined, a low rate (15 L/m) and a high rate (40 L/m) at a low freestream
13

velocity of 0.2 m/s using a full sized torso model, and the results were compared to

Anthony and Flynn (2006B). The full sized CFD model caused vertical velocity to start

moving downward towards the mouth farther upstream than for the realistic facial feature

model, indicating differences in velocity in the facial region for a full sized model

compared to the truncated model in Anthony and Flynn (2006). Bluff body effects in the

near breathing region were found to influence particle trajectories.

Objectives

Anthony and Flynn’s (2006A , 2006B) and Anthony (2010) CFD studies provided

highly detailed, realistic human facial features needed to examine large particle aspiration

efficiency. However, those studies all made simplifications to the human torso. Studies of

airflow around heated and unheated mannequins have identified flow field changes when

the separation of the legs is ignored, where the vertical flow near the torso may increase

relative to geometries with legs. This effect could become especially important in the low

velocities representative of an indoor workplace. Whether these differences in velocity

impact how particles aspirate into an inhaling mouth has not been assessed. Numerical

simulations examining the effect of torso geometry on thermal transport and personal

microclimate environments have not included particle transport. Many of those studies do

not include inhalation in the mannequin; those that have included inhalation used

simplified torso models or lack convergence and mesh density information to evaluate the

quality of the simulations. These studies lead to the conclusion that the effects of torso

geometry simplifications on aspiration efficiency and air flow are not well understood

and need to be investigated further to better understand their effects on aspiration

efficiency.

This research explicitly examines the effects of using simplified torso geometry

on large particle aspiration efficiency. Three torso geometries are evaluated to examine

the effect of using both a simple geometry (elliptical torso matching the 95th percentile
14

female height) and using a truncated simple geometry compared to the baseline of a

realistic model with legs, arms, and torso scaled to the same 95th percentile female

dimensions. The study examines whether the torso size and shape affects (1) the velocity

field near the mouth, (2) the position of upstream critical areas that result in aspirated

particles, and (3) the estimates of aspiration efficiency. The first objective confirms the

findings of other modelers, this time with suction: namely does the velocity change based

on the torso shape. The third object examines whether the impact of velocity differences

affect aspiration, where uniform concentrations of particles are modeled and the position

of the upstream area is irrelevant in these computations. The second objective, while

necessary to generate the data to evaluate the third question, also allows for the

examination of whether simplified torsos are sufficient to model contaminant transport

from a source. This is useful when examining the appropriateness of using torso

simplifications in other exposure assessment applications.


15

CHAPTER II

EFFECT OF TORSO GEOMETRY ON ASPIRATION EFFICIENCY

Introduction

There have been many previous studies investigating aspiration efficiency, both

experimentally and numerically. Most of these studies have used torso simplifications,

although the degree of simplification can vary dramatically from study to study. The

original aspiration studies by Ogden and Birkett (1977) simply used an isolated head.

Studied by Aitken et al. (1999) used a torso truncated at the waist. Many other

experimental studies used full sized torsos truncated at hip or waist height. Although

commonly used, the appropriateness of using simplified torso geometries as surrogates

for human bodies has not been investigated thoroughly. Some studies have identified

differences in airflow or aspiration when using simplified models. Anthony, Flynn and

Eisner (2005) investigated this issue to some extent. They compared an anatomical

mannequin to simplified elliptical torso in a wind tunnel study and identified significant

velocity differences in the facial region. Anthony and Flynn (2006B) created a full-scale

CFD human geometry with detailed facial features to examine aspiration efficiency.

Aspiration efficiencies were compared to facing the wind experimental data from

Kennedy and Hinds (2002). They reported that particles smaller than 52 µm compared

well with the experimental results from Kennedy and Hinds (2002), but for particles

larger than 52 µm the CFD model underestimates aspiration efficiencies compared to

published results. Anthony and Flynn hypothesized that the truncated, simplified torso in

the CFD model may have contributed to velocity differences which may have affected

aspiration efficiency estimates.

Another study investigated the effect of torso geometry on sampler aspiration and

found that the use of a simplified torso (fashioned from a wastebasket) resulted in an

aspiration efficiencies 35% that of samplers placed on a torso-truncated human


16

mannequin (Kennedy et al. 2001). Sampler studies conducted by Smith and Bird (2002)

identified a decrease in vertical velocity in the chest region, with a smaller decrease in the

facial region.

Numerous thermal studies investigating airflow and pollution transport around

human bodies have been conducted. Although these studies lack inhalation, they have

identified that torso complexity can influence airflow around the body and in the facial

region. These studies have also shown the presence of legs causes a decrease in velocity

at the center of the torso (Topp, Hesselholt, Trier, Nielsen, 2003).


Although realistic human facial features have been used to examine large particle

aspiration efficiency, simplifications were made to the human torso (Anthony and Flynn,

2005, 2006). Flow field changes have been identified for both heated and unheated

mannequins when the separation of legs is ignored, where vertical flow may increase

relative to geometries with legs. When looking at indoor workplaces, which have lower

velocities, this could become an important effect. However, the impact of the differences

in velocity on particle aspiration has not been assessed. Studies investigating personal

climate and thermal transport have not included particle transport or suction. Those

studies that have included suction lacked convergence data and mesh density information

to evaluate the quality of the simulations. This leads to the conclusion that the effects of

torso geometry are not well understood and need to be investigated further to better

understand their effects on aspiration efficiency.


This research explicitly examines the effects of using simplified torso geometry

on large particle aspiration efficiency. Three torso geometries are evaluated to examine

the effect of using both a simple geometry (elliptical torso matching the 95th percentile

female height) and using a truncated simple geometry compared to the baseline of a

realistic model with legs, arms, and torso scaled to the same 95th percentile female

dimensions. The study examines whether the torso size and shape affects (1) the velocity

field near the mouth, (2) the position of upstream critical areas that result in aspirated
17

particles, and (3) the estimates of aspiration efficiency. Objective 1 confirms the findings

of other modelers, this time with suction: namely does the upstream velocity change

based on the torso shape. Objective 3 examines whether the impact of velocity

differences affect aspiration, where uniform concentrations of particles are modeled and

the position of the upstream area is irrelevant in these computations. The second

objective, while necessary to generate the data to evaluate the third question, also allows

for the examination of whether simplified torsos are sufficient to model contaminant

transport from a source. This is useful when examining the appropriateness of using torso
simplifications in other exposure assessment applications.

Methods

Torso Model Geometries

As shown in Figure 1, three torso geometries were generated using Gambit, a

CFD modeling software (Ansys, Inc, Lebanon, NH, USA). Each geometry used the same

head with small nose and large lips features, described fully in Anthony 2010. The first

geometry used an elliptical torso 0.2775 m tall, 0.2325 m wide and 0.1725 m depth,

representing truncation at hip height, which was identical to that used in previous work

(Anthony et al., 2005). The second geometry extended the elliptical torso to a height of

1.07 m, for a total height of 1.56 m, approximating the 50th percentile female height. The

final geometry incorporated a realistic anatomical human torso that matched the height of

the second geometry and approximated the 50th percentile female dimensions in other

dimensions (Table 1). The hands and feet were simplified by truncation, assuming their

influence on particle simulations would be negligible. These geometries were fairly

complex and realistic geometry compared to previous inhaling humanoid CFD studies.

These geometries are referred as “Truncated,” “Full” and “Anatomical,” respectively,

throughout this text.


18

This work focused on facing-the-wind orientation. Lateral symmetry was

assumed, to allow only half the domain to be simulated and reduce computational time.

The dimensions of the computational domain, constructed to represent a wind

tunnel, were chosen so that the torso geometries were located far enough away from the

walls that the fluid field was not influenced by the location of the walls. The flow

upstream of the torso geometries would be fully developed, and no acceleration through

the outflow of the domain was a reasonable assumption. Hence, the length (3.6576 m)

and width (1.142 m) of the computational domain and the position of the mouth opening
(origin) were the same for all three geometries. The height of the domain was 2.3 m for

the full and anatomical geometries, but was reduced to 1.23 m for the truncated domain,

with the plane of truncation coinciding with the bottom of the domain.

Mesh Generation and Refinement

Ansys 12.0 (Ansys, Inc, Lebanon, NH, USA) was used to mesh the computational

domain. A paved meshing scheme was used because of the complex geometries involved,

which used triangular surface and tetrahedral volumetric elements. To evaluate the

quality of the simulations’ solutions, three sequential mesh refinements were generated

for each geometry by increasing node counts on each edge by a factor of 1.2. The mesh

densities on the facial features were all assigned similar values for each geometry at each

given mesh condition. For the heavy breathing conditions, coarse mesh simulations did
not converge and a mesh density between the moderate and fine meshes was generated

for additional simulations. The mesh refinement ratio was calculated using equation 2.


(2)

Tables 2 and 3 summarize the mesh refinement details. Refinement ratios between

the coarse, moderate and fine meshes ranged from 1.208 – 1.287. Between the moderate,

mod-fine, and fine meshes, the refinement ratios ranged from 1.097-1.107. Least refined
19

meshes ranged from 300,000 to 700,000 nodes, and most refined meshes ranged from 1.1

to 2.3 million nodes, depending on geometry. Additional details of the mesh refinement

and the number of elements and nodes for each torso geometry are shown in Appendix A.

Computational Method

Fluent 12.0 (Ansys Inc, Lebanon, NH) was used to solve first the fluid flow field,

then particle simulations. Fluent solves the equations of fluid flow for each degree of

freedom at each of the mesh nodes. For this work, six degrees of freedom were

investigated: three velocity components (streamwise=U, lateral=V, vertical=W),


turbulence kinetic energy (KE), dissipation of turbulence kinetic energy (E), and pressure

(P).

The airflow was modeled using steady-state, incompressible, turbulent, Navier-

Stokes equations:

∙u 0 (3)

u∙ u ∙ μ μ u u ρ (4)

The standard κ-ε turbulence model was used:

ρ u∙ ∙ μ μΦ ρε (5)

ρ u∙ ∙ μ c μΦ c Φ c ρ (6)

where: k is the turbulence kinetic energy

ε is the turbulent dissipation

Φ is the viscous dissipation function

The standard constants were used:

cµ = 0.09 σk = 1.00 σε = 1.30

c1 = 1.44 c2 = 1.92
20

Standard wall functions were applied and full buoyancy effects were modeled.

Gravity was set to act downward at 9.81 m/s2. Indoor room air temperature was simulated

(20°C) with the corresponding air density (1.205 kg/m3) and viscosity (1.83692x10-5

kg/m-s). Boundary conditions were set on the surfaces in the computational domain and

are summarized in Table 4.

Four velocity conditions were investigated: two breathing velocity and two

freestream velocities. A uniform velocity of 0.2 m/s or 0.4 m/s was assigned for the

domain entrance. The two breathing velocities were chosen to represent an at-rest (1.81

m/s) and a heavy (12.11 m/s) breathing condition, equivalent to the mean inhalation

velocities for 7.5 and 50.3 L/m cyclical breathing rates. These rates represent the two

extremes likely to be experienced in the workplace.

The turbulent intensity ratio was set at 8%, typical of wind tunnel studies, giving

turbulent kinetic energies of 0.000384 for 0.2 m/s and 0.001536 for 0.4 m/s. The ratio of

eddy viscosity to laminar viscosity was set to 10. The bottom of the computational

domain (the “floor”) was set as symmetry for the truncated model, which allows air to

move along this surface but not through it. For the full and anatomical models the bottom

of the domain was set as a wall, meaning no flow was allowed at the plane’s surface.

A pressure based decoupled segregated solver was used, where the momentum

equation was solved first with pressure solved to enforce continuity, and then the

continuity equation was used to solve for the other degrees of freedom. The SIMPLE

algorithm with second order upwinding was used. Solutions were obtained when the

global solution error (GSE) was below pre-determined levels for each of the six degrees

of freedom. The three levels of GSE were 10-3, 10-4, and 10-5. Typically, the continuity

(pressure) term was the last to reach the GSE tolerance level. Not all of the velocity

conditions converged to the 10-5 solution level, but the simulation was stopped and the

solution evaluated when the last of the degrees of freedom was no longer decreasing.

These simulations were stopped after approximately 2000-3000 iterations.


21

Convergence Studies

The solution for each of the six degrees of freedom were extracted for 1710 points

ranging from 0.015 to 0.75 m upstream of the mouth (X), 0.0 to 0.75 m laterally from the

mouth center (Y) , and 0.3 m below to 0.6 m above the mouth (Z) . The L2 error norm

and the R2 error norm were computed using Equations 9 and 10 to assess the adequacy of

the simulation solutions. The L2 error norm (Equation 7) was assessed within each mesh-

density for each geometry, using solutions between each successively lower GSE

tolerance. The resulting error norm indicates how much change occurred in a value of a

given degree of freedom between solutions; a target of 5% was established a-priori.

Particle simulations could be performed on the solutions with the between GSE L2 error
norms less than 5%.
.

. (7)

where

L2 = the error norm

f = the degree of freedom of interest

lower = indicates the set of data from the solution with a lower

tolerance level

higher = indicates the set of data from the solution with the higher

tolerance level

.

∑ . (8)

where

f = the degree of freedom of interest


subscript indicates the mesh density where fine has the most

number of nodes and coarse has the least


22

The R2 error norm (Equation 8) was used to evaluate whether the solution was

independent of the mesh densities that were used to solve the fluid flow equations (Stern

et al. 2001). When R2 is between 0 and 1, the data from that location was monotonically

converging. The least refined mesh that demonstrated mesh independence (0<R2<1) was

used for subsequent particle simulations.

Velocity Estimates

Vertical velocity was assessed in the head/facial region as well as for areas below

the neck. Vertical velocity data were extracted from the same 1710 data points used for

convergence studies. Linear regression models were developed to examine the

relationship of the velocity estimates between the anatomical geometry, set as our

independent variable, and the other simplified geometries. In addition the relationship

between the two cylindrical geometries was examined.

Particle Release and Tracking

The Euler-Lagrange approach was used to solve for particle motion. The fluid

solutions were first solved, then particle motion was tracked through the calculated fluid

flow. It was not anticipated that the particles would influence the fluid field making it

acceptable to uncouple the particles from the fluid flow. Laminar particle trajectories

were examined: thus, the estimates of aspiration efficiency reflect mean values and

cannot incorporate uncertainty due to turbulent particle behavior. The particle momentum

is described by equation 9:

f (9)

where:

ui = fluid velocity
uip = particle velocity
23

ρp = particle density

ρP = fluid density

fiP = forces acting on the fluid

τ = particle relaxation time, given by


τ (10)

where: µ = fluid viscosity

dp = particle diameter

Rep = particle Reynolds number = dp |ui-uiP|/µ


CD = drag coefficient (a function of Re)

Seven particle sizes were released with an initial velocity of the freestream and

tracked as follows: 7, 22, 52, 68, 82, 100, and 116 m. These particle sizes were chosen

to match the data published in Kennedy and Hinds (2002) and Anthony and Flynn

(2006B).

Non-evaporating, unit density particles were released, which allowed for

reporting in aerodynamic diameters. The release points were located more than four head

diameters away from the torso models to ensure that bluff-body effects were negligible

(Chung and Dunn-Rankin, 1997). As such, the particles were released 0.75 m upstream

of the mouth opening for particles with diameters less than 82 µm and 0.4 m upstream for

particles 82 µm and larger. The release locations for large particles was moved closer to

the mouth opening, but remained overhead, to accommodate the gravitational settling of

the larger particles. Velocities at these locations were checked to ensure no bluff body

effects were seen. To meet the uniform particle distribution assumption, particles were

released at the velocity they would be at the freestream. This was accomplished by

assigning a horizontal velocity equal to the velocity in the wind tunnel at the release
24

position and a vertical velocity equal to the combination of the initial velocity of the

freestream at the release location combined with the terminal settling velocity of the

particle being evaluated.

Particle simulations used a 50-m length scale, which determines the distance the

particle will travel before the particle trajectory is updated. Additional studies will need

to examine the sensitivity of the length scale. To control the error when calculating the

pathlines, the tolerance was set at 1x10-5. The maximum refinement or the largest number

of step size refinements in one single integration step was set to 20 step sizes. The

spherical drag law, as shown in equation 10, was used.

(10)

Where a1, a2 and a3 are constants (Morsi and Alexander, 1972)


Secondary aspiration occurs when a particle strikes an external surface with

enough energy to bounce off and then be aspirated into a suction opening. Bounce and no

bounce conditions were simulated. For the no bounce condition, any particle that

contacted a surface was assumed to deposit there. For the bounce condition, any particle

that struck the torso was allowed to bounce with 100% reflection and continue being

tracked.

Determination of Critical Area

Particle simulations were performed to determine the location of the positions that

resulted in particles that were aspirated, defined as the critical area. The critical area was

determined by a series of particle releases in steps across the mouth. Over a series of

lateral (Y) positions, 100 particles were released across 1-cm vertical (Z) distances at the

given upstream location (X=-0.75, -0.4). The uppermost position of particles going into

the mouth along these lines were recorded as the top of the critical area, and the process

was repeated at sequential lateral distances 0.5 mm away from the previous release

position, until no particles at any height were aspirated. This identified the positions
25

defining the top of the critical area. The process was then repeated along the bottom of

the critical area to determine the position associated with the bottommost position of

particles terminating in the mouth. The critical area was determined using equation 11:

A ∑ Y Y Z , Z , (11)
where

N= the total number of Y positions where particle inhalation was

observed

Yi = the lateral coordinate of the ith position

ZU,i = the upper vertical particle release location associated with


the Yi where a particle is inhaled

ZL,i = the lower vertical particle release location associated with

the Yi where a particle is inhaled

This process was conducted for each particle size (7), velocity condition (4),

bounce condition (2) and torso geometry (3) for a total of 168 critical area

determinations.

Data Analysis
Between-geometry comparisons of velocity, central height of critical areas, and

aspiration efficiencies for both bounce and no-bounce conditions were assessed. The

range and mean of differences were computed by matching particle size, velocity

conditions, and, for velocity, solution positions. In addition, differences were assessed

using one-tailed (critical areas) or two-tailed (velocity, aspiration efficiency) t-tests.

Linear regression techniques were applied to assess the magnitude and practical

significance of these differences.


26

Results

Convergence Studies

The L2 error norms computed from solutions from sequentially reduced global

convergence tolerance limits are given in Table 5 for 6 degrees of freedom: x-velocity

(U), y-velocity (V), z-velocity (W), kinetic energy (K), epsilon (ε), and static pressure

(SP). For nearly all the freestream and breathing velocity conditions between the 1x10-4

and 1x10-5 solutions, the solution changed less than 5%, meeting the a-priori criterion.

The exception was for vertical velocity (W), turbulence kinetic energy (KE) and

dissipation of turbulence kinetic energy (ε) at the 0.2 m/s, at-rest breathing condition for

the full model. In each case, particle simulations were performed using the most refined

GSE solution (1x10-5). The L2 and R2 error norms are shown in detail in Appendix C.
The between-mesh R2 error norms for the 1e-5 GSE solutions were computed

(Table 5). Bold values indicate an acceptable value, while values in italics indicate non-

monotonically converging solutions. For at-rest breathing, the anatomical and truncated

torso were well converged, except for KE at 0.4 m/s freestream. Heavy breathing for

these same geometries demonstrated convergence for only 2 to 3 of the degrees of

freedom. Separate mesh densities were needed to solve heavy breathing velocity field

simulations.

Although mesh dependence was indicated for most of the cases, a more refined

mesh with the same refinement factor of 1.2 was determined to be too computationally

intensive for the current work. Particle simulations were run on the least refined mesh

that indicated mesh independence. For the cases where mesh independence was not

shown, the most refined mesh density solution was used. For heavy breathing

simulations, the upward velocity (W) tended to not demonstrate monotonic convergence

at the 1e-5 tolerance level. In general, the at-rest breathing rate conditions showed

monotonic convergence for more degrees of freedom than the heavy breathing rates.
27

For degrees of freedom that did not show mesh independence, it must be assumed

that an additional mesh refinement would yield mesh independence for the following

particle simulations to be valid.

Velocity Estimates

Table 7 summarizes the range and average for the vertical velocity differences

between the torso models.

The coefficient of determination indicated a stronger relationship between the

velocities from the two elliptical torso geometries than the truncated and anatomical
geometries. The anatomical and full geometries have a better coefficient of determination

than the anatomical and truncated but still does not show a strong relationship between

the two. The relationship between the anatomical and the other two elliptical models

show the weakest relationship. This weak relationship indicates that the inclusion of the

legs within the model causes the difference in velocity.

The differences in vertical velocity were on average 0.004 m/s higher for the

anatomical model compared to the full model, and 0.012 m/s higher than the truncated

model. The vertical velocity for the full model was on average 0.007 m/s higher than the

truncated model. These results show that the complexity of the torso geometry affected

velocity estimates throughout the upstream region. Regression plots for each velocity

condition are shown in Appendix D.

Velocity Data above the Neck

Data was next segregated and analyzed according to whether it was in the face

region (z = -0.05 m, just beneath the chin and above) and the neck/chest region (below z

= -0.05 m). Vertical velocity estimates were paired between torso geometry, by

freestream velocity, breathing rate, and location. Significant differences (p <0.001) were

found between the anatomical and full model for all breathing rates and velocity
28

conditions, shown in Table 8. Linear regression models showed a weak relationship

between torso models (R2 ranged from 0.158-0.449).

The coefficient of determination showed a stronger relation for the velocity data

between the full and truncated models compared to the anatomical and full, but the

relationship was still weak. P-values indicated that the velocities were significant (p

<0.001) for all conditions.

The velocity between the anatomical and truncated model showed the weakest

relationship and all conditions showed significant differences in vertical velocity

(p<0.001).

Velocity Data below the Neck


Velocity data in the region below the neck were also analyzed (Table 9) and

showed significant differences between all models for all conditions except for between

the anatomical and full model at 0.4 m/s, heavy breathing rate (p = 0.17). The higher

freestream velocity and higher rate of suction may negate the bluff body effects that

cause the velocity differences.

Critical Area

The coordinates for critical areas for the 0.4 m/s freestream simulations for 7 m

particles at-rest and 82 m particles at heavy breathing rates are illustrated as examples

of all conditions in Figures 2 and 3. Recall that lateral symmetry was assumed in these

simulations, so these areas present only half of a critical area associated with aspiration.

The similarity in shape of critical area between torso models is apparent in these images.

Note the notch near the center of the critical area, near the top, reported by others as the

reduction in critical area associated with particle impacting the lip and nose (Figure 4,

with both halves of the critical areas shown). The notch was more pronounced for the

truncated torso simulations compared to the anatomical simulations. The difference in


29

area due to the notch, however, is a small proportion of the overall critical area that these

differences may be negligible.

These figures also illustrate differences in the position of the critical areas

between torso geometries. Table 11 presents the central locations of the critical areas, by

geometry, particle size, and velocity condition. The differences between models averaged

10.7 mm higher for the full model relative to the anatomical model and averaged 23.8

mm higher for the truncated model relative to the anatomical model. The location of the

critical area for the truncated torso geometry at all particle sizes was always located
higher than the others. The location of the critical area for the full model was always

higher than for the anatomical torso, which always had the lowest location. The

difference between the locations of the areas between the three torso models decreased

with increasing particle size. The same trend was seen for the heavy breathing conditions,

but the difference in the location of the critical area was less pronounced. These results

show that torso geometry had a greater influence on the location of the critical area for

lower breathing rates.

Linear regression models were fitted to the data and identified a strong

relationship between the critical area positions and the geometric form. The regression

equations indicate that the location of the critical area for the anatomical torso geometry

was shifted approximately 30 centimeters below the critical area for the truncated model.

The critical area for the full torso geometry was approximately 15 cm below that of the
truncated torso geometry, and 15 cm above the anatomical torso geometry.

Aspiration Efficiency

The summary of aspiration efficiencies for each particle size is shown in Table 12

(no bounce) and Table 13 (bounce). The no bounce condition reflects the minimum

anticipated aspiration efficiency, whereas the bounce condition reflects the maximum

anticipated aspiration efficiency.


30

No Bounce Simulation

For all torso models, the aspiration efficiency decreased with increasing particle

size consistently for all velocity and breathing rate conditions. For the heavy breathing

rates, aspiration efficiency decreased to approximately 30% for particles 116 µm. For the

at-rest breathing rates, aspiration efficiency dropped to zero for particles larger than 82

µm. These trends were seen regardless of torso geometry.

For smaller particle sizes, all torso geometries for all velocity conditions and

breathing rates showed aspiration decreased with increasing particle size, similar to the

ACGIH inhalability curve. Deviations from the ACGIH inhalability curve occurred at

larger particle diameters, however. Aspiration efficiency was observed to continue to

decrease with increasing particle size whereas the inhalability curve shows a leveling off

after 50 µm. At-rest breathing rates (1.81 m/s) consistently had smaller mean aspiration

efficiencies with zero aspiration for particle sizes larger than 82 µm. Aspiration

efficiencies for the heavy breathing rate at both freestream velocities were larger for all

three torso models compared to at-rest breathing.

Differences between aspiration efficiencies between the models were calculated

and ranged from -6.4 to 8.8% (Table 13). The differences between the models averaged

1.5% higher for the anatomical model compared to the full cylinder torso but only 0.8%

higher for the anatomical model compared to the truncated torso.

Linear regression models were used to determine the actual magnitude of these

differences, to assess for practical significance and to determine the relationships between

the geometry simulations and the results are shown in Table 6. The aspiration estimates

from the full torso geometry are generally 2% larger than the truncated model. The

anatomical model aspiration estimates compared to the full model are generally larger by

1.35%. The anatomical model aspiration estimates are generally 3.02% larger compared

to the truncated model. The differences in aspiration efficiency estimates are relatively
31

small, indicating that the use of a simplified truncated model will not substantially affect

aspiration estimates.

Bounce Simulation

The results for the aspiration efficiencies at bounce simulations are shown in

Table 13. For the bounce conditions aspirations efficiencies were larger than for no-

bounce conditions for all torso models and velocity conditions. Similar to the no-bounce

simulations, the bounce simulations estimated aspiration efficiencies that decreased with

increasing particle size up to 68 µm. However, at 82 µm the aspiration efficiencies

suddenly increased. When examining where particles traveled before being aspirated,

these larger particle bounced off of the top of the head, as illustrated in Figure 5, which is

not a realistic situation. The presence of hair would be more likely to trap the particles.

For smaller particles, the particles bounced on facial features near the mouth which is a

more realistic situation.

The difference in aspiration efficiency between the bounce and no bounce

conditions was investigated, and the results are summarized in Table 14. The bounce

conditions resulted in higher aspiration efficiencies in every case. Differences ranged

from 0.1% - 96.6%. The differences between the bounce and no bounce conditions

tended to be smaller for particle diameters 52 µm and less (average 3.6%), whereas

differences for particle diameters 68 µm and larger tended to be much larger (average

37.1%). The differences in aspiration efficiency between the bounce and no-bounce

condition was assessed for statistical significance and the p-values were found to be

<0.001 between all models.

Discussion

Convergence studies identified convergence problems with the heavy breathing

conditions, although the at-rest breathing conditions were relatively well solved.
32

Additional work is needed to examine alternate schemes of node spacing to improve the

solution of the heavy breathing models.

Torso geometry resulted in significantly different velocity estimates in the chest

and head region. For this study, when velocity was examined separately for regions

above the chin and regions for the neck and lower, significant velocity differences (p-

value <0.001) were found in both regions between all torso models for all but one

velocity and breathing rate condition between the anatomical and full model (0.2 m/s, at-

rest breathing). It was interesting to note that the velocity data indicated better
relationship in the region below the neck, rather than above. Overall, the velocity

estimates ranged from 0.29 m/s to -0.21 m/s, a fairly narrow range. The differences in

these values between torso geometry are very small and although statistical significance

was found, considering sampling error in many wind tunnel studies, the practicality of the

differences may be considered negligible.

The velocity results from this study compare will with results from other studies.

Several studies have found large velocity changes compared to the freestream velocity in

the chest region, but smaller velocity changes in the facial region. Smith and Bird (2002)

found that velocity was decreased and reached a minimum in the chest region, but in the

face region there was less deceleration of velocity. Topp et al, (2003) also noted velocity

changes in the chest region but not as large changes in the head region. Thermal studies,

although lacking inhalation, have also identified this trend. Those studies found that
while vertical (Z) velocity data at the chest height differed between models, personal

exposures did not.

The presence of legs has been found to influence velocity profiles as well. Topp et

al. (2003) found that when legs were excluded from the model, the velocity at the center

of the torso was significantly lower. Brohus (1996) found similar results that the

inclusion of legs in the model caused a decrease in the upward vertical velocity in the

region of the head. This also compared well with results in this study.
33

Torso geometry was associated with 15 to 30 mm changes in the location of the

area where particles were aspirated. The location of the critical area was found to shift

downward with increasing torso complexity. Although these position differences were

statistically significant, aspiration efficiency studies do not rely on precise position

information in aspiration calculations. Rather, aspiration efficiency work requires the

assumption of a uniform particle concentration across the entire wind tunnel cross

section, including the relatively small critical areas. However, if using CFD techniques to

examine exposures from a point source, these differences in position may be significant.
Hence, torso simplifications are not recommended.

Anthony and Flynn (2006B) proposed that velocity differences associated with

torso simplification (truncation, simple cylindrical shape) may have affected aspiration,

accounting for differences between their computational simulation results and Kennedy

and Hinds (2005) wind tunnel experiments. However, this work concludes that the

complexity of torso geometry had minimal effect on aspiration efficiency, even though

velocity and position of critical areas were influenced by the torso geometry. The

differences in aspiration efficiency by particle size and velocity condition were at most

8.7% and averaged between -0.76 to 1.55% over the test conditions studied here. The

between-geometry differences were less important than between breathing-rate

differences, where 4 to 35% differences were seen. Given the uncertainties and sampling

error in experimental aerosol studies, these between-geometry differences might be


considered negligible.

Although aspiration efficiencies were estimated for four separate velocity

conditions, realistically the conditions a worker encounters would vary across all of these

separate conditions. The conditions would vary from at-rest to heavy breathing and from

less than 0.2 m/s to 0.4 m/s or higher. To more realistically estimate the aspiration

efficiencies a worker might experience during working conditions, aspiration efficiency

for each torso was averaged for all four conditions and compared to facing the wind data
34

from Kennedy and Hinds (2002). Aspiration efficiencies were higher for particles

smaller than 68 µm for all three torso geometries compared to aspiration efficiencies

from Kennedy and Hinds (2002). However, for particles larger than 68 µm, aspiration

efficiencies were lower than those found by Kennedy and Hinds. There is still the

limitation that the results presented here are only for facing-the-wind orientation, not

averaged over 360°. These averaged results are shown in Figure 6. The bars represent the

standard error over all test conditions.

Both bounce and no bounce simulations were investigated in the study. The
estimates of aspiration efficiency for the no-bounce conditions were found to decrease

with increasing particle size for all velocity and breathing rate conditions. This trend

agreed with data published previously (Anthony, 2006), although the aspiration estimates

are larger than those found by King Se et al. (2010). Assuming only no-bounce

conditions could result in an underestimation of aspiration efficiencies in conditions

where particles have the potential to bounce. For the bounce condition, the location

where particles hit the head and bounced in became an important consideration. In the

real world, particles that hit the nose and or lips might realistically bounce into the

mouth, but particles that hit on the forehead or top of the head would not realistically

bounce and be inhaled but could be trapped in the hair. One hundred percent reflection

was assumed, however this is a simplification and the nature of the particle (solid or

liquid) and the facial elasticity would influence the amount of reflection in reality.
The aspiration efficiencies were consistently larger for the bounce condition for

all torso geometries and velocity conditions. Whether particles actually are capable of

bouncing off the facial feature surfaces, from the lips up to the top of the head, are not

quantified in this study. However, the bounce estimates provide the maximum aspiration

efficiency estimate for human aspiration, as 100% reflection was assumed.

For particles released within the bounce critical areas, these particles will either

terminate on the face (dermal deposition, with 0% reflection) or bounce into the mouth
35

(aspiration, with 100% reflection). The nature of both the particle (liquid, solid) and the

facial elasticity would affect whether the particle bounces or deposits. Particles outside

this area would certainly be transported away from the humanoid form and be carried out

of the computational domain. This difference between bounce and no-bounce areas

provides us with an indication of the overall contribution of possible dermal deposition,

although not implicitly studied in this work.

This study confirms that using a simplified torso is a reasonable approximation to

complex human form for aspiration efficiency studies. However, this work was limited
to the facing-the-wind orientation, and as the torso turns relative to the oncoming wind,

wake effects due to different torso effects are less likely to be the same. The standard k-ε

turbulence model was used in this study, which was adequate for facing-the-wind

orientation. However, the standard k-ε model has problems with flow separation so

additional turbulence models should be investigated with torso rotation.

Aspiration efficiency for the bounce condition assumed one hundred percent

reflection which was a simplification. Particle bounce would be dependent on the nature

of the particle (liquid or solid) and the facial elasticity. More research needs to be

conducted with values for reflection to obtain a more accurate estimate of aspiration

efficiency with bounce.

For Reynolds numbers greater than one, the particle is outside of the Stokes

region. Although Stokes law can be used to calculate the terminal settling velocity, the
value will be within 3% for Reynolds numbers ranging from one to six hundred.

Conclusion

This research tested the hypothesis that torso shape and height affect velocity,

location of critical area, and resulting estimates of aspiration efficiency, for both bounce

and no bounce conditions. This work found that vertical velocity and positions of critical

areas were significantly different between torso geometries. However, comparisons of


36

aspiration efficiency estimates between the simplified cylindrical, truncated torso and the

fully anatomical torso yielded a maximum 8.7% difference, but averaged less than 1%

over all velocity and particle sizes studied here. These differences are insignificant

relative to between-velocity conditions, particularly when comparing at-rest to heavy

breathing aspiration efficiency estimates, which can be as large as a difference of 35%.

The results from this study show that the use of the truncated model as a surrogate for a

complex human form is appropriate for aspiration efficiency studies. However, for

studies investigating velocity flow or aspiration from a point source, a more realistic
anthropometrically complex model is necessary.
37

Table 1: Comparison of dimensions between the anatomical model and 50th percentile female
dimensions
Surface Anatomical 50th Percentile Female Difference
Geometry Dimensions (%)
Dimension (m)
Height 1.56 1.625 1.04
Shoulder height 1.34 1.325 0.99
Elbow height 1.11 1.02 0.91
Hip height 0.99 0.835 0.84
Knee height 0.67 0.505 0.75
Width at temple 0.086 0.072 1.19
Chin to top of head 0.19 0.218 0.87
Center of mouth to top 0.154 0.178 0.86
of head
Forehead to back of 0.155 0.18 0.86
head
38

Figuree 1: Illustratiions of trunccated, full, annd anatomiccal geometriees


39

Table 2: Mesh refinement details


Geometry Mesh Density
Coarse Moderate Mod-Fine Fine
Truncated Nodes 291,532 621,272 1,507,780 1,113,039
Elements 1,471,892 3,303,299 8,245,940 6,000,132

Full Nodes 172,028 316,395 na 1,112,918


Elements 841,508 1,634,149 na 6,619,437

Anatomical Nodes 732,788 1,291,572 1,731,694 2,308,915


Elements 3,948,179 7,018,339 9,522,764 12,830,832

Table 3: Mesh refinement ratios


Geometry Coarse- Moderate- Moderate- Mod-Fine-
Moderate Fine Mod-Fine Fine
Truncated 1.287 1.214 1.097 1.107
Full 1.225 1.225 - -
Anatomical 1.208 1.214 1.103 1.101
40

Table 4: Boundary conditions


Surface Normal Tangential Turbulence Turbulent
Velocity velocity Kinetic dissipation
(m/s) (m/s) Energy (m2/s3)
(m2/s2)
Domain entrance 0.2, 0 0.000384 / 8.705661e-5 /
0.4 0.001536 0.001392906

Mouth inlet -1.81 or 0 δk/ δn = 0 δε/ δn = 0


-12.11
Domain Exit δu/ δn = 0 δu/ δτ = 0 δk/ δn = 0 δε/ δn = 0
Walls 0 0 0 0
Symmetry 0 δu/ δτ = 0 δk/ δn = 0 δε/ δn = 0
Mid-sagittal plane 0 δu/ δτ = 0 δk/ δn = 0 δε/ δn = 0
Head and torso 0 0 δk/ δn = 0 δε/ δn = 0
41

Table 5: L2 error norms for global convergence tolerance limits 1e-4 to 1e-5
Velocity/Breathing
Geometry Rate Mesh U V W KE E SP
Anatomical 0.4 m/s, 12.11 m/s Coarse (M1) 0.004 0.013 0.023 0.001 0.002 0.001
Moderate (M2) 0.004 0.014 0.023 0.002 0.003 0.001
Fine(M3) 0.005 0.017 0.029 0.003 0.005 0.001

0.4 m/s, 1.81 m/s Coarse (M1) 0.002 0.008 0.017 0.010 0.013 0.001
Moderate (M2) 0.004 0.013 0.025 0.017 0.018 0.001
Fine(M3) 0.005 0.016 0.028 0.019 0.020 0.001

0.2 m/s, 12.11 m/s Coarse (M1) 0.006 0.018 0.028 0.003 0.004 0.002
Moderate (M2) 0.006 0.020 0.030 0.003 0.004 0.002
Fine(M3) 0.007 0.022 0.035 0.003 0.004 0.002

0.2 m/s, 1.81 m/s Coarse (M1) 0.004 0.010 0.016 0.008 0.011 0.002
Moderate (M2) 0.006 0.017 0.028 0.014 0.021 0.002
Fine(M3) 0.007 0.022 0.036 0.017 0.028 0.002

Full 0.4 m/s, 12.11 m/s Coarse (M1) 0.002 0.006 0.015 0.008 0.011 0.000
Moderate (M2) 0.003 0.009 0.027 0.014 0.023 0.000
Fine(M3) 0.003 0.011 0.031 0.006 0.002 0.000

0.4 m/s, 1.81 m/s Coarse (M1) 0.002 0.006 0.015 0.008 0.011 0.000
Moderate (M2) 0.003 0.009 0.027 0.014 0.023 0.000
Fine(M3) 0.003 0.011 0.031 0.006 0.002 0.000

0.2 m/s, 12.11 m/s Coarse (M1) 0.004 0.012 0.031 0.000 0.000 0.001
Moderate (M2) 0.004 0.013 0.030 0.001 0.002 0.001
Fine(M3) 0.004 0.013 0.032 0.001 0.001 0.001

0.2 m/s, 1.81 m/s Coarse (M1) 0.003 0.008 0.018 0.008 0.003 0.001
Moderate (M2) 0.004 0.012 0.032 0.014 0.012 0.001
Fine(M3) 0.006 0.025 0.104 0.096 0.083 0.004

Truncated 0.4 m/s, 12.11 m/s Coarse (M1) 0.000 0.002 0.005 0.001 0.000 0.000
Moderate (M2) 0.001 0.002 0.005 0.001 0.001 0.000
Fine(M3) 0.001 0.003 0.007 0.000 0.000 0.000

0.4 m/s, 1.81 m/s Coarse (M1) 0.000 0.002 0.004 0.004 0.005 0.000
Moderate (M2) 0.001 0.003 0.007 0.007 0.010 0.000
Fine(M3) 0.001 0.003 0.008 0.006 0.010 0.000
42

Table 5 continued

0.2 m/s, 12.11 m/s Coarse (M1) 0.001 0.003 0.007 0.000 0.000 0.001
Moderate (M2) 0.001 0.003 0.007 0.001 0.001 0.001
Fine(M3) 0.001 0.003 0.008 0.000 0.000 0.001

0.2 m/s, 1.81 m/s Coarse (M1) 0.001 0.002 0.005 0.004 0.002 0.000
Moderate (M2) 0.001 0.003 0.009 0.006 0.002 0.001
Fine(M3) 0.001 0.004 0.010 0.005 0.002 0.001
43

Table 6: Between-mesh R2 error norms for 1e-5 GSE tolerance levels


Torso Velocity Breathing
Geometry Condition Rate Ux Uy Uz KE E SP
Anatomical 0.4 m/s 1.81 m/s 0.49 1
0.63 0.40 0.45 0.16 0.24
Torso 12.11 m/s 1.342 0.95 1.09 1.11 0.39 0.94

0.2 m/s 1.81 m/s 0.45 0.63 0.40 0.20 0.12 0.27
12.11 m/s 1.20 0.97 1.09 1.05 0.44 1.12

Full torso 0.4 m/s 1.81 m/s 0.83 0.73 0.90 0.84 1.64 1.00
12.11 m/s 1.00 1.00 1.01 1.00 1.06 1.00

0.2 m/s 1.81 m/s 1.05 1.09 0.84 1.13 1.09 0.94
12.11 m/s 0.38 0.84 0.97 0.50 0.45 1.33

Truncated 0.4 m/s 1.81 m/s 0.27 0.15 0.58 1.08 0.52 0.44
torso 12.11 m/s 1.05 0.76 1.27 1.10 1.42 0.57

0.2 m/s 1.81 m/s 0.22 0.14 0.47 0.65 0.35 0.29
12.11 m/s 1.11 0.75 1.35 1.12 1.41 0.62
1
Bold values indicate monotonic convergence (i.e. an acceptable value)
2
values in italics indicates a degree of freedom that is not monotonically

converging
44

Table 7: Summary of differences between torso models.


Vertical Velocity Critical Area Aspiration Aspiration
(m/s) Midpoint (mm) Efficiency-No Efficiency-Bounce
Bounce (%) (%)

Range Avg Range Avg Range Avg Range Avg

Anatomical -0.21 - 0.29 0.004 -18 - -3.2 -10.7 -2.9 - 7.2 1.55 -35 - 29 -0.50
-Full

Anatomical -0.18 - 0.28 0.012 -36 - -14 -23.8 -6.1 - 8.7 0.79 -14 - 6.6 -1.00
-Truncated

Full- -0.07 - 0.13 0.007 -17 - -8.3 -13.1 -6.3 - 2.0 -0.76 -22 - 39 -0.50
Truncated
45

Table 8: Between-geometry comparisons. The p-value indicates the results from paired t-tests.
Dependent Independent
variable Slope variable intercept R2 r p-value1

Vertical
Velocity Anatomical 0.706 Full 0.011 0.393 0.627 <0.001

Full 0.916 Truncated 0.009 0.684 0.827 <0.001

Anatomical 0.421 Truncated 0.020 0.114 0.337 <0.001

Critical Area
Midpoint Anatomical 1.020 Full -0.014 0.999 0.999 <0.001

Full 1.012 Truncated -0.016 0.999 0.999 <0.001

Anatomical 1.031 Truncated -0.030 0.999 0.999 <0.001

Aspiration
Efficiency Anatomical 0.986 Full 0.024 0.995 0.998 0.003

(No Bounce
Condition) Full 0.976 Truncated 0.008 0.996 0.998 0.115

Anatomical 0.964 Truncated 0.032 0.992 0.996 0.224

Aspiration
Efficiency Anatomical 1.066 Full -0.063 0.872 0.934 0.773

(Bounce
Condition) Full 0.789 Truncated 0.002 0.880 0.938 0.787

Anatomical 0.947 Truncated 0.036 0.972 0.986 0.251


1
p-value indicates results from pair two-tailed t-tests
46

Table 9: Linear regression models for velocity data above the neck.
Breathing Dependent Independent p-
1
Velocity Condition variable Slope variable Intercept R2 value

0.2 12.11 Anatomical 0.427 Full 0.014 0.158 <0.001

0.2 1.81 Anatomical 0.766 Full 0.009 0.449 <0.001

0.4 12.11 Anatomical 0.543 Full 0.023 0.238 <0.001

0.4 1.81 Anatomical 0.642 Full 0.019 0.318 <0.001

0.2 12.11 Full 0.873 Truncated 0.008 0.695 <0.001

0.2 1.81 Full 0.515 Truncated 0.012 0.258 <0.001

0.4 12.11 Full 0.892 Truncated 0.015 0.720 <0.001

0.4 1.81 Full 0.862 Truncated 0.015 0.691 <0.001

0.2 12.11 Anatomical 0.084 Truncated 0.021 0.006 <0.001

0.2 1.81 Anatomical 0.265 Truncated 0.019 0.052 <0.001

0.4 12.11 Anatomical 0.228 Truncated 0.036 0.038 <0.001

0.4 1.81 Anatomical 0.327 Truncated 0.034 0.077 <0.001


1
p-value indicates results from paired two-tailed t-tests
47

Table 10: Linear regression models for velocity data below the neck
Breathing Dependent Independent p-
1
Velocity Condition variable Slope variable Intercept R2 value
0.2 12.11 Anatomical 1.325 Full 0.001 0.826 <0.001
0.2 1.81 Anatomical -0.013 Full 0.004 0.001 0.167
0.4 12.11 Anatomical 1.265 Full 0.001 0.826 <0.001
0.4 1.81 Anatomical 1.262 Full 0.001 0.829 <0.001

0.2 12.11 Full 0.959 Truncated 0.001 0.864 <0.001


0.2 1.81 Full -0.272 Truncated 0.005 0.006 <0.001
0.4 12.11 Full 0.955 Truncated 0.001 0.867 <0.001
0.4 1.81 Full 0.931 Truncated 0.001 0.868 <0.001

0.2 12.11 Anatomical 1.260 Truncated 0.002 0.702 <0.001


0.2 1.81 Anatomical 1.223 Truncated 0.002 0.723 <0.001
0.4 12.11 Anatomical 1.204 Truncated 0.003 0.711 <0.001
0.4 1.81 Anatomical 1.170 Truncated 0.003 0.714 <0.001
1
p-value indicates results from paired two-tailed t-tests
48

Table 11: Velocity minimum and maximum values


Torso Geometry Anatomical Full Truncated

Min Max Min Max Min Max


Breathing Velocity Velocity Velocity Velocity Velocity Velocity
Velocity Condition (m/s) (m/s) (m/s) (m/s) (m/s) (m/s)

0.2 12.11 -0.080 0.098 -0.105 0.098 -0.091 0.110

0.2 1.81 -0.002 0.104 -0.016 0.107 -0.057 0.094

0.4 12.11 -0.075 0.205 -0.110 0.233 -0.098 0.208

0.4 1.81 0.001 0.201 -0.043 0.220 -0.091 0.192


49

Figure 2: Critical area for 7 m particle in 0.4 m/s freestream velocity with at-rest
breathing inhalation (1.81 m/s). Particles released 0.75 m upstream from mouth opening.

‐0.01
Truncated

‐0.02

‐0.03
Vertical (Z) Position, m

Full
‐0.04

‐0.05

‐0.06 Anatomical

‐0.07

‐0.08

‐0.09
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
Lateral (Y) Position, m
50

Figure 3: Critical area for 82 m particle in 0.4 m/s freestream velocity with heavy
breathing inhalation (12.11 m/s). Particles released 0.75 m upstream from mouth
opening.

0.2

Truncated
0.195

0.19

0.185
Vertical (Z) Position, m

0.18

0.175 Full

0.17

0.165
Anatomical
0.16

0.155

0.15
0 0.01 0.02 0.03 0.04 0.05
Lateral (Y) Position, m
51

Figure 4: Critical area for 7 m particle for 0.4 m/s freestream velocity and at-rest

breathing (1.81 m/s). Particles released 0.75 m upstream from mouth opening.

‐0.01 Truncated

‐0.02

‐0.03
Vertical (Z) Position, m

‐0.04

‐0.05

Anatomical
‐0.06

‐0.07

‐0.08

‐0.09
‐0.045 ‐0.035 ‐0.025 ‐0.015 ‐0.005 0.005 0.015 0.025 0.035 0.045
Lateral (Y) Position, m
52

Table 12: Location of critical area midpoint


Torso Geometry
Particle Velocity and
Truncated Full Anatomical
Diameter (um) Breathing condition
Midpoint location (mm)

7 0.2, 1.81 -29.70 -46.57 -64.49

0.2, 12.11 -30.40 -47.66 -66.54

0.4, 1.81 -31.01 -47.64 -60.71

0.4, 12.11 -32.15 -48.54 -63.89

22 0.2, 1.81 21.24 6.98 -10.77

0.2, 12.11 20.49 4.24 -13.76

0.4, 1.81 -5.00 -21.14 -33.84

0.4, 12.11 -6.36 -22.32 -36.62

52 0.2, 1.81 281.59 270.38 258.13

0.2, 12.11 278.03 264.87 252.53

0.4, 1.81 127.63 113.16 103.56

0.4, 12.11 124.95 111.52 98.76

68 0.2, 1.81 482.71 469.72 466.52

0.2, 12.11 474.02 462.93 453.03

0.4, 1.81 231.87 217.93 209.95

0.4, 12.11 227.02 214.32 203.46

82 0.2, 1.81 375.93 366.89 359.90

0.2, 12.11 356.39 347.93 341.54

0.4, 1.81 183.90 172.05 166.72

0.4, 12.11 178.99 168.96 160.15

100 0.2, 1.81 na na na

0.2, 12.11 493.56 484.05 475.40

0.4, 1.81 na na na

0.4, 12.11 252.22 238.56 235.00

116 0.2, 1.81 na na na

0.2, 12.11 632.02 623.69 618.03

0.4, 1.81 na na na
0.4, 12.11 322.25 311.41 306.06

*for these velocity conditions, there was no aspiration at these particle sizes
53

Table 13: Aspiration efficiency for all torso geometries

Truncated Full Anatomical

Freestream
Velocity
(m/s) 0.2 0.2 0.4 0.4 0.2 0.2 0.4 0.4 0.2 0.2 0.4 0.4

Breathing
Rate (m/s) 1.81 12.11 1.81 12.11 1.81 12.11 1.81 12.11 1.81 12.11 1.81 12.11

Particle Mean Aspiration Mean Aspiration Mean Aspiration


size (um) Efficiency (%) Efficiency (%) Efficiency (%)

7 98.4 99.4 98.2 103.0 97.6 99.4 98.1 98.6 97.9 99.0 97.6 98.8

22 97.3 98.0 97.6 104.3 97.7 98.2 97.5 98.1 96.9 98.3 96.8 98.2

52 85.8 93.3 79.7 97.1 86.4 93.0 81.7 91.6 83.5 93.1 83.5 92.2

68 61.8 89.0 56.4 84.5 63.1 88.7 57.5 78.2 69.9 89.7 58.9 83.2

82 21.1 84.7 28.2 73.3 22.9 84.8 30.1 68.7 26.0 87.3 34.3 74.9

100 0.0 48.2 0.0 50.5 0.0 49.7 0.0 47.5 0.0 56.9 0.0 51.7

116 0.0 26.2 0.0 34.2 0.0 27.1 0.0 32.7 0.0 28.6 0.0 35.1

*No Bounce Conditions


54

Table 14: Difference in aspiration efficiencies between models


Difference between  Difference between  Difference between 
Truncated and Full  Truncated and Anatomical  Full and Anatomical 

Freestream 
Velocity 
(m/s)  0.2  0.2  0.4  0.4  0.2  0.2  0.4  0.4  0.2  0.2  0.4  0.4 

Breathing 
Rate (m/s)  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11 

Particle 
size (um)  Mean Aspiration Efficiency (%)  Mean Aspiration Efficiency (%)  Mean Aspiration Efficiency (%) 

7  ‐0.9  0.1  ‐0.1  ‐4.3  ‐0.5  ‐0.4  ‐0.6  ‐4.1  0.3  ‐0.4  ‐0.5  0.2 

22  0.5  0.2  ‐0.1  ‐6.2  ‐0.4  0.3  ‐0.8  ‐6.1  ‐0.9  0.1  ‐0.7  0.1 

52  0.6  ‐0.3  2.0  ‐5.5  ‐2.3  ‐0.1  3.8  ‐4.9  ‐2.9  0.1  1.8  0.6 

68  1.2  ‐0.3  1.1  ‐6.4  8.1  0.7  2.5  ‐1.4  6.9  1.0  1.4  5.0 

82  1.8  0.1  1.9  ‐4.6  4.9  2.6  6.1  1.7  3.1  2.5  4.2  6.3 

100  0.0  1.6  0.0  ‐3.0  0.0  8.8  0.0  1.2  0.0  7.2  0.0  4.2 

116  0.0  0.9  0.0  ‐1.5  0.0  2.4  0.0  0.9  0.0  1.5  0.0  2.4 
55

Figure 5: Illustration of Bounce


B for 116 µm particcles . Particlles were releeased at the uupper
nd lower bou
an unds of the critical
c area. Particles relleased withiin these two points will eeither
deposit on thee face or be inhaled.
56

1.2

1.0
Aspiration Efficiency Fraction

0.8

Kennedy and Hinds (2002)
0.6
Truncated Model

0.4 Full Cylindrical Model
Realistic Human Torso
0.2

0.0
0 20 40 60 80 100 120
Aerodynamic particle diameter, µm

Figure 6: Comparison of aspiration efficiency for truncated, full, and anatomical torso
models to experimental data from Kennedy and Hinds (2002)
57

Table 15: Aspiration Efficiency for All Torso Geometries with Bounce

Truncated  Full  Annie 

Freestream 
Velocity 
(m/s)  0.2  0.2  0.4  0.4  0.2  0.2  0.4  0.4  0.2  0.2  0.4  0.4 

Breathing 
Rate (m/s)  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11 

Particle 
size (um)  Mean Aspiration Efficiency (%)  Mean Aspiration Efficiency (%)  Mean Aspiration Efficiency (%) 

7  99.9  100.4  99.5  106.4  99.7  100.3  99.2  99.6  99.1  99.8  99.0  99.6 

22  98.7  99.1  99.1  105.7  98.1  99.0  99.0  99.0  98.3  99.0  99.1  98.8 

52  88.9  98.1  90.4  101.5  95.8  98.6  92.0  95.2  93.5  96.5  92.9  95.4 

68  96.1  97.1  82.1  97.1  101.9  98.2  83.3  90.5  92.9  95.5  83.0  83.2 

82  98.2  99.4  76.3  99.4  97.9  100.2  71.7  93.4  102.5  100.1  73.1  94.4 

100  89.9  95.8  64.3  96.2  67.6  95.5  66.2  90.8  96.6  98.2  69.2  91.6 

116  0.0  84.9  0.0  88.9  0.0  85.1  39.1  82.5  0.0  87.0  3.9  83.2 
58

Table 16: Difference in Aspiration Efficiencies between Bounce and No Bounce Simulations
Truncated  Full  Anatomical 

Freestream 
Velocity 
(m/s)  0.2  0.2  0.4  0.4  0.2  0.2  0.4  0.4  0.2  0.2  0.4  0.4 

Breathing 
Rate (m/s)  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11  1.81  12.11 

Particle 
size (um)  Mean Aspiration Efficiency (%)  Mean Aspiration Efficiency (%)  Mean Aspiration Efficiency (%) 

7  1.4  1.0  1.3  3.4  2.1  0.9  1.1  0.9  1.2  0.8  1.4  0.8 

22  1.4  1.1  1.4  1.4  0.4  0.8  1.5  0.9  1.4  0.7  2.3  0.6 

52  3.1  4.8  10.7  4.4  9.4  5.6  10.3  3.6  10.0  3.3  9.3  3.2 

68  34.3  8.1  25.7  12.6  38.9  9.5  25.7  12.3  23.0  5.7  24.0  0.1 

82  77.1  14.6  48.1  26.1  75.0  15.3  41.7  24.7  76.5  12.8  38.8  19.4 

100  89.9  47.7  64.3  45.7  67.6  45.8  66.2  43.3  96.6  41.3  69.2  39.9 

116  0.0  58.7  0.0  54.7  0.0  58.0  39.1  49.8  0.0  58.4  3.9  48.1 
59

CHAPTER III

CONCLUSION

Aspiration efficiencies in this study were found to agree well with published data

(Anthony, 2010 and King Se, et. al., 2010), where aspiration efficiency was found to

decrease with increasing particle size and did not level off at 50% aspiration efficiency as

described by the ACGIH sampling criterion. The same trends seen previously in thermal

studies, increased velocity in the chest with smaller increases in the facial region, were

also observed here. Differences of aspiration efficiency between simplified and


anatomically representative torso geometry models, ranged from -6.9 to 8.8%, which was

considerably less than between breathing rate differences that ranged up to differences of

35%. This gives a degree of confidence that the results produced by previous studies

using simplified torsos are not influenced by torso complexity. The use of a simplified

cylindrical torso, truncated at waist height, is sufficient to model particle aspiration

efficiencies, due to the reliance on the assumption of uniform particle concentration in the

freestream. However, the use of the truncated model is not appropriate when needing to

examine velocity or particle transport from a source: the use of a more anatomically

detailed geometry would be necessary.

This work only examined the facing-the-wind orientation. As the torso turns

relative to the oncoming wind, wake effects due to different torso effects are less likely to

be the same. Further work should investigate different turbulence models, as the standard
k-ε may not be the most appropriate model for the wake effects.

Further study should be conducted examining the mesh for the heavy breathing

condition to achieve better mesh convergence. Additional turbulence models should be

examined, as well as sensitivity tests with the path length in particle trajectory

calculations.
60

APPENDIX A

MESH REEFINEMENT DETAILS

Torso Geometry Anatomical Full

Mesh Density Coarse Moderate Mod-Fine Fine Coarse Moderate Fine

Face Min 1.60E-04 1.30E-04 1.18E-04 1.00E-04 1.60E-04 1.30E-04 1.00E-04

Face Max 3.20E-02 2.60E-02 2.25E-02 2.00E-02 3.20E-02 2.60E-02 2.00E-02

Tet Max 3.20E-02 2.60E-02 2.25E-02 2.00E-02 3.20E-02 2.60E-02 2.00E-02

mouth inlet 1.01E-03 7.80E-04 6.80E-04 6.00E-04 1.01E-03 7.80E-04 6.00E-04

Lips 1.50E-03 1.17E-03 1.05E-03 9.00E-04 1.50E-03 1.17E-03 9.00E-04

unox 1.50E-03 1.17E-03 1.05E-03 9.00E-04 1.50E-03 1.17E-03 9.00E+04

nose 2.00E-03 1.56E-03 1.40E-03 1.20E-03 2.00E-03 1.56E-03 1.20E-03

cheek, eye 4.00E-03 3.12E-03 2.90E-03 2.40E-03 4.00E-03 3.10E-03 2.40E-03

torso 7.58E-03 5.83E-03 5.10E-03 4.48E-03 1.53E-02 1.17E-02 8.90E-03


legs 7.58E-03 5.83E-03 5.10E-03 4.48E-03

arm 6.00E-03 4.60E-03 4.00E-03 3.54E-03

Number of
Nodes 732,788 1,291,572 1,731,694 2,308,915 621,272 1,113,039 2,047,294

Number of
Elements 3,948,179 7,018,339 9,522,764 12,830,832 3,303,299 6,000,132 11,319,228
61

Torso Geometry Truncated

Mesh Density Coarse Moderate Mod-Fine Fine

Face Min 1.60E-04 1.30E-04 1.15E-04 1.00E-04

Face Max 3.20E-02 2.60E-02 2.25E-02 2.00E-02

Tet Max 3.20E-02 2.60E-02 2.25E-02 2.00E-02

mouth inlet 1.01E-03 7.80E-04 6.90E-04 6.00E-04

Lips 1.50E-03 1.17E-03 1.03E-03 9.00E-04

unox 1.50E-03 1.17E-03 1.03E-03 9.00E-04

nose 2.00E-03 1.56E-03 1.38E-03 1.20E-03

cheek, eye 4.00E-05 3.12E-03 2.81E-03 2.40E-03

torso

legs

arm

Number of Nodes 316,395 582,200 1,507,780 1,112,918

Number of
Elements 1,634,149 3,059,413 8,245,940 6,619,437
62

APPENDIX B

RAKE/DATA EXTRATION POINTS

The (X,Y) coordinates, in m, used for data extraction are indicated below. For each

combination, 19 equidistant Z positions were examined (-0.3 to -.6 m). These resulted in a total

of 1710 extraction locations, used to assess mesh convergence, compare velocity fields between

the three torso geometries.


X Y X Y X Y
-0.015 0 -0.03 0 -0.25 0
-0.015 0.025 -0.03 0.025 -0.25 0.025
-0.015 0.05 -0.03 0.05 -0.25 0.05
-0.015 0.075 -0.03 0.075 -0.25 0.075
-0.015 0.1 -0.03 0.1 -0.25 0.1
-0.015 0.125 -0.03 0.125 -0.25 0.125
-0.015 0.15 -0.03 0.15 -0.25 0.15
-0.015 0.25 -0.03 0.25 -0.25 0.25
-0.015 0.5 -0.03 0.5 -0.25 0.5
-0.015 0.75 -0.03 0.75 -0.25 0.75
-0.02 0 -0.05 0 -0.5 0
-0.02 0.025 -0.05 0.025 -0.5 0.025
-0.02 0.05 -0.05 0.05 -0.5 0.05
-0.02 0.075 -0.05 0.075 -0.5 0.075
-0.02 0.1 -0.05 0.1 -0.5 0.1
-0.02 0.125 -0.05 0.125 -0.5 0.125
-0.02 0.15 -0.05 0.15 -0.5 0.15
-0.02 0.25 -0.05 0.25 -0.5 0.25
-0.02 0.5 -0.05 0.5 -0.5 0.5
-0.02 0.75 -0.05 0.75 -0.5 0.75
-0.025 0 -0.1 0 -0.75 0
-0.025 0.025 -0.1 0.025 -0.75 0.025
-0.025 0.05 -0.1 0.05 -0.75 0.05
-0.025 0.075 -0.1 0.075 -0.75 0.075
-0.025 0.1 -0.1 0.1 -0.75 0.1
-0.025 0.125 -0.1 0.125 -0.75 0.125
-0.025 0.15 -0.1 0.15 -0.75 0.15
-0.025 0.25 -0.1 0.25 -0.75 0.25
-0.025 0.5 -0.1 0.5 -0.75 0.5
-0.025 0.75 -0.1 0.75 -0.75 0.75
63

APPENDIX C

RAKE BUILDING AND DATA EXTRACTION PROCEDURE

Rakes were set up to extract velocity data in the mouth region. The basic procedure for

setting up rakes was to click on the Surface button, and choose line/rake. This opened a new

window. In that window the rake type was changed to rake. The number of points was changed

to 19. The x, y z coordinates were imputed manually. Z initial was set to -0.3 and Z1 was set to

0.6 for all x. y combinations. The number of points for each rake was always 19. X and Y

coordinates were imputed by hand for all the points listed in the table in the Appendix.

Each rake was saved with a unique identifying name. After all of the rakes were set up,

the case file was exported. To differentiate from the solution file, the case was saved with –rake

in the name. Only one rake was set up for each mesh density and each geometry and then the

solutions for every velocity condition was imported into the rake case with the matching

geometry and mesh density.

To extract data first a new workbench was set up. The workbench was saved with the

following naming convention: extraction-geometry-mesh density. A new Fluent box was

dragged into the Workbench. The rake file for the same mesh density as the files you wanted to

extract was imported in. Double precision was checked, but the solution was not run another

iteration. The data file for the velocity and solution conditions that was being set up for

extraction was brought it. Since the case file and the data file have different conditions, no

iterations were run and it was not necessary to do so to extract data. Under Plots, the XY plots

was selected and set up clicked. This opened up a new window called Solution XY plots. The

Write to File was selected on the left. Six files were written. The box under the Y axis function

was changed to select the appropriate degree of freedom. All of the rakes under surfaces were

selected and then at the bottom of the window the write button was selected. Total pressure, x

velocity, y velocity, z velocity, kinetic energy, and kinetic dissipation rate were all written to
64

files. After all the files were extracted Fluent was closed without saving anything so as to not

write over either the rake file or the solution data.


65

APPENDIX D

L2 AND R2 ERROR NORMS

Geometry: Anatomical

Freestream 0.2 m/s


velocity

Breathing velocity 12.11 m/s

L2 Error Norms

Mesh GSE Levels U V W KE E SP


-3 -4
Coarse (M1) 10 - 10 0.551% 2.332% 8.202% 3.125% 3.223% 1.861%

10-4 - 10-5 0.570% 1.767% 2.756% 0.277% 0.390% 0.216%

Moderate (M2) 10-3 - 10-4 0.550% 2.418% 8.904% 3.407% 3.264% 2.077%
-4 -5
10 - 10 0.619% 1.958% 3.025% 0.310% 0.448% 0.237%

Fine (M3) 10-3 - 10-4 0.661% 3.006% 10.887% 2.647% 0.848% 1.707%

10-4 - 10-5 0.698% 2.239% 3.474% 0.295% 0.401% 0.205%

R2 Error Norms

GSE level U V W KE E SP

1E-03 1.200367617 0.9680993 1.0904974 1.135137 0.7214302 1.3125667

1E-04 1.1980 0.9671 1.0956 1.0486 0.4354 1.1217

1E-05 1.1958 0.9667 1.0913 1.0526 0.4409 1.1162


66

Geometry: Anatomical

Freestream velocity 0.2 m/s

Breathing velocity 1.81 m/s

L2 Error Norms

Mesh GSE Levels U V W KE E SP

Coarse (M1) 10-3 - 10-4 0.720% 2.785% 7.886% 4.261% 5.838% 0.156%
-4 -5
10 - 10 0.356% 1.018% 1.634% 0.781% 1.057% 0.181%

Moderate (M2) 10-3 - 10-4 0.534% 2.374% 8.802% 9.838% 9.789% 0.316%

10-4 - 10-5 0.591% 1.742% 2.822% 1.415% 2.076% 0.206%

Fine (M3) 10-3 - 10-4 0.656% 3.184% 12.575% 14.282% 11.526% 0.530%
-4 -5
10 - 10 0.734% 2.219% 3.580% 1.733% 2.807% 0.200%

R2 Error Norms

GSE level U V W KE E SP

1E-03 0.443555942 0.6300591 0.3966552 0.1917282 0.1197467 0.2606322

1E-04 0.4517 0.6281 0.4035 0.2011 0.1250 0.2762

1E-05 0.4543 0.6267 0.4015 0.1966 0.1199 0.2739


67

Geometry: Anatomical

Freestream 0.4 m/s


velocity

Breathing velocity 12.11 m/s

L2 Error Norms

Mesh GSE Levels U V W KE E SP

Coarse (M1) 10-3 - 10-4 0.466% 2.033% 6.987% 1.045% 1.746% 0.263%
-4 -5
10 - 10 0.361% 1.309% 2.346% 0.114% 0.183% 0.096%

Moderate (M2) 10-3 - 10-4 0.506% 2.304% 7.880% 1.706% 2.787% 0.393%

10-4 - 10-5 0.411% 1.362% 2.275% 0.196% 0.334% 0.123%

Fine (M3) 10-3 - 10-4 0.490% 2.292% 7.651% 3.023% 4.293% 0.640%
-4 -5
10 - 10 0.514% 1.712% 2.881% 0.289% 0.480% 0.140%

R2 Error Norms

GSE level U V W KE E SP

1E-03 1.35572703 0.946980 1.102498 1.423647 0.618053 0.992741


7 9 8 5 5

1E-04 1.3451 0.9467 1.0992 1.0844 0.3705 0.9442

1E-05 1.3369 0.9477 1.0932 1.1102 0.3861 0.9421


68

Geometry: Anatomical

Freestream velocity 0.4 m/s

Breathing velocity 1.81 m/s

L2 Error Norms

Mesh GSE Levels U V W KE E SP

Coarse (M1) 10-3 - 10-4 0.436% 1.716% 5.170% 2.834% 2.356% 0.129%
-4 -5
10 - 10 0.208% 0.772% 1.659% 1.002% 1.293% 0.062%

Moderate (M2) 10-3 - 10-4 0.429% 1.847% 6.711% 3.501% 2.367% 0.082%

10-4 - 10-5 0.360% 1.292% 2.515% 1.657% 1.799% 0.089%

Fine (M3) 10-3 - 10-4 0.494% 2.257% 7.975% 5.061% 3.313% 0.154%
-4 -5
10 - 10 0.503% 1.591% 2.774% 1.889% 2.009% 0.132%

R2 Error Norms

GSE level U V W KE E SP

1E-03 0.481196166 0.6299237 0.402598 0.4490886 0.1614825 0.2413776

1E-04 0.4862 0.6286 0.4123 0.4498 0.1631 0.2416

1E-05 0.4855 0.6280 0.4010 0.4471 0.1604 0.2383


69

Geometry: Full

Freestream velocity 0.2 m/s

Breathing velocity 12.11 m/s

L2 Error Norms

Mesh GSE U V W KE E SP
Levels

Coarse (M1) 10-3 - 10-4 0.454% 1.588% 7.847% 2.197% 5.207% 3.298%

10-4 - 10-5 0.369% 1.227% 3.101% 0.041% 0.007% 0.076%

Moderate (M2) 10-3 - 10-4 0.501% 1.824% 8.092% 1.523% 0.574% 1.173%
-4 -5
10 - 10 0.380% 1.250% 2.987% 0.094% 0.150% 0.098%

Fine (M3) 10-3 - 10-4 0.592% 2.455% 9.413% 3.612% 8.427% 4.617%

10-4 - 10-5 0.375% 1.271% 3.239% 0.054% 0.073% 0.061%

R2 Error Norms

GSE level U V W KE E SP

1E-03 0.5658 0.92796 0.95803 0.89649 0.88016 1.4941

1E-04 0.3790 0.8355 0.9575 0.5014 0.4503 1.3431

1E-05 0.3816 0.8352 0.9747 0.5004 0.4516 1.3318


70

Geometry: Full

Freestream velocity 0.4 m/s

Breathing velocity 1.81 m/s

L2 Error Norms

Mesh GSE U V W KE E SP
Levels

Coarse (M1) 10-3 - 10-4 0.503% 1.887% 7.170% 5.260% 9.587% 0.176%

10-4 - 10-5 0.255% 0.775% 1.842% 0.759% 0.304% 0.079%

Moderate (M2) 10-3 - 10-4 0.454% 1.650% 9.354% 7.887% 7.265% 0.409%
-4 -5
10 - 10 0.367% 1.179% 3.176% 1.440% 1.201% 0.083%

Fine (M3) 10-3 - 10-4 0.602% 2.491% 11.263% 9.090% 7.698% 0.406%

10-4 - 10-5 0.370% 1.219% 3.304% 1.121% 0.576% 0.062%

R2 Error Norms

GSE level U V W KE E SP

1E-03 1.06749 1.08772 0.78687 1.1398 1.07896 0.97943

1E-04 1.0521 1.0896 0.8215 1.1272 1.0896 0.9319

1E-05 1.0548 1.0888 0.8427 1.1265 1.0881 0.9390


71

Geometry: Full

Freestream velocity 0.4 m/s

Breathing velocity 12.11 m/s

L2 Error

Mesh GSE U V W KE E SP
Levels

Coarse (M1) 10-3 - 10-4 0.433% 1.589% 6.988% 1.103% 0.442% 0.177%

10-4 - 10-5 0.292% 0.981% 2.663% 0.078% 0.073% 0.045%

Moderate (M2) 10-3 - 10-4 3.635% 3.525% 3.263% 2.428% 0.000% 6.721%
-4 -5
10 - 10 1.025% 0.665% 0.127% 0.170% 0.000% 1.690%

Fine (M3) 10-3 - 10-4 0.897% 3.263% 12.021% 1.191% 0.680% 0.275%

10-4 - 10-5 0.000% 0.000% 0.000% 0.000% 0.000% 0.000%

R2 Error Norms

GSE level U V W KE E SP

1E-03 0.99837 1.00287 0.99231 1.00175 1.04715 0.99977

1E-04 1.0006 0.9997 1.0272 1.0018 1.0571 0.9997

1E-05 0.9995 1.0008 1.0119 1.0018 1.0563 0.9997


72

Geometry: Full

Freestream velocity 0.4 m/s

Breathing velocity 4.33 m/s

L2 Error Norms

Mesh GSE Levels U V W KE E SP

Coarse (M1) 10-3 - 10-4 0.466% 1.643% 6.028% 2.286% 3.985% 0.046%
-4 -5
10 - 10 0.168% 0.553% 1.488% 0.726% 1.059% 0.044%

Moderate (M2) 10-3 - 10-4 0.439% 1.581% 7.490% 5.478% 14.564% 0.173%

10-4 - 10-5 0.281% 0.917% 2.678% 1.422% 2.290% 0.002%

Fine (M3) 10-3 - 10-4 0.599% 2.134% 9.489% 6.128% 14.953% 0.222%
-4 -5
10 - 10 0.327% 1.080% 3.144% 1.387% 2.067% 0.039%

R2 Error Norms

GSE level U V W KE E SP TP

1E-03 0.874188625 0.7603224 0.7382383 0.6410194 0.5478818 0.9999724 #DIV/0!

1E-04 0.7569 0.6972 0.7800 0.8151 1.4912 0.9999 #DIV/0!

1E-05 0.8297 0.7341 0.8980 0.8380 1.6396 0.9999 #DIV/0!


73

Geometry: Truncated

Freestream velocity 0.2 m/s

Breathing velocity 12.11 m/s

L2 Error Norms

Mesh GSE U V W KE E SP
Levels

Coarse (M1) 10-3 - 10-4 0.229% 0.825% 2.167% 1.037% 1.222% 0.725%
-4 -5
10 - 10 0.075% 0.301% 0.735% 0.037% 0.046% 0.078%

Moderate (M2) 10-3 - 10-4 0.224% 0.798% 2.153% 0.982% 1.011% 0.590%

10-4 - 10-5 0.075% 0.296% 0.727% 0.079% 0.091% 0.115%

Fine (M3) 10-3 - 10-4 0.263% 0.932% 2.446% 1.183% 1.381% 0.760%
-4 -5
10 - 10 0.087% 0.341% 0.841% 0.034% 0.036% 0.076%

R2 Error Norms

GSE level U V W KE E SP

1E-03 1.10047 0.7778 1.30517 1.0887 1.40087 0.70324

1E-04 1.1090 0.7584 1.3641 1.1214 1.4116 0.6265

1E-05 1.1075 0.7539 1.3528 1.1220 1.4131 0.6246


74

Geometry: Truncated

Freestream velocity 0.2 m/s

Breathing velocity 1.81 m/s

L2 Error Norms

Mesh GSE U V W KE E SP
Levels

Coarse (M1) 10-3 - 10-4 0.299% 1.123% 3.020% 3.542% 6.388% 0.211%
-4 -5
10 - 10 0.050% 0.199% 0.461% 0.387% 0.219% 0.040%

Moderate (M2) 10-3 - 10-4 0.243% 0.920% 2.445% 2.886% 5.722% 0.168%

10-4 - 10-5 0.090% 0.347% 0.852% 0.565% 0.221% 0.055%

Fine (M3) 10-3 - 10-4 0.286% 1.059% 2.865% 3.539% 7.067% 0.164%
-4 -5
10 - 10 0.105% 0.402% 0.991% 0.549% 0.210% 0.060%

R2 Error Norms

GSE level U V W KE E SP

1E-03 0.23895 0.14944 0.4691 0.60325 0.22238 0.29976

1E-04 0.2251 0.1452 0.4688 0.6509 0.3450 0.2967

1E-05 0.2221 0.1439 0.4725 0.6486 0.3477 0.2947


75

Geometry: Truncated

Freestream velocity 0.4 m/s

Breathing velocity 12.11 m/s

L2 Error Norms

Mesh GSE Levels U V W KE E SP

Coarse (M1) 10-3 - 10-4 0.260% 1.039% 2.569% 0.415% 0.307% 0.139%
-4 -5
10 - 10 0.049% 0.194% 0.463% 0.054% 0.030% 0.031%

Moderate (M2) 10-3 - 10-4 0.255% 1.030% 2.686% 0.410% 0.500% 1.689%

10-4 - 10-5 0.057% 0.228% 0.548% 0.052% 0.081% 0.028%

Fine (M3) 10-3 - 10-4 0.220% 0.861% 2.087% 0.226% 0.301% 0.130%
-4 -5
10 - 10 0.069% 0.274% 0.654% 0.028% 0.006% 0.038%

R2 Error Norms

GSE level U V W KE E SP

1E-03 1.061491529 0.7604667 1.4542463 0.9734411 1.3572143 0.8251998

1E-04 1.0525 0.7632 1.2793 1.1144 1.4346 0.5754

1E-05 1.0511 0.7629 1.2687 1.1012 1.4219 0.5692


76

Geometry: Truncated

Freestream velocity 0.4 m/s

Breathing velocity 1.81 m/s

L2 Error Norms

Mesh GSE Levels U V W KE E SP

Coarse (M1) 10-3 - 10-4 0.311% 1.143% 2.909% 1.799% 3.629% 0.166%
-4 -5
10 - 10 0.045% 0.170% 0.415% 0.439% 0.548% 0.029%

Moderate (M2) 10-3 - 10-4 0.256% 0.966% 2.463% 1.553% 2.601% 0.128%

10-4 - 10-5 0.069% 0.263% 0.660% 0.696% 0.961% 0.036%

Fine (M3) 10-3 - 10-4 0.229% 0.862% 2.159% 1.067% 1.346% 0.109%
-4 -5
10 - 10 0.080% 0.303% 0.762% 0.637% 1.024% 0.038%

R2 Error Norms

GSE level U V W KE E SP

1E-03 0.279149905 0.1486746 0.579351 1.0696241 0.5189611 0.4468577

1E-04 0.2787 0.1495 0.5736 1.0738 0.5127 0.4472

1E-05 0.2740 0.1487 0.5753 1.0757 0.5196 0.4437


77

APPENDIX E

VELOCITY REGRESSION PLOTS

Matched by Breathing Rate, Freestream Velocity and Location

0.25
y = 0.7061x + 0.0108
0.2 R² = 0.3928
Vertical Velocity for Anatomical Model

0.15

0.1

0.05

‐0.05

‐0.1

‐0.15
‐0.15 ‐0.05 0.05 0.15 0.25 0.35
Vertical Velocity for Full Model
78

0.25

0.2 y = 0.9161x + 0.0085
R² = 0.6835

Vertical Velocity for Full Model
0.15

0.1

0.05

‐0.05

‐0.1

‐0.15
‐0.15 ‐0.05 0.05 0.15 0.25
Vertical Velocity for Truncated Model

0.25
y = 0.4207x + 0.02
0.2 R² = 0.1135
Vertical Velocioty For Anatomical Model

0.15

0.1

0.05

‐0.05

‐0.1

‐0.15
‐0.15 ‐0.05 0.05 0.15 0.25
Vertical Velocity for Truncated Model
79

0.2 m/s, Heavy Breathing Conditions

0.15

Vertical (Z) Velocity for Anatomical Model
0.1

0.05

‐0.05
y = 0.5269x + 0.0102
R² = 0.2347

‐0.1
‐0.1 ‐0.05 0 0.05 0.1 0.15
Vertical (Z) Velocity for Full Model
80

0.15

y = 0.9188x + 0.0061
R² = 0.7067
0.1

Vertical (Z) Velocity for Full Model
0.05

‐0.05

‐0.1

‐0.15
‐0.15 ‐0.1 ‐0.05 0 0.05 0.1 0.15
Vertical (Z) Velocity for Truncated Model

0.12

0.1
Vertical (Z) Velocity for Anatomical Model

0.08

0.06

0.04

0.02

‐0.02

‐0.04 y = 0.1921x + 0.0161
R² = 0.0261
‐0.06

‐0.08

‐0.1
‐0.15 ‐0.1 ‐0.05 0 0.05 0.1 0.15
Vertical (Z) Velocity for Truncated Model
81

0.2 m/s, At-Rest Breathing Rate

0.12

y = 0.8204x + 0.0062
Vertical (Z) Velocity for Anatomical Model
0.1 R² = 0.5019

0.08

0.06

0.04

0.02

‐0.02
‐0.02 0.03 0.08 0.13
Vertical (Z) Velocity for Full Model
82

0.12

0.1 y = 0.5783x + 0.009
R² = 0.2963

Vertical (Z) Velocity for Full Model
0.08

0.06

0.04

0.02

‐0.02

‐0.04
‐0.08 ‐0.03 0.02 0.07 0.12
Vertical (Z) Velocity for Tuncated Model

0.12
y = 0.2552x + 0.0049
0.1
R² = 0.0986
Vertical (Z) Velocity for Anatomical Model

0.08

0.06

0.04

0.02

‐0.02

‐0.04

‐0.06

‐0.08
‐0.08 ‐0.03 0.02 0.07 0.12
Vertical (Z) Velocity for Truncated Model
83

0.4 m/s, Heavy Breathing Rate

0.25

Vertical (Z) Velocity for Anatomical Model
0.2

0.15

0.1

0.05

‐0.05 y = 0.6356x + 0.0157
R² = 0.3242
‐0.1
‐0.1 0 0.1 0.2 0.3
Vertical (Z) Velocity for Full Model
84

0.3

0.2

Vertical (Z) Velocity for Full Model
0.1

y = 0.9443x + 0.0105
R² = 0.7363
‐0.1

‐0.2
‐0.2 ‐0.1 0 0.1 0.2 0.3
Vertical (Z) Velocity for Truncated Model

0.25

0.2
Vertical (Z) Velocity for Anatomical Mocel

0.15

0.1

0.05

‐0.05 y = 0.347x + 0.0272
R² = 0.0798
‐0.1

‐0.15
‐0.15 ‐0.05 0.05 0.15 0.25
Vertical (Z) Velocity for Truncated Model
85

0.4 m/s, At-Rest Breathing Rate

0.25

y = 0.7262x + 0.0134
Vertical (Z) Velociy for Anatomical Model 0.2
R² = 0.4069

0.15

0.1

0.05

‐0.05
‐0.05 0 0.05 0.1 0.15 0.2 0.25
Vertical (Z) Velocity for Full Model
86

0.3

0.2

Vertical (Z) Velocity for Full Model
0.1

y = 0.9196x + 0.0108
‐0.1 R² = 0.711

‐0.2
‐0.2 ‐0.1 0 0.1 0.2 0.3
Vertical (Z) Velocity for Truncated Model

0.25

0.2
Vertical (Z) Velocity for Anatomical Model

0.15

0.1

0.05

‐0.05
y = 0.4461x + 0.0254
R² = 0.1291
‐0.1

‐0.15
‐0.15 ‐0.05 0.05 0.15 0.25
Vertical (Z) Velocity for Truncated Model
87

APPENDIX F

CRITICAL AREA REGRESSION PLOTS

0.8000
y(truncated critical area) = 1.03x(anatomical 
Critical Area Position for Truncated Geometry, m

0.7000 critical area) ‐ 0.03


R² = 0.99
0.6000

0.5000

0.4000

0.3000

0.2000

0.1000

0.0000

‐0.1000

‐0.2000
‐0.2000 0.0000 0.2000 0.4000 0.6000 0.8000
Critical Area Position for Anatomical Geometry, m
88

0.8000

Critical Area Position for Anatomical Torso Geometry, 
y(anatomical critical area position) = 
0.7000 1.02x(full critical area position) ‐ 0.01
R² = 0.99
0.6000

0.5000

0.4000

0.3000
m

0.2000

0.1000

0.0000

‐0.1000

‐0.2000
‐0.2000 0.0000 0.2000 0.4000 0.6000 0.8000
Critical Area Position for Full Torso Geometry, m

0.8000
Critical Area Position for Truncated Torso Geometry, 

y(truncated critical area position)= 
0.7000 0.99x(anatomical critical area position) + 
0.01
0.6000 R² = 0.99

0.5000

0.4000

0.3000
m

0.2000

0.1000

0.0000

‐0.1000

‐0.2000
‐0.2000 0.0000 0.2000 0.4000 0.6000 0.8000
Critical Area Position for Full Torso Geometry, m
89

APPENDIX G

ASPIRATION EFFICIENCY REGRESSION PLOTS

No Bounce Conditions

120.0%

y = 0.9763x + 0.0078
100.0%
R² = 0.996
Full Model Aspiration Efficiency, %

80.0%

60.0%

40.0%

20.0%

0.0%
0.0% 20.0% 40.0% 60.0% 80.0% 100.0% 120.0%
Truncated Model Aspiration Efficiency, %
90

120.0%

Anatomical Model Aspiration Efficiency, %
100.0% y = 0.9863x + 0.0243
R² = 0.9952

80.0%

60.0%

40.0%

20.0%

0.0%
0.0% 20.0% 40.0% 60.0% 80.0% 100.0% 120.0%
Full Model Aspiration Efficiency, %

120.0%
Anatomical Model Aspiration Efficiency, %

100.0% y = 1.0299x ‐ 0.0276


R² = 0.9923

80.0%

60.0%

40.0%

20.0%

0.0%
0.0% 20.0% 40.0% 60.0% 80.0% 100.0% 120.0%
Truncated Model Aspiration Efficiency, %
91

Bounce Conditions

120.0%
y(full model) = 0.7894x(truncated) + 0.1797
R² = 0.8802
Aspiration Efficiency Full Model, % 100.0%

80.0%

60.0%

40.0%

20.0%

0.0%
0.0% 20.0% 40.0% 60.0% 80.0% 100.0% 120.0%
Aspiration Efficiency Truncated Model, %
92

120.0%
y(Anatomical) = 1.0662x(Full) ‐ 0.0628
R² = 0.8718

Aspiration Efficiency Anatomical Model, %
100.0%

80.0%

60.0%

40.0%

20.0%

0.0%
0.0% 20.0% 40.0% 60.0% 80.0% 100.0% 120.0%
Aspiration Efficiency Full Model, %

120.0%
y(Anatomical) = 0.9471x(Truncated) + 0.0364
R² = 0.9717
Aspiration Efficiency Anatomical Model, %

100.0%

80.0%

60.0%

40.0%

20.0%

0.0%
0.0% 20.0% 40.0% 60.0% 80.0% 100.0% 120.0%
Aspiration Efficiency Truncated Model, %
93

REFERENCES

Aitken, R. J., Baldwin, P. E. J., Beaumont, G. C., Kenny, L. C., & Maynard, A. D. (1999).
Aerosol inhalablity in low air movement environments. Journal of Aerosol Science, 30(5),
613-626.

Anthony, T. R. (2010). Contribution of facial feature dimensions and velocity parameters on


particle inhalability. Annals of Occupational Hygiene, 54(6), 710-725.

Anthony, T. R., & Flynn, M. R. (March 2006). CFD model for a 3-D inhaling mannequin:
Verification and validation. Annals of Occupational Hygiene, 50(2), 157-173.

Anthony, T. R., & Flynn, M. R. (2006). Computational fluid dynamics investigation of particle
inhalability. Journal of Aerosol Science, 37(6), 750-765.

Anthony, T. R., Flynn, M. R., & Eisner, A. (2005). Evaluation of facial features on particle
inhalation. Annals of Occupational Hygiene, 49(2), 179-193.

Armbruster, L., & Breuer, H. (1982). Investigations into defining inhalable dust. Annals of
Occupational Hygiene, 26(1), 21-32.

Baldwin, P. E. J., & Maynard, A. D. (1998). A survey of wind speeds in indoor workplaces.
Annals of Occupational Hygiene, 42(5), 303-313.

Bjørn, E., & Nielsen, P. V. (2002). Dispersal of exhaled air and personal exposure in
displacement ventilated rooms. Indoor Air, 12(3), 147-164.

Brohus, H. (1997). Personal exposure to contaminant sources in ventilated rooms, Ph.D. Thesis,
Aalborg University, Denmark.

Brohus, H., and Nielsen, P.V. (1996). CFD Models of Persons Evaluated By Full-Scale Wind
Channel Experiments. Proccedings of RoomVent, Vol 2, 137-144.

Chang, S.K.W and Gonzalez, R.R. (1993). Air Velocity Profiles around the human body.
ASHRAE Transactions, 99(1), 405-458.

Chung, I. P., & Dunn-Rankin, D. (1992). Numerical simulation of two-dimensional blunt body
sampling in viscous flow. Journal of Aerosol Science, 23(3), 217-232.

Dunnett, S. J. (1997). A numerical study of the flow field in the vicinity of a bluff body with
aspiration oriented to the flow. Atmospheric Environment, 31(22), 3745-3752.

Dunnett, S. J. (1999). An analytical investigation into the nature of the airflow near a spherical
bluff body with suction. Journal of Aerosol Science, 30(2), 163-171.

Dunnett, S. J., & Ingham, D. B. (1986). A mathematical theory to two-dimensional blunt body
sampling. Journal of Aerosol Science, 17(5), 839-853.

Dunnett, S. J., & Ingham, D. B. (1987). The effects of finite reynolds number on the aspiration of
particles into a bulky sampling head. Journal of Aerosol Science, 18(5), 553-561.

Dunnet, S. J., & Ingham, D. B. (1988). The aspiration of non-spherical particles into a bulky
sampling head. Annals of Occupational Hygiene, 32(2), 221-235.
94

Erdal, S., & Esment, N. A. (1995). Human head model as an aerosol sampler: Calculation of
aspiration efficiencies for coarse particles using an idealized human head model facing the
wind. Journal of Aerosol Science, 26(2), 253-272.

Hayashi, T., Ishizu, Y., Kato, S., & Murakami, S. (2002). CFD analysis on characteristics of
contaminated indoor air ventilation and its application in the evaluation of the effects of
contaminant inhalation by a human occupant. Building and Environment, 37(3), 219-230.

Hsu, D., & Swift, D. L. (1999). The measurements of human inhalability of ultralarge aerosols in
calm air using mannikins. Journal of Aerosol Science, 30(10), 1331-1343.

Hyun, S., & Kleinstreuer, C. (2001). Numerical simulation of mixed convection heat and mass
transfer in a human inhalation test chamber. International Journal of Heat and Mass
Transfer, 44(12), 2247-2260.

Ingham, D. B. (1981). The entrance of airborne particles into a blunt sampling head. Journal of
Aerosol Science, 12(6), 541-549.

Ingham, D. B., Hildyard, L. T., & Hildyard, M. L. (1991). On the critical stokes' number for
particle transport in potential and viscous flows near bluff bodies. Journal of Aerosol
Science, 21(7), 935-946.

Kennedy, N. J., & Hinds, W. C. (2002). Inhalability of large solid particles. Journal of Aerosol
Science, 33(2), 237-255.

Kennedy, Nola J., Tatyan, Karina and Hinds, William C. (2001). Comparison of a Simplified and
Full-Sized Mannequin for the Evaluation of Inhalable Sampler Performance. Aerosol Science
and Technology, 35(1), 564-568.

King Se, C. M., Inthavong, K., & Tu, J. (2010). Inhalability of micron particles through the nose
and mouth. Inhalation Toxicology, 22(4), 287-300.

Murakami, S., Kato, S., & Zeng, J. (2000). Combined simulation of airflow, radiation and
moisture transport for heat release from a human body. Building and Environment, 35(6),
489-500.

Murakami, S., Zeng, J., & Hayashi, T. (1999). CFD analysis of wind environment around a
human body. Journal of Wind Engineering and Industrial Aerodynamics, 83(1-3), 393-408.

Ogden, T. L., & Birkett, J. L. (1977). The human head as a dust sampler. in:W.H. Halton (Ed.),
Inhaled particles IV: Proceedings of an international symposium organized by the British
Occupational Hygiene Society (pp. 93–105). Oxford: Pergamon Press, ISBN 0080205607.
Ogden, T. L., Birkett, J. L., & Institute of Occupational Medicine Edinburgh. (1978). An
inhalable dust sampler, for measuring the hazard from total airborne particulate. Annals of
Occupational Hygiene, 21(1), 41-50.

Rodes, C.E., Kamens, R.M., & Wirner, R.W. (1995). Experimental consideration for the study of
contaminant dispersal near the body. American Industrial Hygiene Association, Journal, 56,
535-585.
95

Schmees, D. K., Wu, Y., & Vincent, J. H. (2008). Visualization of the airflow around a life-
sized, heated, breathing mannequin at ultralow windspeeds. Annals of Occupational Hygiene,
52(5), 351-360.

Sideroff, C.N., and Dang, T.Q. (2005). Validation of CFD for the Flow around a computer
simulated person in a mixing ventilated room. Indoor Air, Beijing China.

Sleeth, D. K. and Vincent, J.H. (2009). Inhalability for aerosols at ultra-low windspeeds, Journal
of Physics: Conference Series, 151(1) 012062.

Smith, J. P., & Bird, A. J. (2002). Relationship of sampling efficiency for manikin-mounted
personal samplers to efficiency measurements made independent of manikin. Journal of
Aerosol Science, 33(9), 1235-1259.

Sørensen, D. N., & Voigt, L. K. (2003). Modelling flow and heat transfer around a seated human
body by computational fluid dynamics. Building and Environment, 38(6), 753-762.

Soderholm, S.C. (1989). Proposed International Conventions for Particle Size-Selective


Sampling. Annals of Occupational Hygiene, 33(3), 301-320.

Sreenath, A., Ramachandran, G., & Vincent, J. H. (1997). Experimental investigations into the
nature of airflows near bluff bodies with aspiration, with implications to aerosol sampling.
Atmospheric Environment, 31(15), 2349-2359.

Stern F, Wilson RV, Coleman HW et al. (2001) Comprehensive approach to verification and
validation of CFD simulations—part 1: methodology and procedures. J Fluids Engin;
123:793–802.

Topp, C., Hessellholt, P., Trier, M. R. T., and Nielsen, P.V. (2003). Influence of Thermal
Manikins on Concentration Distribution and Personal Exposure. Proceedings of healthy
buildings, Singapore.

Vincent, J. H.,&Armbruster, L. (1981). On the quantitative definition of the inhalability of


airborne dust. Annals of Occupational Hygiene, 24, 245–248.

Vincent, J.H., Hustson D., and Mark, D., (1982) The nature of airflow near the inlets of blunt
dust sampling probes. Atmospheric Environment, 16, 1243 – 1249.

Vincent, J. H., & Mark, D. (1982). Applications of blunt sampler theory to the definition and
measurement of inhalable dust. Annals of Occupational Hygiene, 26(1), 3–19.

Vincent, J.H., Wood, J.D., Birkett, J.L. and Gibson, H. (1979) Institute of Occupational
Medicine (Edinburgh), Technical Memorandum TM/79/18.

Xing, H., Hatton, A., and Awbi, H.B. (2001). A study of the air quality in the breathing zone in a
room with displacement ventilation. Building and Environment 36, 809-820.

Zhu, S., Kato, S., Murakami, S., & Hayashi, T. (2005). Study on inhalation region by means of
CFD analysis and experiment. Building and Environment, 40(10), 1329-1336.

Potrebbero piacerti anche