mechanical characterization of red blood cells using ...€¦ · relationship between rbc...

152
Mechanical Characterization of Red Blood Cells Using Microfluidic Devices by Zhensong Xu A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Mechanical and Industrial Engineering University of Toronto © Copyright by Zhensong Xu 2018

Upload: others

Post on 06-Jul-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

Mechanical Characterization of Red Blood Cells Using Microfluidic Devices

by

Zhensong Xu

A thesis submitted in conformity with the requirements

for the degree of Doctor of Philosophy

Department of Mechanical and Industrial Engineering

University of Toronto

© Copyright by Zhensong Xu 2018

Page 2: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

ii

Abstract

Mechanical Characterization of Red Blood Cells Using Microfluidic Devices

Zhensong Xu

Doctor of Philosophy

Department of Mechanical and Industrial Engineering

University of Toronto

2018

The mechanical properties of red blood cells (RBCs) have been proven to play important roles in

the regulation of various biological activities and have shown clinical significance. During the

past few decades, a number of research studies have been conducted to understand the

relationship between RBC mechanical properties and diseases. The work described in this thesis

focuses on characterizing mechanical properties of RBCs from sickle cell trait (SCT) carriers and

the degradation/recovery of stored RBCs in transfusion medicine.

The mechanical properties and changes of SCT RBCs under deoxygenated and acidic

environments, two typical conditions present in the circulation of athletes undertaking strenuous

exercise, were measured. The results revealed that SCT RBCs are inherently stiffer than RBCs

from non-SCT healthy subjects, and a lower pH further stiffens the SCT cells. Furthermore, at

Page 3: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

iii

both normal and low pH levels, deoxygenation was found to cause no change of SCT RBCs

stiffness.

In transfusion medicine, compromised RBC deformability can lower the transfusion efficiency or

intensify transfusion complications. This thesis reports microfluidic mechanical measurement of

stored RBCs under the physiological deformation mode. The results revealed that the effective

stiffness of RBCs increases over the storage process. RBCs stored for one week already started

to show significantly higher stiffness than fresh RBCs, and stored RBC stiffness degraded faster

in the last three weeks than in the first three weeks.

Although the stiffness of stored RBCs degrades over the storage process, whether the stiffness of

stored RBCs can be reversed after transfusion remains unknown. A microfluidic platform was

used in this thesis to measure the evolution of RBC stiffness recovery under in vivo-like

conditions. RBCs stored up to 6 weeks (42 days) in the blood bank were measured, revealing that

the degraded stiffness of RBCs over the storage process can be recovered in human serum within

120 minutes. However, the recovered stiffness of older RBCs (stored for 4-6 weeks) is still 1.6 –

2.1 times higher than that of fresh RBCs. Furthermore, ATP, which provides energy to keep

RBC membrane mechanically functional, was also measured. The ATP concentration in stored

RBCs was found to increase in human serum.

Page 4: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

iv

Acknowledgements  

The past four years I spent in my Ph.D. study have been the most rewarding years in my life. I

would like to express my appreciation to all those who provided me the possibility to complete

this degree. My deep gratitude goes first to my advisor, Prof. Yu Sun, for all his advice and

guidance throughout the past four years at the University of Toronto. I have learned a great deal

from Prof. Sun’s energy and critical thinking, which helped me shape this thesis. I often

encountered difficulties to address some of the comments from Prof. Sun during my research,

however, it was those comments that helped me learn fast and significantly improved my

following work. Thinking thoroughly is not only applied in my research, but also expands out of

academia, which allows me to efficiently create solutions for questions in different aspects. Prof.

Sun’s support is also what I am much appreciated. His supportive and knowledgeable comments

always recharge my energy in research, and he encouraged me to attend international

conferences where my presenting skills were improved and my knowledge was broadened. Prof.

Sun’s constant encouragement and tremendous help during my PhD study always inspired and

motivated me to be passionate for research.

My thesis is challenging, and I am glad to have the support from my great collaborators. I would

like to express my great appreciation to Dr. Chen Wang for his close collaboration and valuable

guidance and discussions. Whenever I needed samples or discussions, Dr. Wang was always

willing to help. I thank Prof. Axel Guenther for his knowledgeable advice in my committee

meetings which greatly improved my research. I also thank Prof. Xinyu Liu and Prof. Francis

Lin for serving on my defense committees.

Page 5: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

v

I am glad to be surrounded by many talented people during my graduate study. They did not only

provide me enormous help with my graduate study, they are also my valuable friends who I feel

fortunate to have. I am especially grateful to Yi Zheng, Yuan Wei, Xian Wang and Wenkun who

spent endless hours working with me. I thank Jun Wen for being a fabulous lab mate and friend

for years, who showed me the beauty of Toronto when I first came to this lovable city, and I

enjoyed the fun conversations we had during the past years; Devin Luu for being a reliable friend,

who gave me a lot of support at sin & redemption and helped me edit this thesis; Changhong for

working out together and giving me advice in research. I would also like to thank all past and

present members of the Advanced Micro and Nanosystems Laboratory for all their helpful

discussions and encouragements throughout the years.

Finally but not least, I wish to thank my parents for their encouragement and love over the years.

They always respect my interest and support every decision I make in my life, which gives me

the courage to move forward without hesitation. They are also my best friends who share my joy

and sorrow. I am far away from home but never far away from my parents.

 

   

Page 6: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

vi

Table of Contents

Acknowledgements ...................................................................................................................................... iv

Table of Contents ......................................................................................................................................... vi

List of Figures ............................................................................................................................................ viii

1. Introduction ........................................................................................................................................... 1

1.1. RBC introduction .......................................................................................................................... 1

1.1.1. Membrane lipids ................................................................................................................... 3

1.1.2. Membrane proteins ............................................................................................................... 4

1.1.3. Skeleton proteins ................................................................................................................... 5

1.2. RBC stiffness measurement .......................................................................................................... 7

1.2.1. Filtration ................................................................................................................................ 7

1.2.2. Micropipette aspiration ....................................................................................................... 10

1.2.3. Atomic force microscopy .................................................................................................... 12

1.2.4. Optical tweezers .................................................................................................................. 12

1.2.5. Magnetic twisting ................................................................................................................ 15

1.2.6. Quantitative phase imaging ................................................................................................. 17

1.2.7. Microfluidic measurement .................................................................................................. 18

1.3. Pathophysiological conditions and RBC stiffness ...................................................................... 23

1.3.1. Diseases ............................................................................................................................... 23

1.3.2. Malaria ................................................................................................................................ 24

1.3.3. Sickle cell disease ............................................................................................................... 25

1.3.4. Sickle cell trait .................................................................................................................... 27

1.3.5. Stored RBCs ........................................................................................................................ 28

1.4. Research objectives ..................................................................................................................... 31

2. Stiffening of sickle cell trait red blood cells under simulated strenuous exercise conditions ............. 32

2.1. Introduction ................................................................................................................................. 32

2.2. Methods ....................................................................................................................................... 33

2.2.1. Blood specimens ................................................................................................................. 33

2.2.2. Device and measurement .................................................................................................... 34

2.3. Results and discussion ................................................................................................................ 37

2.3.1. Determination of RBC shear modulus ................................................................................ 37

2.3.2. SCT RBCs become stiffened under lower pH .................................................................... 38

Page 7: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

vii

2.3.3. Hypoxia does not induce SCT RBC stiffening ................................................................... 42

2.4. Discussion ................................................................................................................................... 45

2.5. Conclusion .................................................................................................................................. 46

3. Stiffness increase of red blood cells during storage ............................................................................ 48

3.1. Introduction ................................................................................................................................. 48

3.2. Device fabrication ....................................................................................................................... 50

3.3. Methods ....................................................................................................................................... 52

3.4. Experimental Results .................................................................................................................. 57

3.4.1. Effective stiffness is not velocity dependent under low velocities. .................................... 61

3.4.2. Temperature effect on RBC effective stiffness ................................................................... 63

3.4.3. RBC effective stiffness increases during storage. ............................................................... 65

3.5. Discussion ................................................................................................................................... 67

3.6. Conclusion: ................................................................................................................................. 71

4. Stiffness and ATP Recovery of Stored Red Blood Cells in Human Serum ........................................ 73

4.1. Introduction ................................................................................................................................. 73

4.2. Methods ....................................................................................................................................... 75

4.3. Results ......................................................................................................................................... 78

4.3.1. Stiffness recovery of stored RBCs in human serum ........................................................... 78

4.3.2. Fresher RBCs reaching steady-state shear modulus faster .................................................. 81

4.3.3. Temperature effect on RBC stiffness recovery ................................................................... 83

4.3.4. Shape recovery of stored RBCs in human serum ................................................................ 85

4.3.5. ATP concentration recovery of stored RBCs in human serum ........................................... 87

4.4. Discussion ................................................................................................................................... 90

4.5. Conclusion .................................................................................................................................. 95

5. Conclusions ......................................................................................................................................... 96

5.1. Contributions ............................................................................................................................... 96

5.2. Future directions ......................................................................................................................... 99

6. Bibliography: .................................................................................................................................... 102 

 

Page 8: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

viii

List of Figures Figure 1.1 (a) RBC takes oxygen from lungs and releases the oxygen to tissues. (b) RBC’s thickness at

the thickest point is 2 – 2.5 µm and the minimum thickness in the center is 0.8 – 1 µm. RBC has a

dimeter of 6 – 8 µm. .............................................................................................................................. 2 

Figure 1.2 Current model of the RBC membrane. A phospholipid bilayer is tethered to the spectrin

network (spectrin α and spectrin β proteins) via a number of proteins such as band 3, ankyrin, protein

4.1, and Glycophorin. Spectrin α and spectrin β form the structure underneath cell membrane and

provide mechanical support. Most of the known proteins are shown [26]. .......................................... 6 

Figure 1.3 (a) An array of parallel micro channels. (b-d) Artificial microvascular networks mimicking

vessels can be used to measure the deformability of red blood cells. ................................................... 9 

Figure 1.4 (a) Different micropipette aspiration methods used to measure the mechanical properties of red

blood cell (RBC) membranes. (b)Microfluidic pipette aspiration of RBCs using a funnel channel

chain. ................................................................................................................................................... 11 

Figure 1.5 (a) Optical tweezers measuring the RBC deformability with focused laser beams that transfer

linear momentum or angular momentum of light. To measure the shear modulus of the RBCs, two

microbeads are attached to the opposite sides of a RBC and force was applied by the optical tweezers.

(b) Experimental observations of RBC deformability measurement using optical tweezers, and

simulation results. (c) 3D image of an RBC measure by AFM. ......................................................... 14 

Figure 1.6 RBCs with magnetic beads bound to the surface. SEM image of an RBC with a bead on the

surface. C) Magnetic field is applied onto the bead to generate a torque to deform the RBC. d)

Mechanical simulation to predict the deformation by the force applied by the magnetic bead.

Permission is not required (thesis/dissertation)................................................................................... 16 

Figure 1.7 A network of microfluidic channels to measure RBCs deformability. The transit time of the

individual cells are recorded using a high-speed camera. (b) A microfluidic system for mechanical

Page 9: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

ix

characterization of RBCs. The transit time is obtained from electrical impedance signal captured

when RBCs are passing through the constriction channel. ................................................................. 19 

Figure 1.8 (a)Hydrodynamic RBC deforming microfluidic device. Cells are focused to the center lines of

the channels and deformed by fluid. Cell deformation is measured by analyzing images recorded by

a high-speed camera. (b)Laser diffraction technique incorporating microfluidic rheometer to measure

RBC deformability with elongation index defined. ............................................................................ 22 

Figure 2.1 (a) Schematic of the microfluidic device for RBC mechanical property measurement. RBCs in

solution of different pH levels are loaded into the middle channel and adhere onto the glass channel

bottom. Oxygen level is accurately varied by pumping air or nitrogen into the two side channels.

Schematic illustration of the deformation of a cell element. (b) RBCs are deformed under shear stress

(1.8 kPa). After the release of shear stress, RBCs recover to their original shape .............................. 36 

Figure 2.2 (a) Shear modulus of SCT RBCs under physiological pH level 7.35 (blue) and acid pH level of

6.85 (red). Error bars represent standard error of the mean value. All samples (n=30-70 for each

sample) show a significant difference between different pH levels (*p < 0.05). (b) Shear modulus of

normal RBCs does not reveal significant difference under different pH levels (p value > 0.05; n=30-

70 for each sample). Error bars represent standard deviation. c) Box plot showing the summarized

shear modulus of RBCs from the 7 tested SCT samples and the 7 tested normal samples under

different pH conditions(**p=5.5 × 10-15,***p=2.2 × 10-29,****p=1.9 × 10-82; n=200-300). ........ 42 

Figure 2.3 At physiological pH 7.35 (a) and low pH 6.85 (b), normal RBCs elastic modulus did not

change significantly when the channel was deoxygenated (*p>0.75). (c)(d) For SCT RBCs,

although shear modulus became higher at pH 6.85 than at pH 7.35, deoxygenation did not induce

further increase (*p>0.75). (e) Summarized shear modulus of SCT RBCs under different oxygen

levels, and no difference was observed. Error bars represent standard deviation. .............................. 44 

Page 10: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

x

Figure 3.1(a) Microfluidic device with a constriction channel where RBCs are bent by drag force. (b)

RBC’s shape change over storage showing that up to 12% cells change their shapes from discocytes

to echinocytes. ..................................................................................................................................... 54 

Figure 3.2 (a) Simulation shows that the highest stresses occur in RBC membrane regions closest to the

channel walls. (b) At a flow velocity of 0.02 m/s, shear stress becomes significantly higher than 300

Pa for a channel width smaller than 10 µm, which can lead to RBC lysis. When channel width is

larger than 14 µm, shear stress is too low to deform RBCs. ............................................................... 56 

Figure 3.3 Schematic force diagram. AFM experiment and force displacement curve. Contact width

estimation by applying a force of 150 pN. .......................................................................................... 60 

Figure 3.4 Experimentally measured deformation index DI =L/d and effective stiffness values of RBCs

from a fresh sample. (a) DI significantly decreases at higher flow velocities (*p < 0.001). (b) In

comparison, effective stiffness of RBCs is not flow velocity dependent. Error bars represent the

standard deviation. .............................................................................................................................. 63 

Figure 3.5 Effective stiffness of RBCs at different temperatures. RBCs under 4°C show higher effective

stiffness. RBC stiffness measured at 25°C shows no significant difference from that measured at

37°C (*p < 0.001). Error bars represent the standard deviation. ......................................................... 65 

Figure 3.6 Effective stiffness of RBCs at different storage time points. RBCs stored for one week already

started to show significantly higher stiffness than fresh RBCs. Stored RBC stiffness degraded faster

in the last three weeks than in the first three weeks(*p < 0.001 and **p < 0.0001). Error bars

represent the standard deviation. ......................................................................................................... 66 

Figure 3.7 (a) RBC membrane skeletal network. The affinities and interactions of key proteins are

regulated by biochemical parameters such as SNO and ATP, leading to RBC membrane stiffness

changes. (b) SNO (blue) and ATP (red) degradation during RBC storage. Data shown here are from

[30][31][32]. Approximately 90% SNO is depleted during the first week of storage, and ATP

Page 11: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

xi

remains unchanged during the first three weeks of storage and starts to decrease significantly in

Week 4. ............................................................................................................................................... 69 

Figure 4.1 (a) RBCs strongly adhered to glass substrate after settling for 15 minutes. RBCs in

microchannel were perfused with 37 °C human serum for 120 minutes. (b) RBCs were deformed by

shear flow (flow rate of 10 µL/min). Images of RBC deformation were recorded, and the shear

modulus of each RBC was quantified. ................................................................................................ 77 

Figure 4.2 Stiffness of RBCs perfused for 120 minutes in human serum vs. end-point measured stiffness

of RBCs incubated in human serum in an incubator without perfusion (n=1300-3000 RBCs for each

condition). No difference was observed between the perfusion group and incubating group (*p>0.05),

showing that shear stress from perfusion did not induce bias in RBC stiffness recovery. ................. 78 

Figure 4.3 (a) Shear modulus evolution of two-week old RBCs. PBS perfusion (blue line) did not cause

RBCs to recover their shear modulus. With perfusion of 37 °C human serum (red line), RBCs shear

modulus remained unchanged at 4.3 µN/m for 60 minutes; however, between 70-80 minutes, the

shear modulus value continuously decreased. By 90 minutes, stead-state was reached, and shear

modulus became 2.7 µN/m. (b) Steady-state shear modulus values after 120-min perfusion. With

human serum perfusion, RBCs stored for one week or two weeks were able to recover their shear

modulus close to the level of fresh RBCs. Older RBCs (four-six weeks) revealed limited capability

of stiffness recovery. *p <0.05&**p<0.001. Error bars represent the standard deviation. For each

condition, n= 1800-3000 RBCs. ......................................................................................................... 80 

Figure 4.4 Stiffness of RBCs perfused up to 120 minutes in human serum vs. perfused up to 8 hours

(n=790-3000 RBCs for each condition). No difference was observed (*p>0.05), showing that longer

perfusion did not induce further recovery of shear modulus. ............................................................. 82 

Figure 4.5 Steady-state shear modulus of stored RBCs that were PBS or serum perfused at 25 °C or 37 °C

(n=1260-3000 RBCs for each condition). The results show that 25 °C and 37 °C did not cause a

statistically significant difference in the stiffness of PBS-perfused or serum-perfused RBCs. .......... 84 

Page 12: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

xii

Figure 4.6 (a) Higher percentages of one-week and three-week old RBCs, compared to five-week old

RBCs, reached their steady-state shear modulus by 80 minutes. b-d) Compared to 25 °C, 37 °C

accelerated the stiffness recovery process for stored RBCs in human serum. Blue and orange lines

are the second degree polynomial fitting trend lines. Error bars represent the standard deviation. .... 85 

Figure 4.7 Before incubation in human serum (blue line), a higher percentage of echinocytes existed in

old RBC samples. After 120-min incubation in human serum, the percentage of echinocytes

decreased from 3.8 % to 2.1 % for one-week old RBCs, from 4.3% to 3.3% for two-week old RBCs,

from 4.5% to 3.3% for three-week old RBCs, from 6.2% to 4.3% for four-week old RBCs, from 8.6%

to 6.2% for five-week old RBCs, and from 9.3% to 7.3% for six-week old RBCs. For each condition,

n=2200-2800 RBCs. ........................................................................................................................... 86 

Figure 4.8 (a) Lack of ATP increases the spectrin-membrane affinity, leading to higher RBC stiffness.

When ATP is re-synthesized, spectrin-membrane binding becomes dynamic, causing the RBC to be

more deformable. (b) ATP concentration in RBCs that were stored from one week to six weeks. For

each data point, n=3 samples with each sample containing about 8000 RBCs. Older RBCs showed

lower ATP concentration than fresh RBCs, and their ATP concentrations increased during the

serum-incubation process. Steady-state values were reached by ~30 minutes. .................................. 89 

Figure 4.9 RBCs were cultured in human serum for 8 hours. For each sample, the intracellular ATP

concentration by 120 minutes and at the end of the 8-hr incubation was compared. For each

condition, n=3 samples with each sample containing about 8000 RBCs. For all stored RBCs (one-

week to six-week old), no difference was observed, indicating that longer incubation (>120 minutes)

did not induce further re-synthesize of ATP. ...................................................................................... 90 

Page 13: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

1

Chapter 1 

1. Introduction

1.1. RBC introduction

Human body is composed of about 37 trillion cells [1]. They provide structure for the body and

carry out specialized functions. Cells need energy to carry out their functions, which consumes

oxygen, and this requires the oxygen to be delivered to every part of the human body after the

oxygen is breathed in. The transportation of oxygen is carried out by the circulatory system

where the heart pumps oxygenated blood to the body and deoxygenated blood to the lungs. The

most abundant cells in blood are red blood cells (RBCs) which are the principle entities for

delivering oxygen, as shown in Figure 1.1(a). Among all the 37 trillion cells, 30 trillion of them

are RBCs, and approximately 2.4 million new RBCs are produced per second in human adults.

RBCs develop in the bone marrow and circulate for about 100–120 days in the body before their

components are recycled by macrophages.

Although RBCs are the most abundant cells in human body, they are unique in that unlike other

cells, mature RBCs lack a cell nucleus and most organelles. RBCs are produced in the bone

marrow with a nucleus and when they almost reach maturity, they then undergo enucleation

process when RBC’s nucleus is removed. The enucleation process cuts off a segment of the cell

containing the nucleus, which is then extruded from the RBC and further swallowed by a

macrophage [2]. The absence of a nucleus allows the RBCs to provide the maximum space for

hemoglobin to carry more oxygen. It also allows the cell to have its distinctive bi-concave shape

and be flexible, as shown in Figure 1.1(b). Such distinguishing feature provides RBCs the ability

to undergo large passive deformations during repeated passage through the narrow capillaries of

Page 14: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

2

the microvasculature to fully deliver oxygen. Due to RBC’s lack of nucleus and most organelles,

RBC’s membrane, its only structural component, accounts for most of cell’s antigenic, transport,

and mechanical characteristics [3]. Abnormal RBCs with compromised deformability accounts

for decreased RBC life span, and leads to anemia in different RBC disorders.

 

Figure 1.1 (a) RBC takes oxygen from lungs and releases the oxygen to tissues. (b) RBC’s thickness at the thickest point is 2 – 2.5 µm and the minimum thickness in the center is 0.8 – 1 µm. RBC has a dimeter of 6 – 8 µm.

 

The RBC membrane plays a great role in keeping deformable. There are two major components

of RBC membrane: 1) lipid bilayer formed by its lipid main constituents with many

transmembrane proteins embedded; and 2) a spectrin-based membrane skeleton located on the

inner surface of the lipid bilayer [3].

Page 15: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

3

1.1.1. Membrane lipids

The lipid bilayer composition participates in defining physical properties such as fluidity and

deformability of a cell. The lipid bilayer is composed of cholesterol and phospholipids in equal

proportions by weight. Cholesterol is evenly distributed between the inner and outer leaflets.

Five major phospholipids are asymmetrically disposed. Phosphatidylcholine and sphingomyelin

are located in the outer leaflet, while phosphatidylethanolamine and all phosphatidylserine,

together with the phosphoinositol, are confined in the inner layer [4][5]. The phospholipid

asymmetry is maintained by a group of both energy-dependent and energy-independent transport

activities. Energy-dependent activities include “flippases” which move phospholipids from the

outer layer to the inner layer, and “floppases” that do the opposite. Both “flippases” and

“floppases” describe the movement of phospholipids against a concentration gradient.

“Scramblases”, on the other hand, move the phospholipids bi-directionally down their

concentration gradients, which is an energy-independent activity [6][7]. These phospholipids

movement activities involve few proteins such as P-type ATPases, ATP-binding cassette

transporters and a putative scramblase [8][9].

The maintenance of the asymmetric distribution of phospholipids, plays a great role in regulating

biological function, especially the localization of phosphatidylserine to the inner layer.

Phosphatidylserine can act as an apoptosis signal when they are flipped to the outer layer of the

membrane, then macrophages can recognize the exposed phosphatidylserine and engulf the

RBCs [10]. Normally phosphatidylserine is actively held in the inner side of the cell membrane

by “flippase”, and the confinement of this lipid in the inner layer is essential for the RBC to

survive its frequent encounter with macrophages, especially in the spleen where the pores of the

Page 16: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

4

red pulp meshwork in spleen are quite small (5 µm), which give RBCs more chance to interact

with macrophages. However, when the “flippase” activity is lost, which could happen when an

RBC undergoes apoptosis or with some disease, phosphatidylserine is no longer restricted in the

inner layer of the membrane. “Scramblase” will take place and phosphatidylserine will exchange

between the two sides of the cell membrane, resulting more of phosphatidylserine exposed on the

outer layer. This expose of phosphatidylserine has been suggested to play a role in destruction of

thalassemia and sickle cell disease RBCs [11][12][13]. Furthermore, phosphatidylserine shows

higher adhesion, and the restriction of phosphatidylserine to the inner layer of membrane also

inhibits the adhesion of RBCs to vascular endothelial cells [14]. When phosphatidylserine is kept

in the inner layer of the membrane, the interaction between phosphatidylserine and inner spectrin

network can also regulate membrane mechanical property [15][16].

1.1.2. Membrane proteins

There are more than 50 types of transmembrane proteins existing on RBC membrane, as shown

in Figure 1.2. A large fraction of them define the various blood group antigens, and other

proteins serve as transport proteins, signalling receptors, and etc. [17][18]. Among all

transmembrane proteins, two macromolecular complexes of membrane proteins are the most

relevant to structural integrity of RBC membrane, one Ankyrin-based, and the other 4.1R

protein-based. The structural integrity related proteins are usually responsible for the linking of

membrane to the skeleton network underneath. For example, Band 3 and RhAG link the lipid

Page 17: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

5

bilayer to the skeleton through the interaction with Ankyrin. Glycophorin C, Rh, and other

proteins, on the other hand, build their connection to the skeleton through protein 4.1R

[19][20][21]. The membrane protein linkages with skeletal proteins play an important role in

regulating the interaction between the lipid bilayer and the membrane skeleton, and thus keep the

RBCs to maintain the favorable membrane surface area for the maximum oxygen exchange.

Meanwhile, the lodging and dislodging between the membrane proteins to the skeleton ensure

the RBC membrane to be deformable enough so they could squeeze through the capillaries

which are only few micro meters wide [22]. In addition, Band 3 also plays a key role in

regulating RBC metabolism in ion and gas transport functions [23].

1.1.3. Skeleton proteins

RBC membrane skeleton is composed majorly of spectrin (mostly spectrin α and spectrin β),

actin and some associated proteins including protein 4.1R, Ankyrin, adducin, and etc. (Figure

1.2)[24][25]. Spectrin is a long, flexible, helix like protein composed of two chains (spectrin α

and spectrin β) orientated in opposite directions. Each chain contains multiple spectrin repeats

with a “head” end and a “tail” end [26]. The functional domains at the “head” end are for

spectrin dimer-tetramer association and for Ankyrin binding. The domains at the “tail” end are

for binding to protein 4.1R, protein 4.2R and actin [27][28]. Isolated spectrin α or spectrin β

chains bind to each other at a pair of nucleation sites by repeats near the spectrin tail, while the

binding between spectrin α and spectrin β is through the association of two proteins called dimer

and tetramer [29]. The tetramers dissociate and reform when the membrane is deformed by shear

forces caused by the blood shear flow [30]. This allows the RBCs to sustain great distortions

when RBCs are passing through the micro vessels. The skeleton network is tethered to the RBC

Page 18: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

6

membrane at two sites: one mediated by Ankyrin that couples spectrin to Band 3 and the other

mediated by protein 4.1R that couples to Glycophorin C [31]. The affinity of the spectrin

skeleton and membrane can be regulated by the proteins connecting between them, and the

continuous de- and reattachment of the spectrin-membrane binding results in a deformable RBC

membrane.

 

Figure 1.2 Current model of the RBC membrane. A phospholipid bilayer is tethered to the spectrin network (spectrin α and spectrin β proteins) via a number of proteins such as band 3, ankyrin, protein 4.1, and Glycophorin. Spectrin α and spectrin β form the structure underneath cell membrane and provide mechanical support. Most of the known proteins are shown(reproduced by permission of [26]).

Page 19: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

7

1.2. RBC stiffness measurement

1.2.1. Filtration  

Stiffer RBCs can enhance the chance of being trapped and cleared in the spleen, which has been

shown in animal in vivo studies [32]. The spleen consists of the following parts: 1) the white pull,

a lymphoid tissue that contains the most of the immune effector cells, 2) the red pulp, a reticular

meshwork where aberrant RBCs are trapped and destructed, 3) and a marginal zone between the

white pulp and red pulp [33]. Healthy RBCs, which are quite deformable, are able to pass

through the meshwork in the spleen easily and continue travelling in the human body. However,

due to the small pore size (~5 µm) of the red pulp meshwork [34], less deformable RBCs can be

trapped and further cleared by macrophages. This mechanism is usually used for aged RBCs to

be cleared, since after about 120 days of circulation, aged RBCs become stiff and recycled by the

spleen. Inspired by the mechanism of spleen, the filtration method was introduced as a technique

for measuring RBCs deformability. This method examines the ability of multiple RBCs to pass

through membrane with small sized pores as the filter [35]. During filtration, whole blood (or

partially diluted blood) is passed through the holes in a membrane filter by applying a certain

positive or negative pressure. Blood containing more stiff RBCs takes longer time or requires

higher pressure to pass through membrane filter.

The filtration technique is easy to implement for RBC deformability measurement. However, to

compare the deformability between samples, it requires the testing samples to have the same

concentration of RBCs [36]. Additionally, if whole blood sample is used, white blood cells

contained in the sample are stiffer than RBCs, which could also potentially induce the clogging

of the pores, result in a longer passing time, and require a higher pressure [37][38]. Besides the

Page 20: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

8

stiffness of RBCs or white blood cells, other factors could also influence the filtration results.

For example, diabetic patients may have a higher blood plasma viscosity [39], resulting in the

high viscosity of whole blood, which could cause a longer passing time in filtration measurement.

Thus, measuring RBCs deformability by the filtration method could result in significant

variations across experiments.

Microfluidic filtration was designed to solve the issues of traditional filtration techniques by

fabricating multiple micro channels and observing RBCs under microscopy imaging. Whole

blood or PBS-diluted RBCs can be introduced into the microchannels with a certain pressure or

flow rate to flow RBCs through the microchannels. The behaviors (e.g., deformation and rate of

shape recovery) of RBCs when passing through the microchannels are recorded by camera, and

RBC stiffness is analyzed by data processing and mechanics modeling. Stiffer RBCs take longer

time to pass through the microchannels or get trapped in the channels [40][41]. By using reversal

flow, stiffer RBCs can also be trapped at a certain area [42] and the number of stiff RBCs can be

counted and used to represent the percentage of stiff RBCs in the sample. By using microfluidic

channels, only a small amount of blood is required, and high-throughput deformability

measurement of RBCs can be realized. Microfluidic devices have also been used to mimic in

vivo capillary blood flow system (Figure 1.3) [43][44]. Deformability is determined by

quantifying the threshold pressure required to traverse the constrictions. Microfluidic filtration

devices have been used to measure the deformability of malaria-infected RBCs [45] or sickle cell

disease RBCs [46]. This method could provide deformability assessment of an RBC population,

and only phenomenological metrics of RBCs deformability instead of inherent stiffness or

modulus of RBCs membrane can be measured.

Page 21: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

9

 

Figure 1.3 (a) An array of parallel micro channels. (b-d) Artificial microvascular networks mimicking vessels can be used to measure the deformability of red blood cells (reproduced by permission of [47]).

Page 22: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

10

1.2.2. Micropipette aspiration  

Population-based measurements are able to assess the relative deformability change of RBCs;

however, they also mask the behavior of minority subpopulations and the differences between

cells. Single RBC deformability measurement is required to understand the heterogeneity in an

RBC sample. Micropipette aspiration is one of the first single cell measurement techniques and

has been used to measure the mechanical properties of RBC membrane [48]. A micropipette

aspiration measurement system consists of a glass micropipette with an inner diameter of 1-3 µm.

A negative pressure is applied through the micropipette, as shown in Figure 1.4 (a). The

deformation of the RBC membrane aspirated into the pipette is observed under a microscope.

Several RBC membrane stiffness measurement methods have been introduced [49]: 1) With a

given pressure, to measure the ratio between the length of RBC membrane aspirated and the

radius of the micropipette radius; 2) To measure the pressure required to aspirate the RBC

membrane to a certain length into the micropipette; and 3) To measure the pressure necessary for

the whole RBC to be aspirated into the micropipette. By measuring the amount of RBC

membrane aspirated with different negative pressure, Evans determined that the shear modulus

of the RBC membrane to be ~9 µN/m [50][51]. By using microfabrication techniques, an array

of microfluidics-based micropipette aspiration channels were constructed for micropipette

aspiration measurement of RBC deformability with a higher throughput, as shown in Figure 1.4

(b). The pressure applied on each RBC was calculated through hydrodynamic equivalent circuit,

and the deformation of each RBC in the channel was measured by microscope. Stiffness of

RBCs parasitized by Plasmodium falciparum was measured along with other types of cells

Page 23: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

11

including RT4 human bladder cancer cells and L1210 mouse lymphoma cells [52][53].

 

Figure 1.4 (a) Different micropipette aspiration methods used to measure the mechanical properties of red blood cell (RBC) membranes. (b)Microfluidic pipette aspiration of RBCs using a funnel channel chain (reproduced by permission of [54]).

Page 24: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

12

1.2.3. Atomic force microscopy  

Atomic force microscopy (AFM), which was invented to achieve high resolution topography

imaging of an object, was adapted to perform RBC deformability measurement [55]. It uses a

cantilever with a tip of various shapes (triangular, parabolic or cylindrical) as a probe [56]. The

tip is placed on the sample surface and as the tip raster scans across the sample, the cantilever is

bent according to the surface topography, as shown in Figure 1.5. The minor vertical

deformation of the cantilever could cause position change of the laser beam reflected by the tip.

The reflected laser dot can be precisely detected by photodiodes, providing information of the

vertical deflection change of the cantilever [57].

Soon after the invention of AFM, it was also used to measure the mechanical properties of a

sample with a nanometric resolution, for example RBC membrane stiffness [58][59][60]. The

measurement was conducted by using AFM as an indenter and by monitoring the cantilever

deflection during indentation. By using this method, Young’s modulus of RBC membrane under

different physiological conditions was measured. The Young’s modulus of a healthy RBC’s

membrane Young’s modulus was found to be around 4.4 kPa [61]. Stiffness of RBCs from

thalassemia and diabetes mellitus was also measured by using AFM, and the effect of the disease

on RBC membrane stiffness was understood [61][62].

1.2.4. Optical tweezers  

The stiffness of RBCs can be measured by using optical tweezers. Optical tweezers can be used

to manipulate micro objects that could be as small as a single molecule. Optical tweezers use

highly focused laser beams to create an optical trap which could hold a micrometer or

Page 25: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

13

nanometer-sized dielectric particle at its centre [63]. When the particles are near to the center of

the trap, the focused laser beam applies a force on particles in the beam along the direction of

beam propagation, which is caused by the conservation of momentum. Photons that are scattered

by the tiny dielectric particles impart momentum to the dielectric particle, and the force

generated is also referred as scattering force. The scattering force can be determined by the

refractive indices of sample, surrounding medium, laser power, and sample size [64][65]. By

adjusting the power applied by the optical tweezers, a force with an order of pN can be achieved,

which is enough to deform the RBC. By measuring the deformation of the RBC membrane at the

given force, the mechanical properties of the RBCs could be determined (Figure 1.5).

The most common way of using optical tweezers is to use the scattering force to deform beads

which are attached to the RBC membrane. Two microbeads are first attached to RBCs from

opposite directions. The laser beam is then applied to generate a scattering force [66]. The

deformation of the RBC membrane is recorded by a microscope and the shear modulus can be

determined by using mathematical models. The shear modulus values of RBCs measured by

using this method vary from 10 to 30 µN/m [67][68].

Page 26: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

14

 

Figure 1.5 (a) Optical tweezers measuring the RBC deformability with focused laser beams that transfer linear momentum or angular momentum of light. To measure the shear modulus of the RBCs, two microbeads are attached to the opposite sides of a RBC and force was applied by the optical tweezers. (b) Experimental observations of RBC deformability measurement using optical tweezers, and simulation results. (c) 3D image of an RBC measure by AFM (reproduced by permission of [47]).

Page 27: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

15

1.2.5. Magnetic twisting  

Magnetic twisting cytometry is able to apply direct mechanical loading onto the membrane of a

single RBC. A ferromagnetic microbead is first tightly bound to the RBC surface, and then

magnetic field is generated to precisely control the torque applied to the bead. RBC membrane

deforms in response to the torque applied (Figure 1.6). By controlling the magnetic field, the

microbeads exhibit wide ranges of forcing time scale and forcing amplitude to RBC membrane.

The motion of the beads and the deformation of the RBC membrane are recorded by a camera.

Then the shear and loss moduli of single RBCs can be determined by varying oscillating

frequency from 0.1 Hz to 100 Hz and applied magnetic field from 1 Gauss to 10 Gauss [69]. By

using magnetic twisting cytometry, healthy RBC membrane shear modulus is measured to be

around 6-12 µN/m, while the loss modulus increases as the frequency increases, which is around

0.2-0.8 pNꞏµm. Magnetic twisting cytometry has also been used to determine the stiffness

change of diseased RBCs, for example Plasmodium falciparum parasitized RBCs. Stiffness

change of RBCs infected with Plasmodium falciparum has been measured over all the stages of

parasite maturation. Temperature range from room temperature to febrile condition (41 °C) was

also studied. The results showed a dramatic increase in the stiffness of Plasmodium falciparum-

infected RBCs at a temperature of 41 °C [70].

Page 28: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

16

 

Figure 1.6 RBCs with magnetic beads bound to the surface. SEM image of an RBC with a bead on the surface. c) Magnetic field is applied onto the bead to generate a torque to deform the RBC. d) Mechanical simulation to predict the deformation by the force applied by the magnetic bead. (Permission of [69] is not required for thesis/dissertation use).

Page 29: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

17

1.2.6. Quantitative phase imaging  

RBC membrane fluctuation is closely correlated with its mechanical property [71]. RBC can be

clearly seen under microscope and the biconcave shape can be clearly observed. RBC samples

are optically transparent in visible light, and only the light intensity can be measured by

conventional bright field microscopy, which does not provide enough information to determine

RBC membrane fluctuation. However, quantitative phage imaging techniques can measure the

amplitude and also phase information [72]. Since optical information is related to the physical

property of a sample, and there is a significant optical phase delay through the transparent

samples, quantitative phase imaging could provide high contrast images of the measured samples.

By using quantitative phase imaging, dynamic fluctuations of RBCs can be determined [73][74].

Fluctuations of RBCs membrane are used to characterize the membrane stiffness by determining

the in-plane shear modulus, and the stiffness difference between healthy RBCs and Plasmodium

falciparum-infected RBCs was analyzed [75].

Page 30: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

18

1.2.7. Microfluidic measurement  

Microfluidic devices have emerged as a promising tool to precisely control fluid and manipulate

cells. There has been growing evidence that cell stiffness can be used to provide a label-free

biomarker for determining cells states [76][77][78]. Thus, with the ability to manipulate a small

volume of samples and provide the heterogeneity information by measuring hundreds of

thousands single cells, microfluidic channels have been widely used to study cells stiffness as an

important biomarker. A real-time deformability cytometry was designed by flowing cells through

a microfluidic channel [79]. The deformability information was captured by a CMOS camera and

analyzed real time. This method was able to detect the cytoskeletal alteration, identify cell-cycle

phases and track stem cell differentiation into distinct lineages.

Since RBCs pass through micrometer-sized capillaries in vivo, microfluidic channels with a size

of micrometers can be used to mimic capillaries for measuring the stiffness of RBCs.

Constriction channels that are smaller than RBCs provide an efficient method to deform RBCs

for stiffness measurement. RBC’s deformability affects several parameters including elongation

and transit time when RBCs pass through the channels, and these parameters can be recorded by

high-speed imaging. A constriction channel device was first used to study the stiffness difference

between healthy RBC samples and malaria parasite-infected samples [80]. The results showed

that healthy RBCs showed higher elongation in the constriction channels and could pass through

small channels more easily than malaria parasite infected RBCs. RBC’s size can cause a

difference in the elongation of RBCs in the constriction channel. A two-stage microchannel was

designed to measure the size and the deformability of RBCs separately [81]. Besides RBC

deformation in the constriction channel, transit time can also be used to evaluate RBC

Page 31: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

19

deformability (Figure 1.7). This is usually applied when RBCs are driven into the channel with a

certain pressure. Due to the resistance from the constriction channel, stiffer RBCs take longer

time than softer RBCs to pass through the channels (transit time). Transit time is usually

measured by analyzing images taken by a high-speed camera [82]. By applying electric voltage

on two ends of the constriction channel, the impedance difference caused by the RBCs passing

through the channels can also be measured to determine the transit time and further the

deformability of the RBCs [54].

As mentioned before, RBC deformability measured by using constriction channels is affected by

the cell size. Along with cell size, adhesion between cell membrane and channel walls is also

coupled with RBC deformability. Thus, transit time does not necessarily reveal RBC’s

deformability since larger RBCs or more adhesive RBCs also show longer transit time. Since the

constriction channel provides mechanical stimulation by physically contacting RBCs and is

normally smaller than RBCs, the channel is susceptible to clogging.

 

Figure 1.7 A network of microfluidic channels to measure RBCs deformability. The transit time of the individual cells are recorded using a high-speed camera. (b) A microfluidic system for

Page 32: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

20

mechanical characterization of RBCs. The transit time is obtained from electrical impedance signal captured when RBCs are passing through the constriction channel (reproduced by permission of [54]).

Page 33: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

21

Besides the force provided by the direct contact of the microfluidic channel walls, the shear

stress generated by the flow can provide mechanical stimulation for the RBCs to deform [83].

Ektacytometry, for instance, is one of the primary methods for measuring RBC deformability.

Suspension of RBCs is subjected to the flow stress generated by different shearing geometries

(Couette flow, plate-plate, Poiseullie flow). RBCs are elongated by the shear stress, and the

deformation of RBCs can be captured by a technique termed laser diffractometry. When a laser

beam is applied on an RBC suspension, light is scattered by the RBCs and forms a single image,

referred to as a diffraction pattern. The shape of this diffraction pattern reflects the average shape

of a high number of cells (e.g., thousands). When determining deformability of RBCs, the

pattern is fitted to an elliptical shape with a long axis L and short axis W. RBC deformability is

described by the elongation index, EI=(L-W)/(L+W).

RBC deformability change due to glutaraldehyde treatment, which stiffens RBC membrane by

crosslinking membrane lipids and proteins, or heat treatment has been well studied by using the

Ektacytometry method. Based on this method, LORCA as a commercial product has been

manufactured. LORCA consists of two concentric cylinders with a gap of 0.3 mm which can

hold 1 mL whole blood for RBC measurement. The outer cylinder can rotate with varying speeds,

and the fixed inner cylinder is integrated with a laser light source along with a light sensor and

temperature control unit. Not only RBC deformability in terms of EI, but also RBC aggregation

behaviour has been studied by using this instrument, as shown in Figure 1.8. One disadvantage

of using EI to quantify deformability is that EI is phenomenologically defined and is strongly

dependent on flow velocity.

Page 34: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

22

 

Figure 1.8 (a)Hydrodynamic RBC deforming microfluidic device. Cells are focused to the center lines of the channels and deformed by fluid. Cell deformation is measured by analyzing images recorded by a high-speed camera. (b)Laser diffraction technique incorporating microfluidic rheometer to measure RBC deformability with elongation index defined (reproduced by permission of [47][54]).

Page 35: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

23

1.3. Pathophysiological conditions and RBC stiffness

1.3.1. Diseases  

RBC stiffness has been associated to many blood related diseases such as sepsis, diabetes, and

malaria infection. Sepsis is usually caused by immune response triggered by an infection due to

the injury or disease. It has been associated with hemodynamic alternations and microcirculatory

disturbances [84]. Although cardiac output has been observed to increase in sepsis patients,

nutrient blood flow to the liver, kidneys, and muscles is markedly reduced [85]. Thus, sepsis can

lead to poor organ function and insufficient blood flow which could further cause septic shock.

Deformability of RBCs of sepsis patients has been well studied, and the results show that

compared to healthy RBCs, a significant decrease in deformability of sepsis patients’ RBCs was

observed [86][87]. As discussed before, RBC membrane is a composite of a lipid bilayer and an

underlying spectrin skeleton. The connection between the skeleton and lipid bilayer determines

the RBC’s response to external force when passing through vessels. Adducin, an RBC membrane

protein, is a target for the calcium-dependant regulating protein which could promote spectrin-

actin interactions. Increased calcium cytosolic concentrations in sepsis is suspected to lower the

spectrin-actin interaction, which could further lower deformability of the RBCs in sepsis

[86][88].

Diabetes mellitus is a metabolic disorder characterized by hyperglycemia and usually caused by

low insulin levels. It is one of the most common causes of renal failure or cardiovascular

diseases [89][90]. It has been shown that diabetes is associated with higher blood viscosity and

reduced RBC deformability [91][92]. Lower RBC deformability can result in impaired perfusion

at the tissue level, which is a major complication of diabetes mellitus [93][94]. The abnormal

Page 36: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

24

RBC deformability of diabetes mellitus is believed to be caused by the lipid-protein interactions

and increased glycosylation [95][96]. Filtration technique has been used to study the

deformability of diabetes RBCs, where a constant negative pressure was given to diluted

suspension of RBCs to pass through membranes with straight channels. The passing rate was

used to indicate the RBCs deformability, and a compromised RBC deformability was observed

in diabetes RBCs [97]. Ektacytometry, as discussed before, has also been used to investigate

diabetes RBC deformability, and impaired diabetes RBC deformability was observed in terms of

EI [98][99].

1.3.2. Malaria  

Malaria causes the death of over 1 million people each year globally[100]. Malaria is a disease

caused by Plasmodium parasites. When affected blood enters the human body, sporozoites (a

motile infective form) enters the bloodstream and migrate to the liver where they infect liver

cells. During the potential dormant period in the liver, they multiply and differentiate to

thousands of merozoites, which will escape into blood and infect RBCs. The parasites wrap

themselves in the cell membrane to avoid being detected by immune system [101]. Within the

RBCs, the parasites further reproduce themselves, periodically breaking out of the host cells to

invade more fresh RBCs [102].

Although the parasites protect themselves from immunes system’s attack by hiding themselves

within liver and blood cells, parasites in the RBCs will lead to a more vulnerable RBC

membrane. Vulnerable RBCs will be easily destroyed by the spleen during the circulating. The

malaria parasites cleverly secrete some adhesive proteins on the RBC membrane, helping RBCs

stick to the blood vessels, preventing infected RBCs from entering the spleen [103]. However,

Page 37: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

25

the sticky RBCs will clog the vessels and the blockage of the micro vessels can cause symptoms

such as Placental Malaria [104][105].

Malaria infected RBCs’ membrane skeleton is strongly interrupted by parasite protein, which

could lead to abnormal RBC membrane stiffness [106]. As discussed before, RBC membrane

lipid bilayer is connected to the spectrin skeleton through proteins. When the RBCs are invaded

by malaria parasites, parasite-encoded proteins such as KAHRP (knob-associated histidine-rich

protein), PfEMP3 (P. falciparum Erythrocyte Membrane Protein 3) and RESA (ring-infected

erythrocyte surface antigen) will form some knob structures underneath the membrane

[107][108]. The knob structures modify RBC membrane and lower RBCs’ deformability.

The mechanical properties changes caused by malaria have been well studied through multiple

techniques including the methods introduced previously like Ektacytometry [109][110] and

micropipette aspiration [111][112]. Results have shown that malaria infected RBCs have much

lower deformability than healthy RBCs, and the compromised deformability contributed by

RESA is significantly more severe at febrile temperature condition (41 °C) [113].

1.3.3. Sickle cell disease  

Malaria-resistant genes have been identified in tropical populations at risk for the disease,

suggesting that naturally occurring mutations are an evolutionary response. Genetic factors

including sickle cell disease, thalassaemia, and etc. have provided different levels of malaria

resistance. Sickle cell disease, which describes a condition when RBCs can turn into rigid and

sickle-like shape, illustrates some trade-offs that have occurred due to the malaria [114]. Sickle

Page 38: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

26

cell disease RBCs are more rigid and vulnerable than healthy RBCs, which provides little

protection for the malaria parasites and easily exposes the parasites to immune system. Sickle

cell disease RBCs also have shorter life span (~40 days) than healthy RBCs (~120 days), which

leaves malaria parasites no time to fully grow and reproduce themselves to infect more cells.

Without endemic malaria, however, the sickle cell mutation is a total disadvantage. Sickle cell

disease describes the inheritance of two mutated β1-globin genes, the sickle hemoglobin gene

(HbS), from both parents. The mutated genes will encode abnormal hemoglobin S instead

normal hemoglobin A. Hemoglobin S causes no apparent effects in conditions of normal oxygen

concentration. However low oxygen concentration will cause hemoglobin S to aggregate and

form long chain of fibrous precipitates [115], known as polymerization. The presence of the long

chain polymers distorts the RBC from bi-concave shape to sickle-like shape, which is known as

sickling. Although sickled RBCs recover to bi-concave shape when the oxygen concentration is

rebuilt, cycling between sickled and unsickled causes RBCs to lose their ability to recover and

stay permanently sickled. Besides the shape change, sickled RBCs membrane is also modified

due to the higher concentration of Ca+2 , which leads to a higher RBC stiffness [116]. Sickled

RBCs possess abnormal shape and more rigid membrane, preventing themselves from passing

through narrow capillaries smoothly, leading to vessel occlusion and ischaemia [116][117].

Microfluidic experiments have been used to understand sickle cell disease RBCs morphological

and rheological changes [118]. Channels with size of capillaries were fabricated where sickle cell

disease RBCs were infused into the channels with pressure controlled, when the oxygen level

was controlled to investigate the change of RBCs under different oxygen concentration. The

Page 39: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

27

results show that sickle cell disease RBCs pass through the channels smoothly under normal

oxygen conditions, while blockage was observed when the oxygen concentration was lowered.

1.3.4. Sickle cell trait  

As described in the previous section, sickle cell disease patient inherits two mutated β1-globin

genes, the sickle hemoglobin gene (HbS), from both parents. Sickle cell trait carrier, on the other

hand, inherits one abnormal gene from one parent along with one normal hemoglobin gene (HbA)

from the other parent. Since individuals with sickle cell trait carry one normal hemoglobin gene,

they can generate normal hemoglobin, so they do not display the severe symptoms of sickle cell

disease patients. Thus SCT is typically considered to be benign and harmless [119], and they can

protect the carriers from malaria. However, exercise-related collapses within SCT athletes is a

serious complication. SCT as a risk factor for sudden death in physical training was first

comprehensively studied in 1987 [120]. Since then, there have been numerous reports

associating SCT and sudden death in young athletes engaged in vigorous exertion [121][122].

Autopsy observations showed vascular occlusion in brain, heart, lungs and livers in SCT sudden

death cases, indicating that the cause of death was vaso-occlusion due to SCT RBCs change

during physical exertion [123]. RBCs from SCD patients are known to become stiffer and even

sickled under deoxygenated or acidic conditions, causing the blockage of blood vessels

[116][124]. However, whether SCT RBCs also become stiffer under deoxygenated or acidic

conditions remain unknown.

Page 40: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

28

1.3.5. Stored RBCs  

Sickle cell disease or sickle cell trait related RBCs abnormality is the result of natural selection;

while storage-induced RBC degradation, referred to as “storage lesion”, occurs in RBC

preservation for transfusion. After major blood groups (O, A, and B) were discovered, which

made blood transfusion much safer, the preservation of blood donation was put onto studies to

realize non-direct transfusion [125]. The purpose of RBC storage system is to allow longer

storage with more RBCs circulating 24 hours after transfusion [126], and nowadays the “gold

standard” issued by Food and Drug Administration requires a minimum level of post-transfusion

RBC survival of 75 % after 24 hours [127].

During storage, RBCs undergo biophysical changes, referred to as storage lesions. Storage-

associated mechanical changes (e.g., decrease in deformability) can cause a decrease in

transfusion efficacy and an increase in harmful effects [128][129]. Stiff RBCs result in a higher

clearance by the spleen and are known to contribute to respiratory distress and systemic sepsis

[130][131]. Tissue ischemia, for instance, is believed to be contributed by microcirculatory

occlusion caused by poorly deformable RBCs [132]. Furthermore, among the critical ill patients

with sepsis who had older RBCs transfused, sepsis can be aggravated [133][134]. Since septic

patients have constricted vessels, poorly deformable RBCs can be trapped in the microcirculation,

leading to tissue hypoxia and exacerbating patients’ health conditions [135][136]. Thus,

mechanical changes of RBCs over the storage process are important to understand.

It is known that the storage process induces RBCs mechanical property changes and more

significant storage lesions have been observed in longer stored RBCs. However the clinical

outcomes between patients transfused with shorter stored RBCs and longer stored RBCs remain

Page 41: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

29

controversial in terms of higher rate of mortality or longer ICU staying time. More than 40

observational studies have concluded that use of older stored RBCs is associated with a

significantly increased risk of adverse complications after transfusion [137][138]. These studies

observed patients with patient number from 52 to 387,130 and examined the effect of RBC

storage on various clinical outcomes including mortality, rates of infection, and length of ICU

Stay in the hospital [139][140][141]. The time point to distinguish younger and older RBCs was

mostly chosen to be 14 days or 21 days [142][143]. In these studies, patients transfused with

older RBCs showed higher rate of mortality and infection, or longer ICU staying time than

patients transfused with younger RBCs, which seems to conclude that younger RBCs are used

exclusively might save lives.

Results from more recent clinical studies, on the other hand, reported no significant difference in

clinical consequences between fresher and older RBCs [144]. Five clinical studies detected no

important clinical consequences of older RBCs [145][146][147][148][149]. In two larger trials,

transfusion of younger RBCs, as compared with older RBCs, did not significantly reduce the

complications of prematurity in very-low-birth-weight infants (patient number 377) or reduce the

rates of organ failure or adverse events among 1098 patients undergoing cardiac surgical

procedures [150][151].

The storage of RBCs is known to cause cell degradation (e.g., RBC stiffness and ATP), referred

to as “storage lesion”. For instance, mechanical stiffness changes of stored RBCs have been

widely studied, and the results consistently revealed that the stiffness of RBCs increases over the

storage process [152][153][154][155]. However, the controversy in clinical outcome of

Page 42: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

30

transfusing fresher and older RBCs suggests that the degradation of stored RBCs can possibly be

reversed. Whether in vivo condition helps stored RBCs recover stiffness remains unknown.

Page 43: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

31

1.4. Research objectives

The overarching goal of the thesis is to develop microfluidic systems for characterizing

mechanical properties of RBCs and understand RBC property changes in pathological and

storage conditions. Specific objectives include:

To design a microfluidic system for sickle cell trait RBC mechanical characterization

under typical exercise-induced conditions. To measure stiffness change of RBCs over

blood storage using human-capillary like microchannels.

To quantify the stiffness recovery evolution of stored RBCs under in vivo-like conditions

(i.e., in human serum).

Page 44: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

32

Chapter 2

2. Stiffening of sickle cell trait red blood cells under simulated strenuous exercise conditions

2.1. Introduction

Athletes are often seen as the healthiest entities in the human population; hence, the occurrence

of sudden deaths in this group [156] can be shocking and have a devastating impact on

communities and the public [157][158]. According to an analysis conducted by the National

Collegiate Athletic Association (NCAA) from 2004 to 2008 [156], 36 out of 80 medical causes

of athletes’ death (45%) were identified to be exertional sudden deaths [159]. Among all the

exertional sudden deaths, sickle cell trait (SCT)-related cases caused most controversies

[160][161][162].

SCT describes the inheritance of one normal hemoglobin gene (HbA) from one parent along with

one mutated β1-globin gene, the sickle hemoglobin gene (HbS), from the other parent [163]. By

the end of 2009, there were approximately 300 million people worldwide with SCT.

SCT is typically considered to be benign and harmless [119]. However, as more cases of SCT

athletes’ sudden deaths were reported [159][164], heated debates over whether SCT should be

considered as a death cause during exercise and whether athletes should be screened for SCT

arose. Rationally analyzing these issues requires clear understanding of the properties of SCT

RBCs [159][165][166].

Vaso-occlusion, usually caused by the block of blood vessels, is one of the most fatal and

common symptoms in sickle cell disease (SCD). It also appears to be a crucial contributor to

Page 45: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

33

sudden deaths in SCT individuals [166]. RBCs from SCD patients are known to become stiffer

and even sickled under deoxygenated or acidic conditions, causing the blockage of blood vessels

[124][116]. However, whether SCT RBCs also become stiffer under deoxygenated or acidic

conditions is not known.

During strenuous exercise, athletes’ muscles are under maximal oxygen consuming condition,

which lowers the oxygen level in circulation. Furthermore, although human blood pH is

normally 7.35, in strenuous exercise the higher concentration of hydrogen ions in human body

makes blood pH drop below 7.0 and even to 6.8 under extreme conditions for a short time period

[167][168]. Along with deoxygenation, the acidic blood condition has also been confirmed to

trigger the sickling of SCD RBCs and is hypothesized to stiffen SCT RBCs [169][170].

In this chapter, we focused on determining whether lowered pH and deoxygenation conditions

can trigger the stiffening of SCT RBCs which could be associated with higher vaso-occlusion

risks. The measurements were made on both normal RBCs and SCT RBCs, using a microfluidic

system that is capable of controlling oxygen and pH levels.

2.2. Methods

2.2.1. Blood specimens  

The study was performed in accordance with the institutional guidelines for using human tissue

samples. Blood samples were collected for routine tests and used for study only after they were

completed for clinical tests and would be otherwise discarded. The study protocol was approved

by the Mount Sinai Hospital Research Ethics Board, in which the informed consent was not

required because the samples were selected retrospectively and no patient identification

Page 46: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

34

disclosed to the study, and the study had no effect on the clinical test or patient management.

SCT was confirmed by sickle cell test and standard hemoglobin electrophoresis in the clinical

laboratory. Blood samples including seven normal blood samples and seven SCT blood samples

were stored with ethylenediaminetetraacetic acid (EDTA, 1.5 mg ml−1) and used within 48

hours. Before introduced into the device under room temperature, blood samples tested under

normal pH level were diluted 200 times in phosphate-buffered saline (PBS, pH= 7.35± 0.05),

while blood samples tested under acid conditions were diluted 200 times in acid-adjusted PBS

(pH=7.10±0.05 and pH=6.85±0.05).

2.2.2. Device and measurement  

As shown in Figure 2.1(a), the microfluidic device consists of three parallel channels. The

middle channel is used for loading and testing RBCs, and the other two channels are used to

control the oxygen level in the middle channel. Diluted blood sample is introduced to the middle

channel and left settling for 15 minutes. Due to the presence of the carboxyl group of sialic acids

in the cell membrane, RBCs show negative charge, while glass slide has positive charge, so

RBCs strongly adhere to the glass substrate due to the difference in electrical charge on the cell

membrane and glass surface [171]. In experiments, when the pressure was varied from 0 Pa to 9

Pa, none of the RBCs was detached or revealed noticeable displacements. Two water tanks

containing PBS are connected to the inlet and outlet of the middle channel to maintain the

osmolality and pH level. Pumping either air or nitrogen into the two side channels controls the

oxygen level in the device due to the gas permeability characteristic of polydimethylsiloxane

(PDMS) [172]. The cross-sectional area of the three channels is 60 µm 300 µm. The gap

Page 47: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

35

between neighboring channels is 100 µm for facilitated gas exchange between the gas channel

and the cell testing channel. Details of fabrication and design of the microfluidic device have

been described in our previous work [173]. In addition, pH level is controlled by adding

hydrochloric acid and confirmed before and after each experiment by using pH measurement

instrument (Hanna FC 240B pH electrode). RBCs are deformed under shear stress (0.9 Pa)

generated by a regulated vacuum source (pressure difference is 1.8 kPa) and this shear stress is

comparable to in vivo condition [174]. After the release of the shear stress, the RBCs recover to

their original shape. The dynamic recovery process is captured using a CCD camera connected to

a microscope. Mechanical models are developed to extract the shear modulus of each RBC.

Page 48: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

36

 

Figure 2.1 (a) Schematic of the microfluidic device for RBC mechanical property measurement. RBCs in solution of different pH levels are loaded into the middle channel and adhere onto the glass channel bottom. Oxygen level is accurately varied by pumping air or nitrogen into the two side channels. Schematic illustration of the deformation of a cell element. (b) RBCs are deformed under shear stress (1.8 kPa). After the release of shear stress, RBCs recover to their original shape

Page 49: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

37

2.3. Results and discussion

2.3.1. Determination of RBC shear modulus  

When an RBC is in a shear flow, its membrane undergoes deformation by shear stress.

According to the Kelvin-Voigt (KV) model [175][176],

T λ (1)

where T (µN/m) is the average tension force acting on RBC membrane, μ (µN/m) is the elastic

shear modulus expressed in force per unit length, and λ is the extension ratio of RBC membrane.

λ where l is the RBC’s length when deformed under shear stress, and l is the RBC’s

original length [177].

The flow in microfluidic channel in this work is driven by a pressure difference (ΔP). The

velocity profile of pressure-driven flow is

ν 1 (2)

where η is dynamic viscosity of the fluid. Since the microchannel has a rectangular cross-section

and its width (w) is much larger than the channel height (h), hydraulic radius R is approximately

equal to h. Since w>>h, the error caused by the approximation is minor and doesn’t affect the

conclusion in this work. Shear stress on the microchannel bottom where RBCs are located is

τ η │ ∆ (3)

Page 50: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

38

To calculate tension force T in the direction of extension, a small element (dA) is taken for force

equilibrium analysis, as shown in Figure 2.1. Since only the steady-state behavior is considered,

tension force and shear force on each element are balanced.

T x 2τ Y x (4)

where Y x is the half-width of the element dA. Thus, tension force T is

T (5)

where A is the surface area of the RBC. Shear modulus μ is determined by substituting Equation

(5) into Equation (1).

2.3.2. SCT RBCs become stiffened under lower pH  

RBCs from healthy donors (control) and SCT individuals were first tested using the microsystem

without adjusting the pH level. Figure 2.2 shows that under normal pH levels, SCT RBCs are

significantly stiffer than normal RBCs, which is in agreement with previous results [173][178].

The higher stiffness of SCT RBCs could lead to a higher blood viscosity in the vascular system

[179]. In large blood vessels, lower RBC deformability limits cell orientation in flow and thus

increase blood viscosity [180]. In small blood vessels, stiffer RBCs lead to a lower Fahraeus-

Lindqvist effect, which increases the flow resistance and blood viscosity [181]. In healthy non-

SCT carriers, endothelial cells lining the blood vessels can generate vasodilators (e.g., Nitric

Oxide) to mediate increased blood viscosity. Differently, SCT carriers are known to develop

impaired vascular functions and generate less vasodilators [166][182].

Page 51: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

39

We next measured the shear modulus of both SCT and normal RBCs at lower pH levels. As

shown in Figure 2.2(a), the shear modulus of all the seven tested SCT RBC samples consistently

became higher under a low pH level of 6.85 (p < 0.05). This low pH level was chosen because it

is known that in strenuous exercise the higher concentration of hydrogen ions in human body can

make pH drop to approximately 6.8 [167][168]. We then aimed to investigate whether moderate

exercise could induce SCT RBC stiffening. Hence, we conducted experiments under an

intermediate pH value (pH 7.10) which mimics the moderate exercise condition and no

significant difference in shear modulus was observed, as shown in Figure 2.3 (c).

In contrast, RBCs from normal subjects did not respond significantly to low pH levels. When pH

was reduced to 6.85, as shown in Figure 2.2(b) including seven samples’ data, the shear modulus

of normal RBCs increased slightly. However, the difference was not statistically significant (p≥

0.05). In our experiments, normal RBCs only started to reveal stiffness changes with statistical

significance at pH levels lower than 6.0 (data not shown) which is a physiologically irrelevant

condition [183][184][185].

As can be seen in Figure 2.2(c) where all of the seven samples are summarized together and

reported in box plots, when pH was reduced from 7.35 to 6.85, SCT RBCs were stiffened

significantly while control RBCs’ stiffness only increased slightly. The average shear modulus of

normal RBCs increased from 2.09 ± 0.67 µN/m to 2.27 ± 0.83 µN/m while the average shear

modulus of RBCs from SCT individuals increased significantly from 2.81 ± 0.7 µN/m to 4.05 ±

1.08 µN/m. The results indicate that SCT RBCs are inherently stiffer than control/normal RBCs,

and SCT RBCs are more sensitive to lowered pH levels than normal RBCs. Statistical analysis

confirms a significant difference of the shear moduli of normal RBCs and SCT RBCs under the

Page 52: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

40

acidic condition (2.27 ± 0.83 µN/m vs. 4.05 ± 1.08 µN/m) (p=1.9 × 10-82). The significant

stiffening of SCT RBCs under the acidic condition could cause difficulties for the SCT RBCs to

pass through minuscule vessels and capillaries. Along with the increased blood viscosity, this

could lead to a higher chance of vessel blockage. The resulting vaso-occlusion events could

result in acute ventricular failure and contribute to sudden death in SCT carriers.

Page 53: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

41

 

Page 54: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

42

Figure 2.2 (a) Shear modulus of SCT RBCs under physiological pH level 7.35 (blue) and acid pH level of 6.85 (red). Error bars represent standard error of the mean value. All samples (n=30-70 for each sample) show a significant difference between different pH levels (*p < 0.05). (b) Shear modulus of normal RBCs does not reveal significant difference under different pH levels (p value > 0.05; n=30-70 for each sample). Error bars represent standard deviation. c) Box plot showing the summarized shear modulus of RBCs from the 7 tested SCT samples and the 7 tested normal samples under different pH conditions(**p=5.5 × 10-15,***p=2.2 × 10-29,****p=1.9 × 10-82; n=200-300).

2.3.3. Hypoxia does not induce SCT RBC stiffening  

We then tested the oxygen effect on SCT and normal RBCs. As shown in Figure 2.3, the

stiffness of SCT RBCs was not found to increase by deoxygenation under physiological pH

(7.35), which is in agreement with previously reported result [173]. We speculated that under

acidic conditions deoxygenation might cause SCT RBCs to become even stiffer compared to the

condition of low pH only. However, the measurement results revealed that deoxygenation is not

capable of stiffening SCT RBCs further. In the deoxygenation experiments, we infused Nitrogen

into the two side channels for 20 minutes to reduce the oxygen concentration from 20% (air) to 0%

(pure Nitrogen). The validation of Oxygen depletion was described in our previous work [173].

As discussed in the previous section, the shear modulus of SCT RBCs increased from 2.81 ± 0.7

µN/m to 4.05 ± 1.08 µN/m under pH 7.35 and 6.85. These shear modulus values largely

remained the same when the channel was deoxygenated (for pH 7.35, oxygenated: 2.81 ± 0.7

µN/m vs. deoxygenated: 2.82 ± 0.7 µN/m; for pH 6.85, oxygenated: 4.05 ± 1.08 µN/m vs.

deoxygenated: 4.02 ± 1.04 µN/m). The differences were confirmed by using Mann-Whitney

nonparametric analysis to be insignificant (p>0.75). In order to decouple the effect of

Page 55: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

43

deoxygenation and pH levels, the RBC samples were diluted 200 times, and it was confirmed

after each experiment that there was no significant difference in pH after deoxygenation.

Page 56: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

44

 

Figure 2.3 At physiological pH 7.35 (a) and low pH 6.85 (b), normal RBCs elastic modulus did not change significantly when the channel was deoxygenated (*p>0.75). (c)(d) For SCT RBCs, although shear modulus became higher at pH 6.85 than at pH 7.35, deoxygenation did not induce further increase (*p>0.75). (e) Summarized shear modulus of SCT RBCs under different oxygen levels, and no difference was observed. Error bars represent standard deviation.

Page 57: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

45

2.4. Discussion

 

SCT was recently reported to be associated with strenuous exercise-related mortality

[159][160][164]. Existing results on mechanical properties of RBCs from SCT individuals are

limited. Here we examined SCT RBCs’ stiffness change under controlled oxygen and pH levels.

The results reveal that SCT RBCs are significantly stiffer than RBCs from non-SCT healthy

subjects. Lower pH resulted in 28% increase in SCT RBCs’ shear modulus (Figure 2.2).

The stiffness increase of SCT RBCs could be due to several physiological alterations. It was

reported that SCT RBCs contain a higher concentration of Ca+2, which can enhance the binding

of cytoplasmic domain of band 3 (CDB3) to the cytoskeleton bound Ankyrin [178][186]. This

stronger binding caused by increased Ca+2 can possibly contribute to the higher rigidity of the

membrane of SCT RBCs. At lower pH, CDB3 becomes even more compact [183][187], which

can further enhance the binding and stiffening of SCT RBCs. In addition, Monocarboxylate

transporter 1 (MCT-1) activity has been speculated to impact RBC stiffness [188][189][190].

Since MCT-1 activity is inherently stronger in SCT RBCs than normal RBCs at low pH, the

higher concentration of hydrogen ions leads to even stronger MCT-1 activity, the enhancement

of MCT-1 activity could also be responsible for the significant increase of shear modulus

measured on SCT RBCs at pH 6.85. However, since no significant difference was observed

when pH was lowered to 7.10, we speculate that at pH 7.10, cytoplasmic domain of band 3

(CDB3) is not sufficiently compact to enhance the binding to the cytoskeleton bound Ankyrin,

leading to insignificant stiffening of RBCs. The results also suggest that under normal exercise

conditions where pH only slightly decreases, the stiffening of SCT RBCs would not be

significantly evident.

Page 58: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

46

Experimental results also show that deoxygenation did not induce further stiffening of SCT

RBCs (Figure 2.3). During strenuous exercise, both blood oxygen and pH levels become lower.

Although deoxygenation did not directly impact the stiffness of SCT RBCs, low oxygen levels

during exercise can contribute to the lowering of blood pH [191]. Our data measured in the

simulated strenuous exercise condition indicate that low pH rather than hypoxia is effective in

triggering SCT RBC stiffening. In addition, although SCD RBCs are known to sickle under

acidic and/or hypoxia conditions, no sickling of SCT RBCs was observed under these conditions

in this work. Besides the stiffening of SCT RBCs, other exercise-induced physiological changes

can also possibly lead to a higher risk. For example, higher epinephrine during exercise has an

effect on SCT RBCs’ adhesion [192][193]. Increased adhesion can contribute to stronger

interactions between RBCs and epithelial cells and thus, trigger inflammatory pathways.

Dehydration, a common condition occurring during exercise, has also been reported to affect

RBCs’ physical properties [194]. The stiffening of SCT RBCs, inflammation, and dehydration

individually and together can be associated with a higher risk among SCT individuals during

strenuous exercise.

2.5. Conclusion

This study aims to address the question whether RBCs of SCT individuals become stiffened

during strenuous exercise. RBCs from SCT individuals and non-SCT subjects were tested under

simulated strenuous exercise conditions (i.e., low oxygen and low pH). The results show that

RBCs from SCT individuals are inherently stiffer and are sensitive to low pH which induces

significant stiffness increase in SCT RBCs, implying that the stiffening of RBCs could occur in

SCT individuals during strenuous exercise. Furthermore, the experimental results revealed that

Page 59: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

47

deoxygenation alone did not cause SCT RBCs to increase their stiffness. However, since low

oxygen levels contribute to the lowering of blood pH, the stiffening of SCT RBCs in vivo could

result from a combined effect of low oxygen and low pH.

Page 60: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

48

Chapter 3

3. Stiffness increase of red blood cells during storage

3.1. Introduction

More than 108 million blood donations are collected globally every year. Regulations in many

countries specify 42 days (6 weeks) as the shelf life for stored RBCs, and a first-in-first-out

inventory management approach is standard. Large-scale clinical studies involving 200-1,800

patients indicated that patients transfused with older RBCs tend to have a higher risk of mortality

than those receiving fresher RBCs [195][196][197][198].

During storage, RBCs undergo several biochemical and biophysical changes, referred to as

storage lesions. Storage-associated biomechanical changes (e.g., decrease in deformability) can

cause a decrease in transfusion efficacy and an increase in harmful effects [128][129]. Poorly

deformable RBCs result in a higher clearance by the spleen and are known to contribute to

respiratory distress and systemic sepsis [130][131]. Clinical research has also identified a

number of disease conditions such as splanchnic ischemia developed in patients who had been

transfused with older RBCs. Tissue ischemia, for example, is believed to be contributed by

microcirculatory occlusion which is caused by poorly deformable RBCs [132]. Furthermore,

among the critical ill patients with sepsis who had older RBCs transfused, sepsis can be

aggravated [133][134]. Since septic patients have constricted vessels, poorly deformable RBCs

can be trapped in the microcirculation, leading to tissue hypoxia and exacerbating patients’

health conditions [135][136]. Thus, mechanical changes of RBCs over the storage process are

important to understand.

Page 61: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

49

Micropipette aspiration was first used to investigate the deformability change during RBC

storage [199]. The negative pressure required to aspirate an RBC into the micropipette was used

as an indicator of RBCs’ deformability. The result showed that the longer stored RBCs required

a higher negative pressure to be aspirated into the micropipette, suggesting a poorer

deformability. Optical tweezers were used to stretch an RBC, revealing that 35-day RBCs were

more difficult to stretch than fresh RBCs [200]. More recently, ektacytometry was used to

characterize the deformability of stored RBCs by stretching RBCs with shear stress induced by a

rotating plate. The extent of RBC elongation, defined as elongation index (EI), was used to

indicate the deformability under a certain shear stress [152][153]. Microfluidic measurement was

also reported for investigating stored RBCs’ deformability, wherein the deformation index (DI)

did not show significant differences at different storage time points [201].

The deformation of an RBC contains three modes: area expansion, shear, and bending of the cell.

Area expansion describes the isotropic area dilation or compression of the membrane surface

under a force. Shear describes the extension of the in-plane extension of the membrane surface

with the same membrane area. Bending characterizes the deforming behavior of a membrane

under an out–of-plane force. Micropipette aspiration is suitable for measuring the area expansion

stiffness, while shear stiffness (named as shear modulus in many cases, although with a unit of

N/m) can be measured by using optical tweezers or shear flow. Both area expansion stiffness and

shear stiffness reflect an RBC’s in-plane properties. When an RBC is deformed under an in vivo-

like flow condition, bending must also be considered in the deformation. In this case, RBC

deformation results from the collective effects of area expansion, shear, and bending. Although

an in vivo-like flow condition was created on a microfluidic device, the defined deformation

index (DI) is a phenomenological parameter which strongly depends on flow velocities, which is

Page 62: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

50

unsuitable to use as a metric for evaluating RBC mechanical degradation during storage. More

detailed discussion on DI’s dependence on flow velocities is provided in the Experimental

Results section.

This chapter reports microfluidic measurement of stored RBCs, wherein RBCs are deformed in

the folding mode and velocity-independent effective stiffness is used to evaluate RBCs’

mechanical degradation during storage. The results reveal that RBCs’ effective stiffness already

becomes significantly higher in the first week of storage and consistently increases over the 6-

week storage period. Interestingly, the time points of effective stiffness increase were found to

coincide well with the degradation patterns of S-nitrosothiols (SNO) and adenosine triphosphate

(ATP) in RBC storage lesion.

3.2. Device fabrication

The PDMS microfluidic device was fabricated by using standard lithographic techniques. The

masks for the device were designed in AutoCAD (Autodesk, Inc., USA) and were printed as

transparencies by CAD/Art Services, Inc.. Table 3.1 shows the details of the fabrication of

alignment marker. Alignment of features on the photoresist layers was accomplished by first

patterning a chromium layer on a glass slide. The SU-8 negative photoresist (MicroChem,

Newton, MA, USA) was used both as a seeding layer and feature master with two different

heights (20 µm and 60 µm). Details on fabricating the SU-8 layers are shown in Table 3.2.

PDMS was made with the standard 10:1 mixing ratio of base to curing agent. PDMS was then

poured onto the mold and cured in an oven at 80 °C for 40 minutes. The entire PDMS structure

was peeled off the SU-8 master. The structure was washed in acetone, methanol, and DI water,

dehydrated on a hotplate at 150 °C for 10 min, and then O2 plasma bonded to a clean glass slide.

Page 63: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

51

Table: 3.1: Fabrication of alignment markers

Step Procedure

1 Starting with chromium-coated slides, clean with acetone, methanol, and DI water

2 Dehydrate glass slides on hotplate at 150 °C for 30 min

3 Pour positive photoresist S1811 onto the slide

4 Spin: step 1: 500 rpm, 5 s, 1 acl; step 2: 3000 rpm, 30 s, 8 acl

5 Pre-bake slide on hotplate at 95 °C for 2 min to remove solvent

6 Place slide and mask in mask aligner. Soft contact UV exposure for 6 s

7 Develop slide in MF-321 for 2 min. Rinse in DI water and dry with N2 gun

8 Hard bake slide on hotplate at 95 °C for 1 min

9 Etch chromium layer in CR-2 for 1 to 2 min. Rinse in DI water and dry with N2 gun

10 Develop in AZ-300T for about 5 min. Rinse in DI water and dry with N2 gun

Table 3.2: Fabrication of SU-8 seeding, 20 µm and 60 µm feature layers

Step Procedure

1 Dehydrate glass slides on hotplate at 150 °C for 30 min

2 Pour SU-8-5 on entire slide (seeding layer)

3 Spin: step 1: 500 rpm, 5 s; 1 acl, step 2: 3000 rpm, 30 s, 3 acl

4 Pre-exposure bake: 65 °C for 1 minute, 95 °C for 3 min

5 Place slide in mask aligner. Flood exposure for 6 s

Page 64: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

52

6 Post-exposure bake: 65 °C for 1 minute, 95 °C for 1 min

7 Hard bake at 175 °C for 2 hours

8 Pour SU-8-25 on entire slide (20 µm layer)

9 Spin: step 1: 500 rpm, 5 s; 1 acl, step 2: 4000 rpm, 30 s, 3 acl

10 Pre-exposure bake: 65 °C for 1 min, 95 °C for 5 min

11 Place slide and mask in mask aligner. Soft contact exposure for 10 s

12 Post-exposure bake: 65 °C for 1 min, 95 °C for 5 min

13 Develop in SU-8 developer for several min

14 Pour SU-8-25 on entire slide (60 µm layer)

15 Spin: step 1: 500 rpm, 5 s; 1 acl, step 2: 1000 rpm, 30 s, 3 acl

16 Pre-exposure bake: 65 °C for 5 min, 95 °C for 15 min

17 Place slide and mask in mask aligner. Soft contact exposure for 12 s

18 Post-exposure bake: 65 °C for 1 min, 95 °C for 4 min

19 Develop in SU-8 developer for several min

20 Hard bake at 175 °C for 2 hours

3.3. Methods

The device (Figure 3.1) consists of wide channels (500 µm × 60 µm) for introducing cells and a

constriction channel (12 µm × 20 µm) for inducing shear force to deform RBCs. Two focusing

channels were used to center and reorient RBCs to ensure that most of the cells were deformed

symmetrically in the center of the constriction channel [201]. Before introduced into the device,

fresh RBCs and RBCs (from 5 subjects) stored for one week to six weeks were diluted 200 times

Page 65: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

53

in phosphate-buffered saline (PBS) to minimize the coincidence occurrences of multiple cells in

the channel. Although pH of RBC storage medium decreases to 6.5 by the sixth week [202],

stored RBCs were tested in PBS (pH: 7.35) since pH 7.35 is more physiologically relevant [203].

Before the RBCs were introduced into the microfluidic device, their shape at rest was first

evaluated as shown in Figure 3.1(b). During storage, RBCs progressively change their shape

from smooth discs called discocytes to bumpy discs called echinocytes, and the percentage of

echinocytes over the storage process increases to around 12% by the end of six-week storage.

The deformation of the bent RBCs in the channel was recorded by a camera with a frequency of

5,000 frames per second and shutter time of 30 µs (HiSpec 1, Fastec Imaging Corp., U.S.).

Page 66: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

54

Figure 3.1(a) Microfluidic device with a constriction channel where RBCs are bent by drag force. (b) RBC’s shape change over storage showing that up to 12% cells change their shapes from discocytes to echinocytes.

 

The size of the constriction channel was chosen via finite element simulation and experimental

validation. Figure 3.2 shows that the highest shear stress occurs on the RBC membrane closest to

the channel wall. The flow velocity used in this study was 0.02 m/s, which is limited by the

Page 67: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

55

shutter time of the camera. When the channel width is smaller than 10 µm, the highest shear

stress increases significantly to 300 Pa, sufficiently high for causing RBC lysis [204][205].

When the channel width is larger than 14 µm, the shear stress is not sufficient to bend the RBCs.

Thus, the channel width was chosen to be 12 µm in this study.

After an RBC enters the constriction channel, the cell reaches its steady-state shape before

approaching the end of the channel which is 160 µm long. The shear stress acts on the cell and

imposes a drag force to deform the RBC. The drag force can be approximated analytically

according to [206]

𝐹 3𝜋𝜇𝑑𝑣𝑓 (1)

where µ is medium’s viscosity; 𝑣 is the flow velocity; d is the diameter of the deformed RBC

(see Figure 3.1); wall factor 𝑓 [207] with the ratio λ=d/D, where D is the size of the

constriction channel on the microfluidic device; then shear force acting on the end of the cell

𝐹 𝐹. Finite element simulation in COMSOL, using a model of a rectangular microchannel

and deformed biconcave RBC shape, confirmed that the error in drag force quantification with

Eq. (1) is consistently within 10% for all flow velocities. The RBC’s deformed diameter d and

the deflection L were both measured from 5,000 Hz imaging. The effective stiffness of the RBC,

reflecting the collective effects of area expansion, shear, and bending, is hence

𝑘 (2)

Page 68: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

56

 

Figure 3.2(a) Simulation shows that the highest stresses occur in RBC membrane regions closest to the channel walls. (b) At a flow velocity of 0.02 m/s, shear stress becomes significantly higher than 300 Pa for a channel width smaller than 10 µm, which can lead to RBC lysis. When channel width is larger than 14 µm, shear stress is too low to deform RBCs.

 

 

 

 

Page 69: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

57

3.4. Experimental Results

Microfluidic measurement of effective stiffness was first validated by AFM indentation. In

microfluidic measurement, RBC is located in the center of the microchannel, and shear-induced

force is symmetrically applied to the cell. As shown in Figure 3.3, the distributed shear stress 𝜏

is equivalent to a force 𝐹 acting on a position, y=y’ such that the two force systems are

equivalent with the same resultant force and the same resultant moment. The AFM experiment

was designed accordingly, and individual RBCs with half adhered on a substrate and the other

half suspended were measured by AFM indentation. Comparable effective stiffness values from

microfluidic measurement and AFM measurement of 6-week old RBCs were obtained (95 ± 7

µN/m, n = 220 vs. 108 ± 18 µN/m, n = 11).

3.4.1. AFM validation  

To mimic the situation of an RBC experiencing shear force in the microfluidic channel, in the

validation experiments, we fixed the center of the RBC and used an AFM (Bioscope Catalyst,

Bruker) tip to deform an RBC edge.

In microfluidic measurement, RBC is located in the center of the microchannel, and shear-

induced force is symmetrically applied to the cell. As shown in Figure 3.3, the distributed shear

stress (𝜏 ) is equivalent to a force (𝐹 ) acting on a position, 𝑦 𝑦′ such that the two force

systems are equivalent with the same resultant force and the same resultant moment.

τ y y′ dy τ y′ y dy (3)

τ at each position y was quantified from finite-element simulation and satisfies

Page 70: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

58

τ dy F (4)

Combining Eq. (3) (4) gives

y'=0.7385r

where r is the radius of the RBC, which is within the range of 3-4 µm but varies from one cell to

another. In experiments, each RBC’s radius was measured via imaging. These results mean that

the microfluidic shear situation is equivalent to the application of F , along a line, on the position

y' of the RBC body.

Correspondingly, in experiments, each RBC was accurately positioned with a micromanipulator

to have half of the cell adhered on a glass slide while the other half suspended (Figure 3.3) and a

rectangular AFM cantilever tip (MSNL-10 B: spring constant 0.023 N/m, length 210 µm, width

20 µm, thickness 0.5 µm) was used to deform the free end. The cantilever tip applied F on the

position y' of each RBC. An example force-displacement curve from AFM measurement of an

RBC is shown in Figure 3.3. The same RBC sample was tested in our microfluidic device. For

direct comparison, effective stiffness for the AFM measured RBCs was calculated as Fs/L.

The results from both groups are summarized in Table 1. It can be seen that the effective

stiffness measured by AFM is slightly higher than the microfluidic device-measured results.

Several error sources could have contributed to the difference. (1) The AFM cantilever’s spring

constant was carefully calibrated with the thermal tune method integrated in the AFM by Bruker.

The error of the calibrated spring constant is within 5%, directly reflected in the AFM applied

force. (2) The exact position of AFM cantilever tip for applying the force Fs also contains errors,

although the best experimental care was used. The position error is approximately one pixel

Page 71: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

59

(0.1613 µm/pixel), and for a 4 µm RBC, this position error y' = 2.9 ± 0.1613 µm causes an error

of 3.8% in the quantified effective stiffness of the RBC. (3) Finally, in the derivation of

equivalent loading position (Eq. (3) (4)), the force is supposed to be applied along a line (dy →

0). However, in experiments, when the cantilever tip contacts the RBC, the width of the

contacting area is small but not zero. Finite element simulation of the experimental situation

reveals that the contact width is approximately 0.13 µm, as shown in Figure 3.3. Putting this

width back to the theoretical calculation (dy = 0.13 µm), it caused an error of 3.5% in the

quantified effective stiffness of the RBC. In summary, the validation experiments contained

identifiable error sources as any measurement; however, the results support the validity of our

microfluidic measurement.

Table 3.3: Microfluidic device and AFM measured effective stiffness of RBCs (6 weeks)

Technique Microfluidic (n=220 RBCs) AFM (n=11 RBCs)

Effective stiffness 95 ± 7 µN/m 108 ± 18 µN/m

Page 72: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

60

 

Figure 3.3 Schematic force diagram. AFM experiment and force displacement curve. Contact width estimation by applying a force of 150 pN.

Page 73: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

61

3.4.2. Effective stiffness is not velocity dependent under low velocities.

To understand the dependence of deformation index (DI, defined as L/d [201]) and effective

stiffness on flow velocity, measurements were made on RBCs from a fresh sample at ambient

temperature. Figure 3.4 shows the RBCs’ DI and effective stiffness data measured at different

flow velocities. DI significantly changed (0.79 ± 0.09 vs. 1.01 ± 0.11, *p < 0.001) when the flow

velocity was increased from 0.01 m/s to 0.02 m/s. Similarly, in an ektacytometry study [208]

where RBCs were deformed in the stretching mode, the extent of RBC elongation defined as

elongation index (EI) was also found to increase at higher flow velocities. These results confirm

that the indices, DI and EI, strongly depend on flow velocity, and the use of different flow

velocities in experiments can lead to different conclusions in DI or EI change during RBC

storage. The drag force quantification in Eq. (1) takes into account the effect of flow velocity, v.

At low flow velocities when the parachute shape is sustained, both the drag force F and RBC

deformation/deflection L increase as flow velocity increases. Experimental measurement on

fresh RBCs confirmed that their effective stiffness remained unchanged under different flow

velocities (28.5 ± 8.4 µN/m; p value > 0.1) as shown in Figure 3.4(b), indicating that effective

stiffness is largely flow velocity independent under low velocities and is a more appropriate

metric for characterizing RBCs’ mechanical degradation than the phenomenological parameters

of DI and EI.

To further investigate the effect of flow velocity on RBC’s effective stiffness, 3D finite element

simulation (COMSOL) was conducted. RBC was modeled as a shell (i.e., cell membrane;

thickness: 10 nm, Young’s modulus: 1 kPa [26]) encapsulating fluids (i.e., hemoglobin, as

incompressible [27]). It was modelled as a biconcave disk (diameter: 8 µm) with a thickness at

the thickest point of 2 µm and a minimum thickness in the center of 1 µm. In simulation, flow

Page 74: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

62

velocities were varied from 0.01 m/s to 0.1 m/s in the constriction channel (12 µm × 20 µm in

cross section, the same as in experiments). Experimentally varying drag force and measuring

RBC deflection is difficult because of the practical limitation of the high-speed camera’s shutter

time for clearly measuring RBC deformation, limiting the flow velocity to < 0.03 m/s. The

simulation results reveal that the drag force increases linearly with the flow velocity; however,

RBC deflection increases linearly only when the flow velocity is lower than 0.06 m/s, after

which nonlinearity occurs. This is because the stress induced by a flow velocity higher than 0.06

m/s exceeds RBC’s yield stress (250 Pa to 300 Pa) [204][205][209]. Thus, the flow velocity <

0.03 m/s used in experiments was not considered to be sufficiently high to induce nonlinearity in

RBC effective stiffness.

Finite element simulation was also conducted to investigate the effect of microfluidic channel

width on RBC effective stiffness (flow velocity: 0.02 m/s, as used in experiments). The results

reveal that when the channel width is larger than 11 µm, the effect of channel width on RBC

effective stiffness becomes negligible. Simulation also shows (data not shown here) that when

the microfluidic channel width is larger than 14 µm, RBC deformation is only approximately 0.1

µm for 6-week old RBCs, due to the small flow-induced force, making RBC deformations

difficult to measure via imaging. Thus, a channel width of 12 µm was chosen in our microfluidic

device.

Page 75: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

63

 

Figure 3.4 Experimentally measured deformation index DI =L/d and effective stiffness values of RBCs from a fresh sample. (a) DI significantly decreases at higher flow velocities (*p < 0.001). (b) In comparison, effective stiffness of RBCs is not flow velocity dependent. Error bars represent the standard deviation.

3.4.3. Temperature effect on RBC effective stiffness

In blood banks, RBCs are stored at 4°C. We next measured the effective stiffness of stored RBCs

at different temperatures to understand temperature effect on the stiffness of stored RBCs. In the

fresh RBC group (Figure 3.5), freshly collected RBCs were stored at 4°C for 5 hours and then

diluted in PBS that had been stored at 4°C; or diluted in PBS that had been incubated at 37°C for

microfluidic measurement. After the 37°C measurements, RBCs were cooled down to ambient

temperature of 25°C and then measured. The same protocol was used for collecting data in the ‘2

weeks’ and ‘5 weeks’ groups (Figure 3.5) for RBCs stored at 4°C for two weeks and five weeks,

respectively. In all measurements, pH was maintained consistently at 7.35 ± 0.05.

As shown in Figure 3.5, within each of the three groups (fresh, ‘2 weeks’, and ‘5 weeks’, n>

2,000 RBCs), the effective stiffness of RBCs measured at 4°C was always significantly higher

Page 76: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

64

than at 25°C and 37°C. However, the effective stiffness of RBCs showed no difference when

measured at 25°C and 37°C. It is known that as temperature is increased from 4°C, the lipid tails

in the RBC membrane become unsaturated, resulting in extra free space within the lipid bilayer

[210][211]. Furthermore, as temperature increases, phosphatidylcholine lipids in the RBC

membrane turn more from a crystal-like arrangement to a liquid-like state. This transition is

largely completed at 22°C when lipid tails are fully unsaturated [210][212][213]. Our results, for

the first time, quantitatively reveal how RBCs’ effective stiffness decreases from 4°C to 25°C

and 37°C. The stiffness data supports previous findings of RBC membrane lipid packing changes

and the transition of saturation states at low to high temperatures. Additionally, data in all three

groups (fresh, ‘2 weeks’, and ‘5 weeks’) also indicated no significant difference in measuring

RBCs’ effective stiffness at 25°C and at 37°C.

Page 77: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

65

 

Figure 3.5 Effective stiffness of RBCs at different temperatures. RBCs under 4°C show higher effective stiffness. RBC stiffness measured at 25°C shows no significant difference from that measured at 37°C (*p < 0.001). Error bars represent the standard deviation.

3.4.4. RBC effective stiffness increases during storage.  

Fresh RBCs and RBCs stored up to 6 weeks were tested under ambient temperature. Based on

the measurement of over 5,000 RBCs from five different samples, the effective stiffness of RBCs

at different storage points was quantified and is summarized in Figure 3.6. RBCs stored for one

week already started to show significantly higher stiffness than fresh RBCs (one week: 37.2 ±

8.6 µN/m vs. fresh: 26.5 ± 8.3 µN/m; **p < 0.0001). No significant difference for RBCs stored

from 1 week to 3 weeks was observed until the fourth week (60.5 ± 12.3 µN/m) when the

stiffness of stored RBCs increased drastically. RBC stiffness then further increased to 79.9 ±

17.2 µN/m (5 weeks) and 86.2 ± 17.5 µN/m (6 weeks). These results indicate that stored RBC

Page 78: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

66

stiffness degrades faster in the last three weeks than in the first three weeks, and 6-week old

RBCs have an effective stiffness almost four times that of fresh RBCs (86.2 ± 17.5 µN/m vs.

26.5 ± 8.3 µN/m).

 

Figure 3.6 Effective stiffness of RBCs at different storage time points. RBCs stored for one week already started to show significantly higher stiffness than fresh RBCs. Stored RBC stiffness degraded faster in the last three weeks than in the first three weeks(*p < 0.001 and **p < 0.0001). Error bars represent the standard deviation.

Page 79: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

67

3.5. Discussion

RBC stiffness is determined by the membrane skeletal network of RBCs. In the RBC membrane,

a phospholipid bilayer is tethered to the spectrin network (mainly the spectrin α and spectrin β

proteins) via a number of proteins such as band 3, ankyrin, protein 4.1, and Glycophorin, as

shown in Figure 3.7(a). Spectrin α and spectrin β form the structure underneath cell membrane

and provide mechanical support [214][215]. Ankyrin is responsible for the bridging of spectrin to

band 3, which is one of the major RBC membrane-spanning proteins in the lipid bilayer

[216][217]. The affinity between ankyrin and spectrin is modulated by S-nitrosylation, a process

involving post-translational protein modifications [218][219]. Another protein, protein 4.1, also

plays a key role in regulating membrane stiffness by interacting with spectrin and Glycophorin,

and their interactions are modulated by phosphorylation, a process that involves the addition of

phosphate to an organic compound by consuming adenosine triphosphate (ATP) [220]. The

structure and the interactions between these proteins are changed during RBC storage due to the

degradation of several biochemical parameters such as S-nitrosothiols (SNO) and ATP

[221][222][223]. The degradation of these biochemical parameters over the RBC storage process

has been widely reported, and the depletion of SNO and ATP has been speculated to play

important roles in regulating RBCs’ mechanical properties [220][224][225][226].

During RBC storage, the SNO level becomes lower, leading to lower activity of S-nitrosylation

[221]. The detachment between spectrin and ankryin, which is induced by S-nitrosylation,

becomes less frequent, and the detached end of spectrin thermally diffuses back to ankyrin and

reattach [219][227]. Higher affinity between the spectrin and ankyrin proteins can contribute to a

higher stiffness of the RBC membrane. In the meanwhile, gradual ATP depletion is known to

Page 80: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

68

occur during RBC storage [222]. Due to insufficient energy provided by ATP for the detachment

of Glycophorin from the spectrin network [220], Glycophorin tends to re-attach to spectrin by

protein 4.1 [225], and the higher affinity between Glycophorin and spectrin can also contribute to

the increase of RBC membrane stiffness.

Our data show that the stiffness of stored RBCs increased significantly during the first week of

storage. Interestingly, this timing well matches the degradation pattern of SNO. It is known that

the SNO level decreases by 70% after only one day of RBC storage, and by the end of the first

week of storage, up to 90% SNO is depleted [221], as shown in Figure 3.7(b). Thus, we reason

that the significant depletion of SNO in the first week of RBC storage has caused the significant

RBC stiffness increase by Week 1 as our data revealed (one week: 37.2 ± 8.6 µN/m vs. fresh:

26.5 ± 8.3 µN/m). Since little SNO is left at the end of first week of storage, the effect of SNO

depletion on RBC stiffness becomes less obvious in Week 2-Week 6. Compared to Week 1-

Week 3, RBC stiffness increase was significantly more rapid in the last three weeks (Week 4-

Week 6). This faster increase in stiffness during the last three weeks’ storage can attribute to the

degradation timings of ATP. The ATP level has been shown to remain largely unchanged during

the first three weeks’ storage but start to drastically decrease in Week 4 [222][223], as shown in

Figure 3.7. It is thus likely that the increase of stored RBCs’ stiffness in Week 4-Week 6 is

mostly caused by the depletion of ATP.

Page 81: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

69

 

Figure 3.7 (a) RBC membrane skeletal network. The affinities and interactions of key proteins are regulated by biochemical parameters such as SNO and ATP, leading to RBC membrane stiffness changes. (b) SNO (blue) and ATP (red) degradation during RBC storage. Data shown here are from [30][31][32]. Approximately 90% SNO is depleted during the first week of storage, and ATP remains unchanged during the first three weeks of storage and starts to decrease significantly in Week 4.

Page 82: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

70

Whether and how RBCs change their stiffness during storage has critical relevance to patient

safety and treatment effectiveness in transfusion medicine [129][228]. RBCs over the storage

process have been mechanically characterized in both stretching and folding modes, and

different metrics for indicating RBC deformability were used [153][201][229]. Elongation index

(EI) and deformation index (DI) were phenomenologically defined and are strongly dependent

on flow velocity. Instead of measuring EI or DI, shear modulus of RBCs was measured using the

optical tweezers method to characterize stored RBC deformability changes [154]. Quantifying

the laser tweezers-applied forces was obtained by comparing the experimental data with finite

element simulation. Older RBCs were found to be stiffer than fresh ones. Unfortunately, no shear

modulus change data were presented beyond 21-day storage in [154] although 42 days is

specified as the shelf life in most countries. Recently, a deformability-based microfluidic device

was reported for sorting stiff and less stiff RBCs [230]. It was observed that significant

difference in the sorting results existed between RBCs stored less than and longer than 28 days,

indicating that RBCs became stiffer after 28-day storage. This result agrees well with our

quantified effective stiffness changes that the effective stiffness of RBCs starts to increase

drastically from Week 4 (28 days). Note that the sorting device is not capable of quantitating the

mechanical property changes of RBCs; instead, it leverages RBC stiffness changes for RBC

sorting.

Our work, for the first time, measured RBCs’ inherent stiffness change over the storage process

by deforming RBCs in the bending mode (an in vivo-like deformation mode). Effective stiffness,

a flow velocity independent parameter, was defined and used to quantify the mechanical

degradation of stored RBCs. Effective stiffness reflects the resistance of an RBC against bending

Page 83: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

71

deformation under external forces, for instance, induced by blood flow in vivo or fluidic flow in

microfluidic channels. The results reveal that RBCs stored for one week already started to show

significantly higher stiffness than fresh RBCs, and stored RBC stiffness degraded faster in the

last three weeks than in the first three weeks. These results and the interesting coincidence

between the time points of effective stiffness increase and the degradation patterns of SNO and

ATP in stored RBCs motivate us to pursue a systematic correlation between biomechanical and

biochemical parameters in RBC storage lesion; they can also likely trigger deeper analyses of

patient data gathered from previous and present large-scale clinical studies in transfusion

medicine to better understand RBC storage age and clinical results, for instance, do 1-week and

3-week old RBCs cause a mortality difference in transfused patients? There also exists some

technical issue that deserves more investigation. For instance, to better understand the effect of

biochemical degradation on RBCs’ stiffness changes, different from existing studies on RBC

populations [221][222][223], performing biochemical measurements on single RBCs and

stiffness measurements on the same RBCs would permit a more precise correlation.

3.6. Conclusion:

This chapter presented microfluidic measurement of effective stiffness changes of RBCs over the

storage process. Instead of using phenomenological metrics such as deformation index (DI) and

elongation index (EI) as in existing studies, effective stiffness that reflects RBC’s inherent

mechanical property and is flow velocity independent is used to quantitatively describe the

mechanical degradation of stored RBCs. Fresh RBCs and RBCs stored up to 6 weeks were

measured on a microfluidic device in the bending mode mimicking their deformation when they

Page 84: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

72

pass through microcapillaries in vivo. The results revealed the pattern of effective stiffness

increase and the time points where drastic stiffness increase occurred. The coincidence with the

degradation patterns of SNO and ATP was also discussed.

Page 85: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

73

Chapter 4

4. Stiffness and ATP Recovery of Stored Red Blood Cells in Human Serum

4.1. Introduction

According to the World Health Organization, 92 million units of blood are collected annually in

164 countries [231]. More than 21 million units of red blood cells (RBCs) as the main type of

blood product are transfused every year in the United States [232]. Regulations stipulated by the

US Food and Drug Administration (FDA) and many other countries specify 42 days as the shelf

life for stored RBCs [127]. This “gold standard” was established based on the criteria of post-

transfusion RBC survival of 75 % or above after 24 hours [233], where post-transfusion RBC

survival was measured by monitoring transfused RBCs labeled with 51Cr [234]. However, 25%

RBC clearance is higher by orders of magnitude than the normal daily RBC clearing ratio

(~0.8%) in the human body [235].

In transfusion medicine, there has been a decades-long debate about whether older RBCs (i.e.,

RBCs that are stored longer) can cause worse transfusion outcome than fresher ones

[236][237][238]. More than 40 clinical studies reported that the use of older RBCs was

associated with a significantly increased risk of adverse complications after transfusion

[137][138][129]. These studies examined the effect of RBC storage on several clinical outcomes

including mortality, rates of infection, and length of ICU stay in hospital [139][140][141]. The

time point to distinguish fresher and older RBCs was mostly chosen to be 14 days or 21 days

Page 86: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

74

[129][142]. In these studies, patients transfused with older RBCs showed higher rates of

mortality and infection, or longer ICU staying time than patients transfused with fresher RBCs.

Results from more recent clinical studies, on the other hand, reported no significant difference in

clinical consequences between fresher and older RBCs [144][145][146][147][148][149]. In two

larger-scale clinical trials, transfusion of fresher RBCs, as compared with older RBCs, did not

significantly reduce the complications of prematurity in very-low-birth-weight infants (patient

number 377) or reduce the rates of organ failure or adverse events among 1,098 patients

undergoing cardiac surgical procedures [150][151].

In the human body, RBCs must be transported to every tissue to deliver oxygen; otherwise,

tissue hypoxia occurs, which can result in pulmonary hypertension, stroke, and cardiovascular

dysfunction [239][240][241]. A proper stiffness must also be maintained by RBCs to allow them

to pass through capillaries in microcirculation [242]. Poorly deformable RBCs (i.e., higher

stiffness) result in a higher clearance by the spleen and are known to contribute to respiratory

distress and systemic sepsis [130][131]. Clinical research has also identified other disease

conditions such as splanchnic ischemia developed in patients who had been transfused with older

RBCs that are known to have higher stiffness than fresher ones [132]. Furthermore, among the

critically ill patients with sepsis who had older RBCs transfused, sepsis became aggravated

[134][133]. Since septic patients have constricted vessels, poorly deformable RBCs can be

trapped in microcirculation, leading to tissue hypoxia and exacerbating patients’ health

conditions [135][136].

Page 87: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

75

The storage of RBCs is known to cause cell degradation (e.g., RBC stiffness and ATP), referred

to as “storage lesion” [243][244]. For instance, mechanical stiffness changes of stored RBCs

have been widely studied [199][200][155], and the results consistently revealed that the stiffness

of RBCs increases over the storage process. However, the controversy in clinical outcome of

transfusing fresher and older RBCs suggests that the degradation of stored RBCs can possibly be

reversed, and parameters such as RBC stiffness and ATP can recover after transfusion.

Therefore, the questions that motivated the present study are as follows. Can the stiffness of

stored RBCs recover in vivo? To what extent? How does storage duration cause differences in

the recovery of stiffness? Furthermore, since ATP concentration in RBCs plays an important role

in regulating the mechanical stiffness of RBCs [244][220][222], how well and quickly can the

ATP level recover after RBC transfusion? This chapter reports microfluidic measurement of the

evolution of stiffness and ATP recovery of stored RBCs under in vivo-like conditions (i.e., in 37

C human serum) and provides quantitative evidence to these questions.

4.2. Methods

The microfluidic device, constructed via standard PDMS soft lithography, had a microchannel of

60 µm in height and 1,000 µm in width. With the flow rate of 10 µL/min used in our

experiments, the generated shear stress of 0.5 Pa is comparable to in vivo condition [174], is

sufficiently high to deform RBCs (Figure 4.1), and is far below the yield stress of RBCs (~100

Pa) [205]. RBCs (from 7 different subjects) stored for different periods of time were first diluted

in PBS by 200 times and then introduced into the microfluidic channel of the device. They

strongly adhered to the glass substrate after settling for 15 minutes [245]. The RBCs were heated

by a heating plate (37 °C, HWPT-384S) throughout experiments. Human serum (type AB, male,

Page 88: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

76

Sigma-Aldrich) of 37 °C was continuously perfused in the microchannel for 120 minutes, and

images of RBC deformation [Figure 4.1(b)] were recorded at 1 Hz. The shape of each individual

RBC was measured via image processing in Matlab, and the shear modulus of each RBC was

quantified, according to our previously reported method [203]. Briefly, according to the Kelvin–

Voigt (KV) model, 𝑇 𝜆 , where T (µN/m) is the average tension force acting on the

RBC membrane, which was calculated from the shear stress applied on the RBC and the area of

the RBC [246][176]; and λ is the extension ratio of the RBC membrane, 𝜆 , where 𝑙 is the

RBC’s length when deformed under shear stress, and 𝑙 is the RBC’s original length. RBC’s

elastic shear modulus (𝜇) expressed in force per unit length (µN/m) was determined, and higher

shear modulus indicates a higher RBC membrane stiffness. Non-stored fresh RBCs were also

tested in the experiment as a benchmark to evaluate how stored RBCs recover their properties in

37 C human serum.

Page 89: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

77

 

Figure 4.1 (a) RBCs strongly adhered to glass substrate after settling for 15 minutes. RBCs in microchannel were perfused with 37 °C human serum for 120 minutes. (b) RBCs were deformed by shear flow (flow rate of 10 µL/min). Images of RBC deformation were recorded, and the shear modulus of each RBC was quantified.

 

 

Page 90: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

78

4.3. Results

4.3.1. Stiffness recovery of stored RBCs in human serum  

PBS and human serum were both used to perfuse into the micro channel, and the stiffness change

of individual RBCs was monitored. Figure 4.3 shows the shear modulus evolution of two-week

old RBCs (i.e., stored for two weeks). During 120 minutes of perfusion, no stiffness change was

observed in the PBS-perfusion group. In the serum-perfusion group, the shear modulus of RBCs

remained unchanged at 4.3 µN/m for approximately 60 minutes and then started to decrease.

When the shear modulus decreased to approximately 2.7 µN/m, steady-state was reached, and no

further change was observed. These results confirmed that stored RBCs are capable of

recovering their shear modulus in human serum, and the recovery process was in the time scale

of tens of minutes.

 

Figure 4.2 Stiffness of RBCs perfused for 120 minutes in human serum vs. end-point measured stiffness of RBCs incubated in human serum in an incubator without perfusion (n=1300-3000

Page 91: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

79

RBCs for each condition). No difference was observed between the perfusion group and incubating group (*p>0.05), showing that shear stress from perfusion did not induce bias in RBC stiffness recovery.

One-week to six-week old RBCs (3,000 RBCs from 7 different subjects) were tested, and their

stiffness recovery after 120-min perfusion with PBS and human serum are summarized in Figure

4.3 (b). The RBCs stored for one week and two weeks recovered their shear modulus from 4.1 ±

0.5 µN/m and 4.3 ± 0.5 µN/m to 2.9 ± 0.3 µN/m and 3.2 ± 0.4 µN/m, which are about 1.1 and

1.2 times the shear modulus of fresh RBCs (2.6 ± 0.3 µN/m). However, RBCs stored for four

weeks and longer were only able to recover from 5.5 ± 0.6 µN/m to 4.2 ± 0.9 µN/m (four

weeks), from 6.2 ± 0.7 µN/m to 5.1 ± 1.1 µN/m (five weeks), and from 6.9 ± 0.6 µN/m to 5.5 ±

1.1 µN/m (six weeks). These results indicate that RBCs stored longer than three weeks have

limited capability of stiffness recovery. For instance, six-week old RBCs after recovery have a

shear modulus twice that of fresh RBCs (5.5 ± 1.1 µN/m vs. 2.6 ± 0.3 µN/m), i.e., very poor

deformability. Note that for stored RBCs of all ages (one week to six weeks), stiffness recovery

was not biased by the shear stress from perfusion (Figure 4.2).

Page 92: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

80

 

Figure 4.3 (a) Shear modulus evolution of two-week old RBCs. PBS perfusion (blue line) did not cause RBCs to recover their shear modulus. With perfusion of 37 °C human serum (red line), RBCs shear modulus remained unchanged at 4.3 µN/m for 60 minutes; however, between 70-80 minutes, the shear modulus value continuously decreased. By 90 minutes, stead-state was reached, and shear modulus became 2.7 µN/m. (b) Steady-state shear modulus values after 120-min perfusion. With human serum perfusion, RBCs stored for one week or two weeks were able to recover their shear modulus close to the level of fresh RBCs. Older RBCs (four-six weeks)

Page 93: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

81

revealed limited capability of stiffness recovery. *p <0.05&**p<0.001. Error bars represent the standard deviation. For each condition, n= 1800-3000 RBCs.

4.3.2. Fresher RBCs reaching steady-state shear modulus faster  

The time fresher and older RBCs took to reach their steady-state shear modulus was also

different. For instance, by 80 minutes, 85% of one-week old RBCs and 80% of three-week old

RBCs reached their steady-state shear modulus. In contrast, the percentage for five-week old

RBCs was only 47% [Figure 4.6 (a)]. Although the measurement was made continuously for 120

minutes, most of the stored RBCs reached their steady-state shear modulus between 60 minutes

and 90 minutes. Between 90 minutes and 120 minutes, for all the stored RBC samples (one-week

to six-week), almost no RBC revealed a change in its shear modulus. As shown in Figure 4.6 (a),

by the end of the 120-min human serum perfusion, there were still ~30% of the five-week old

RBCs showing no recovery at all. We also conducted additional experiments to continuously

perfuse stored RBCs for 8 hours (Figure 4.4), finding no further recovery of the RBCs and the

results were consistent with those obtained from the 120-min perfusion experiments. These data

show that older RBCs require longer time to reach steady-state shear modulus, and a higher

percentage of older RBCs cannot recover their stiffness in human serum.

Page 94: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

82

 

Figure 4.4 Stiffness of RBCs perfused up to 120 minutes in human serum vs. perfused up to 8 hours (n=790-3000 RBCs for each condition). No difference was observed (*p>0.05), showing that longer perfusion did not induce further recovery of shear modulus.

Page 95: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

83

4.3.3. Temperature effect on RBC stiffness recovery  

We tested the shear modulus of PBS-perfused RBCs and serum-perfused RBCs at both 25 C

(room temperature) and 37 C. As shown in Figure 4.5, when the stored RBCs were perfused

with PBS, after two hours of perfusion, no significant difference in the steady-state shear

modulus values was found between the 25 C group and the 37 C group, indicating that 37 C

alone cannot induce RBC stiffness recovery. The lipid tails in the RBC membrane become more

unsaturated as temperature increases, resulting in more free space within the lipid bilayer and

thus, a more deformable membrane [210]. At 22 C, the stiffness of the RBC membrane has

reached the steady state because the lipid tails are fully unsaturated [212]. This phenomenon is

consistent with our finding that 25 C and 37 C did not cause a statistically significant

difference in the stiffness of PBS-perfused RBCs.

Page 96: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

84

 

Figure 4.5 Steady-state shear modulus of stored RBCs that were PBS or serum perfused at 25 °C or 37 °C (n=1260-3000 RBCs for each condition). The results show that 25 °C and 37 °C did not cause a statistically significant difference in the stiffness of PBS-perfused or serum-perfused RBCs.

For serum-perfused RBCs, we also found no difference in the steady-state shear modulus

between the 25 C group and the 37 C group. However, the RBCs perfused at 37 C showed a

faster recovery process than at 25 C. Figure 3(b-d) shows the data collected on one-week, three-

week and five-week old RBCs. Higher percentages of 37 C RBCs than 25 C RBCs reached

their steady-state values by 80 minutes. This can be attributed to higher temperature-caused

faster ATP-regulated unbinding of RBC membrane proteins [247]. Our results suggest that,

compared to 25 C, 37 C does not produce an additional effect on the stiffness recovery of

stored RBCs; however, it accelerates the stiffness recovery process in the human serum

environment.

Page 97: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

85

 

Figure 4.6 (a) Higher percentages of one-week and three-week old RBCs, compared to five-week old RBCs, reached their steady-state shear modulus by 80 minutes. b-d) Compared to 25 °C, 37 °C accelerated the stiffness recovery process for stored RBCs in human serum. Blue and orange lines are the second degree polynomial fitting trend lines. Error bars represent the standard deviation.

4.3.4. Shape recovery of stored RBCs in human serum  

During storage, RBCs change their shape from biconcave to echinocyte [248]. To investigate the

shape recovery of stored RBCs (i.e., reversal from echinocytes to biconcave), RBCs were

cultured in 37 C human serum for 120 minutes. The percentage of echinocytes before and after

human serum incubation was measured, as shown in Figure 4.7. Before incubation in human

serum, the percentage of echinocytes was approximately 4% for the one-week to three-week old

RBCs. The percentage increased to 6% for the four-week RBCs, and by the end of the storage

Page 98: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

86

(i.e., six weeks) the percentage increased to about 9%, agreeing well with previously reported

results of storage lesion [248][249]. After 120 minutes of incubation in human serum, the

percentage of echinocytes for one-week to three-week old RBCs decreased to around 2% while a

higher percentage of older RBCs remained echinocytes (4%, 6%, and 7% for four-week, five-

week, and six-week old RBCs, respectively), indicating that RBCs stored for a shorter time

period are more capable of recovering their shape from echinocytes back to biconcave.

 

Figure 4.7 Before incubation in human serum (blue line), a higher percentage of echinocytes existed in old RBC samples. After 120-min incubation in human serum, the percentage of echinocytes decreased from 3.8 % to 2.1 % for one-week old RBCs, from 4.3% to 3.3% for two-week old RBCs, from 4.5% to 3.3% for three-week old RBCs, from 6.2% to 4.3% for four-week old RBCs, from 8.6% to 6.2% for five-week old RBCs, and from 9.3% to 7.3% for six-week old RBCs. For each condition, n=2200-2800 RBCs.

 

 

Page 99: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

87

4.3.5. ATP concentration recovery of stored RBCs in human serum  

RBC membrane is composed of a phospholipid bilayer tethered to the underneath spectrin

network (spectrin and spectrin ), as shown in Figure 4.8(a). RBC stiffness is dependent on the

interactions between the phospholipid bilayer and the underlying spectrin cytoskeleton [244].

The interactions are regulated by the binding between spectrin and 4.1R, a protein embedded in

the phospholipid bilayer [250][215]. The spectrin/4.1R binding is modulated by phosphorylation

which consumes intracellular ATP [220]. Luciferin technique (ATP Bioluminescent Assay kit;

Promega, Madison, Wisconsin, United States) was used to measure RBC intracellular ATP

concentration. For each sample, about 8000 RBCs (containing enough total amount of ATP to

emit light for measurement) were first washed by PBS at least three times. Then the RBCs were

lysed by 5% trichloroacetic acid (TCA) to extract the ATP in the RBCs. Tris-acetate, as a buffer

solution, was then added to neutralize the TCA. The rL/L reagent from the assay kit was added

into the ATP exacted solution and react with ATP to emit light with a wavelength of 560 nm

measured by a fluorescent camera. The light intensity was used to quantify the amount of ATP.

The assay kit contains a vial of ATP Standard (10-7 M) which was used as the reference to

quantify measured ATP in each experiment following the assay kit’s protocol. Then the average

ATP concentration with a unit of M (mol/L) was calculated from the measured amount of ATP

(with a unit of mol) and the total volume of the RBCs (each RBC has a volume of around 90 pL).

Fresh RBCs and stored RBCs (one-week to six-week) were incubated in human serum at 37 C.

For each time point during serum incubation, ~8,000 RBCs were used for quantifying their

average ATP concentration. We also measured the intracellular ATP concentration of fresh

RBCs (3.50 ± 0.09 mM) as the control group. As shown in Figure (b), one-week storage did not

Page 100: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

88

induce significant reduction of ATP concentration (3.50 ± 0.08 mM). However, for the RBCs

stored for four weeks, the ATP concentration was only 3.10 ± 0.08 mM; and for six-week old

RBCs, it was as low as 2.35 ± 0.04 mM. These results of ATP degradation are consistent with

the values previously reported on RBC storage lesion [222]. Mitochondria are the organelles that

use oxygen to synthesize ATP in most types of cells. RBCs, which do not contain mitochondria,

synthesize ATP by glycolysis, an oxygen-independent metabolic pathway that converts glucose

into lactate. Glycolysis contains ten enzyme-catalyzed reactions, and in two of these reactions

(converting 1, 3-Bisphosphoglycerate to 3-Phosphoglycerate, and converting

Phosphoenolpyruvate to Pyruvate) ATP is synthesized in the cytoplasm. It is known that RBC

storage causes a consistent decrease of glucose over the storage duration [222], resulting in

reduced glycolysis activities and thus the reduction of intracellular ATP concentration.

In human serum incubation, the increase of ATP concentration of stored RBCs, especially the

older RBCs (e.g., four to six weeks), was already apparent within the first 10 minutes of serum

incubation, and by 30 minutes their ATP concertation started to plateau. Longer incubation (8

hours) was also conducted, and it was verified that the steady-state ATP concentration of the

stored RBCs was indeed already reached by 120 minutes after which no further change occurred

(Figure S4). Our data reveal that the ATP concentration of stored RBCs increased in human

serum but cannot fully recover for older RBCs. For instance, for two-week old RBCs, serum

incubation increased their ATP concentration to full recovery, compared to that of fresh RBCs,

from 3.37 ± 0.05 mM to 3.50 ± 0.05 mM by 120 minutes. However, the steady-state ATP

concentration of six-week old RBCs after 120-min serum incubation only increased from 2.35 ±

0.04 mM to 2.74 ± 0.05 mM (vs. 3.50 ± 0.09 mM for fresh RBCs).

Page 101: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

89

 

Figure 4.8 (a) Lack of ATP increases the spectrin-membrane affinity, leading to higher RBC stiffness. When ATP is re-synthesized, spectrin-membrane binding becomes dynamic, causing the RBC to be more deformable. (b) ATP concentration in RBCs that were stored from one week to six weeks. For each data point, n=3 samples with each sample containing about 8000 RBCs. Older RBCs showed lower ATP concentration than fresh RBCs, and their ATP concentrations

Page 102: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

90

increased during the serum-incubation process. Steady-state values were reached by ~30 minutes.

 

Figure 4.9 RBCs were cultured in human serum for 8 hours. For each sample, the intracellular ATP concentration by 120 minutes and at the end of the 8-hr incubation was compared. For each condition, n=3 samples with each sample containing about 8000 RBCs. For all stored RBCs (one-week to six-week old), no difference was observed, indicating that longer incubation (>120 minutes) did not induce further re-synthesize of ATP.

4.4. Discussion

 

Despite the many clinical studies of RBC transfusion that investigated whether older RBCs cause

worse clinical outcomes, these clinical trials provided conflicting information [129][228] with

some showing that older blood is less effective [139][140][142] but with others showing no such

difference [149][150]. Up to today, there still exist strong controversies regarding whether the

age of stored RBCs is a factor in transfusion efficacy. Intuitively, fresher RBCs may function

better than older RBCs because RBC storage is known to induce biomechanical and biochemical

Page 103: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

91

degradations to RBCs, and longer storage leads to more severe degradations [199][200][155].

The controversial clinical trial results could have been biased by the vastly different conditions

of the patients involved in the clinical studies; however, the elusiveness is also due to the lack of

understanding of how well and quickly stored RBCs after transfusion can recover their key

parameters, such as stiffness and ATP concentration.

After blood donation and processing, RBCs are placed in the preservation medium for storage at

4 C. The low temperature is to keep the rate of glycolysis at a lower limit and minimize the

proliferation of bacteria that might have entered the blood unit [129]. In the preservation

medium, acid citrate functions as anticoagulant, mannitol is for preventing RBCs from

swelling/hemolysis, and a strictly protocoled amount of glucose is also added into the

preservation medium to help maintain the metabolism of RBCs [251]. Over the storage duration,

the glucose concentration significantly decreases due to glycolysis. This leads to a lower ATP

concentration in stored RBCs [222] and causes RBCs’ stiffness to increase, as shown in Figure

2(b) and Figure 5(b). It should be noted that RBC preservation protocols do not allow a high

amount of glucose to be added into the preservation medium since adding a high amount of

glucose can cause impairing complications such as the glycosylation of RBC membrane and

skeletal proteins [252].

Storage-induced degradations of RBCs can lead to the clearance of transfused RBCs in vivo. In a

healthy human being, only ~0.8% RBCs are cleared within 24 hours [253]; in contrast, a

significantly higher percentage of stored RBCs after transfusion are cleared (e.g., 25% for six-

week old RBCs) [253]. The clearance of RBCs mainly occurs in the spleen [254] which consists

Page 104: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

92

of a meshwork with many tiny pores (e.g., 5 µm) [34]. RBCs with a higher than normal stiffness

have a higher chance to be trapped within the meshwork and cleared [255]. A higher stiffness

itself has been shown to serve as a signal for macrophages to induce phagocytosis for RBC

clearance [256][257].

When stored RBCs are transfused into the human body, the major microenvironment they

encounter is the human serum. In this work, we quantitatively studied the stiffness and ATP

recovery of stored RBCs in 37 C human serum. The results showed that in 37 C human serum,

stored RBCs are able to recover their stiffness and ATP concentration to varying extents

depending on their age of storage. As summarized in Table I, one can see that for one-week old

RBCs, although the shear modulus before recovery was 1.6 times that of fresh RBCs, 97% of the

cells had their stiffness recovered in human serum to be 1.1 times that of fresh RBCs; and the

ATP concentration of one-week old RBCs after recovery showed no difference from that of fresh

RBCs. For three-week old RBCs, 89% RBCs recovered their stiffness to be 1.3 times that of

fresh RBCs; and the recovered ATP concentration was only 5% lower than that of fresh RBCs.

However, for six-week old RBCs, only about 70% of the RBCs showed stiffness recovery in

human serum; their shear modulus after recovery was still 2.1 times that of fresh RBCs; and their

ATP concentration after recovery was 25% lower than that of fresh RBCs. Overall, the results

indicate that fresher RBCs (one-week to three-week) have significantly higher capability of

stiffness and ATP recovery in human serum than older RBCs (four-week to six-week).

Page 105: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

93

Table 4.1: RBC shear modulus and ATP concentration before and after recovery

Age of the

stored RBCs

Shear modulus before

recovery (µN/m)

Shear modulus after

recovery (µN/m)

Recovered RBCs

percentage (%)

ATP concentration

before recovery (mM)

ATP concentration

after recovery (mM)

Fresh (benchmark) 2.6 ± 0.3 2.6 ± 0.3 N/A 3.50 ± 0.09 3.50 ± 0.09

One week 4.1 ± 0.5 2.9 ± 0.3 97 ± 2.2 3.50 ± 0.08 3.53 ± 0.02

Two weeks 4.3 ± 0.5 3.2 ± 0.4 95 ± 1.7 3.37 ± 0.05 3.50 ± 0.06

Three weeks 4.5 ± 0.5 3.3 ± 0.4 89 ± 8.3 3.28 ± 0.06 3.39 ± 0.02

Four weeks 5.5 ± 0.6 4.2 ± 0.9 85 ± 5.7 3.10 ± 0.08 3.30 ± 0.06

Five weeks 6.2 ± 0.7 5.1 ± 1.1 72 ± 3.8 2.76 ± 0.05 3.15 ± 0.04

Six weeks 6.9 ± 0.6 5.5 ± 1.1 70 ± 3.2 2.35 ± 0.04 2.79 ± 0.05

Our experiments showed that neither stiffness nor ATP concentration recovered when the stored

RBCs were incubated in PBS. Compared to PBS, the compositions of human serum are more

diverse [222]. For instance, human serum contains much glucose and pyruvate while these

components are absent in PBS [258]. Glucose is converted into lactate by the metabolic pathway

of glycolysis. Among the ten reactions involved in glycolysis, 1, 3-bisphosphoglycerate is

converted to 3-phosphoglycerate, and phosphoenolpyruvate is converted to pyruvate. The

subsequent conversion of pyruvate to lactate provides nicotinamide adenine dinucleotide (NAD),

which promotes one of the reactions in glycolysis, i.e., glyceraldehyde 3-phosphate conversion to

1,3-bisphosphoglyceric acid, and the molecules produced in this reaction are utilized by RBCs to

synthesize ATP [259]. Our data quantitatively revealed the recovery of ATP concentration in

stored RBCs when they were incubated in human serum. Since it is known that ATP regulates

the tension of the spectrin-membrane connection and thus RBC stiffness [244][220], it is likely

that the stiffness recovery of stored RBCs in human serum is proceeded by the recovery of ATP

concentration.

Page 106: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

94

Our experiments also revealed that the process of stiffness recovery and ATP recovery were on

the scale of tens of minutes. For instance, for one-week old RBCs, the recovered steady-state

ATP concentration and stiffness value were reached by 10 minutes and 80 minutes, respectively.

For six-week old RBCs, the recovery of ATP concentration and stiffness respectively took 30

minutes and 90 minutes. To put these numbers in context, an RBC completes the circulation for

one cycle in the human body within a minute [260]. The long recovery time (tens of minutes)

could increase the chance for stored RBCs to be cleared by the spleen. This might imply that

when feasible in RBC transfusion, pre-treatment of stored RBCs in the target patient’s serum

might help achieve a higher transfusion efficacy. Our results also show that, comparing six-week

old RBCs and three-week old RBCs, about 20% less RBCs can recover in stiffness. The

recovered six-week old RBCs had a stiffness that is 1.7 times that of recovered three-week old

RBCs, and the recovered ATP concentration was 20% lower than in three-week old RBCs. This

significantly poorer recovery capability might alert a revisit of the policy for the 42-day shelf life

of RBC storage. Admittedly, we note that this study used human serum to mimic the in vivo

environment, but the environment in the human body is more than what serum provides. For

instance, can other factors in the human body, such as adrenaline, which is absent in the serum

used in our study, speed up the recovery of stored RBCs? It is our hope that this study could

trigger the next steps of more comprehensive characterization of the recovery behaviors of stored

RBCs (e.g., 2,3DPG and SNO) and the quantitation of the in vivo recovery of stored RBCs in

transfusion medicine.

Page 107: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

95

4.5. Conclusion

This chapter reports microfluidic measurement of the evolution of stiffness and ATP recovery of

stored RBCs under in vivo-like conditions (i.e., in 37 C human serum). The results provide

quantitative evidence to the following questions. Can the stiffness and ATP of stored RBCs

recover in human serum? To what extent? How does storage duration cause differences in the

recovery of stiffness and ATP concentration? It was found that despite the degradation induced

by storage, stored RBCs were able to recover their stiffness and ATP concentration in human

serum. RBCs stored for one to three weeks were capable of recovering their stiffness and ATP

concentration close to fresh RBCs, while after the recovery older RBCs (four to six weeks) had a

significantly higher stiffness and lower ATP concentration than that of fresher RBCs. The results

also revealed that the process of stiffness recovery and ATP recovery were on the scale of tens of

minutes. These findings bring new insight to how well and quickly stored RBCs recover after

transfusion.

Page 108: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

96

Chapter 5

5. Conclusions  

RBC’s membrane accounts for most of RBC’s antigenic, transport, and mechanical

characteristics. Abnormal RBCs with compromised deformability accounts for decreased RBC

life span, and leads to anemia in different RBC disorders or RBC products. In this thesis, a

microfluidic device that allows for measuring RBC mechanical property with in vivo-like

conditions applied and adjusted, was developed and applied to characterize mechanical

properties of RBCs from sickle cell trait (SCT) carriers and the degradation/recovery of stored

RBCs in transfusion medicine. The main findings and contributions of the thesis are summarized

as following:

5.1. Contributions

 

1. Developed microfluidic systems to address the question whether RBCs of SCT

individuals become stiffened during strenuous exercise. RBCs from SCT individuals and

non-SCT subjects were tested under simulated strenuous exercise conditions (i.e., low

oxygen and low pH). The results indicate that RBCs from SCT individuals are sensitive

to low pH which induces significant stiffness increase in SCT RBCs, implying that the

stiffening of RBCs could occur in SCT individuals during strenuous exercise.

2. Designed a microfluidic device for measurement of effective stiffness changes of RBCs

over the storage process. Compared to existing techniques using phenomenological

Page 109: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

97

metrics such as deformation index (DI) and elongation index (EI), effective stiffness that

reflects RBC’s inherent mechanical property and is flow velocity independent is used to

quantitatively describe the mechanical degradation of RBCs during the blood storage.

The results show that stored RBC stiffness degraded faster in the last three weeks than in

the first three weeks.

3. Characterized stored RBCs stiffness recovery by using a microfluidic device. Stiffness

evolution of RBCs stored up to 6 weeks was measured under an in vivo-like condition

(37 °C human serum perfusion). The results revealed the storage time’s effect on stored

RBCs’ steady-state stiffness. The time required for RBCs to recover also differs between

RBC samples stored for various weeks. ATP, as a crucial organic chemical that provides

energy to keep RBC membrane mechanically functional, was also measured with RBCs

treated by human serum.

In conclusion, this thesis introduced the mechanical characterization of RBCs using microfluidic

devices, and also provided knowledge to understand RBC property changes in pathological

condition such as sickle cell trait and storage conditions. The results showed that although sickle

cell trait RBCs do not sickle or become stiffer under low oxygen condition, lower pH during

strenuous exercise could lead to the stiffening of RBCs. Blood storage condition was also studied

and the results revealed that longer stored RBCs have higher stiffness, which could be partially

recovered in in vivo-like condition, and the findings bring new insight to how well and quickly

stored RBCs recover after transfusion.

The devices used in this thesis are also easy to be implemented in a clinical setting. After the SU-

8 layer is fabricated, as discussed in the device fabrication section, the PDMS curing and

Page 110: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

98

bonding can be easily done within few hours. It takes only one microscope and image processing

software to test the stiffness of RBCs. This can be used to perform a quick screening of RBCs’

mechanical properties before the blood transfusion. Meanwhile, the devices used in this thesis

allow the monitoring of single RBCs under different micro environment, which could be

implemented in clinical setting for drug testing.

Page 111: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

99

5.2. Future directions

5.2.1. RBC modulus measurement in bending mode  

In this thesis, we measured the effective stiffness of RBCs which reflects the resistance of an

RBC against bending deformation under external forces. However, studies are still limited

where systematic approaches are taken to investigate the cell membrane modulus along with

viscosity when the RBCs are deformed in bending mode. To further extract the modulus and

viscosity information, a thorough model of the dynamic process of an RBC passing through a

capillary-like channel needs to be built and applied to the experiments. In future studies,

elastic spring model can be used to represent the RBC membrane [261]. In this model, the

skeleton structure of an RBC membrane can be simulated by 80 to 120 (depending on the

experiment sensitivity) mass particles, which are interconnected by springs. I expect by

correlating simulation and experimental result, an RBC membrane modulus can be extracted.

Furthermore, by monitoring the dynamic shape changes when an RBC enters/leaves the

channel, we can also use the spring model to determine RBC’s viscosity [262].

5.2.2. Single RBC biochemical properties measurement  

In this thesis, we measured the ATP changes of RBCs during the storage as well as during

the recovery. Although population-based ATP measurements are able to assess the relative

ATP change of RBCs and show the recovery of ATP, they also mask the behavior of

minority subpopulations and the differences between single cells. Single RBC ATP

measurement is required to understand the heterogeneity in an RBC sample. What is needed

Page 112: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

100

now are effective ways to isolate and process large numbers of individual RBCs for ATP

measurements. This requires cell isolation under uniform conditions. One possible strategy is

droplet microfluidics, which is an effective method to isolate and analyze thousands of

individual cells for ATP measurement [263][264]. 2,3-DPG and SNOs are also crucial

biochemical properties for RBC to perform functionally in human bodies, for example

delivering oxygen [265], and single RBC measurement of 2,3-DPG and SNOs can also be

used to assess the quality of stored RBCs. 2,3-DPG concentration of RBCs can be

determined by using the detection kit from Roche Diagnostics (Mannheim, Germany), and

the SNOs can be selectively converted to NO which can be detected as the chemiluminescent

product of its reaction with ozone [266].

5.2.3. In vivo study of stored RBC recovery  

We have studied the stored RBCs recovery under in vivo-like conditions (i.e., in 37 C

serum) in this thesis; however, we note that this study used serum to mimic the in vivo

environment, but the body environment is more complicated than serum [267]. To study the

recovery of stored RBCs in an in vivo condition, animal tests should be considered in the

future studies. Both swine and sheep have proven to act as good test subjects for RBC studies

in transfusion medicine [268][269]. To track the RBCs in animal subjects, stored RBCs can

be labeled with 51Cr [234] before the transfusion. When the blood sample is collected after a

certain time of circulation in animal subjects, labeled RBCs can be sorted by flow cytometry

[270] for further biophysical and biochemical properties study.

Page 113: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

101

5.2.4. In vitro production of RBCs  

RBC donation and transfusion have been used for various medical conditions to replace patients’

lost RBCs. However, a number of problems persist, including insufficiency of supply and the

debate about whether older RBCs (i.e., RBCs that are stored longer) are inferior than fresher

ones, as discussed in this thesis. It would be valuable if mature RBCs can be produced in vitro

rather than collected from donors. Immature hematopoietic progenitor cells have shown to be

able to develop into mature RBCs [271]. However, present in vitro produced RBCs still contain

large number of nucleated RBCs which can cause harm to human body. In future work, adjusting

the enucleation medium or co-culturing with macrophages [255][272] may have the potential to

further promote the enucleation of RBCs in in vitro production studies.

Page 114: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

102

6. Bibliography:

[1] E. Bianconi, A. Piovesan, F. Facchin, A. Beraudi, R. Casadei, F. Frabetti, L. Vitale, M. C.

Pelleri, S. Tassani, F. Piva, S. Perez-Amodio, P. Strippoli, and S. Canaider, “An

estimation of the number of cells in the human body,” Ann. Hum. Biol., vol. 40, no. 6, pp.

463–471, 2013.

[2] P. Ji, S. R. Jayapal, and H. F. Lodish, “Enucleation of cultured mouse fetal erythroblasts

requires Rac GTPases and mDia2,” Nat. Cell Biol., vol. 10, no. 3, pp. 314–321, 2008.

[3] N. Mohandas and P. G. Gallagher, “Red cell membrane : past , present , and future,”

Blood, vol. 112, no. 10, pp. 3939–3948, 2008.

[4] A. J. Verkleij, R. F. A. Zwaal, B. Roelofsen, P. Comfurius, D. Kastelijn, and L. L. M. van

Deenen, “The asymmetric distribution of phospholipids in the human red cell membrane.

A combined study using phospholipases and freeze-etch electron microscopy,” BBA -

Biomembr., vol. 323, no. 2, pp. 178–193, 1973.

[5] By Robert F.A. Zwaal and Alan J. Schroit M, “Pathophysiologic implications of

membrane phospholipid asymmetry in blood cells.,” Blood, vol. 93, no. 7, pp. 2143–2148,

1997.

[6] D. L. Daleke, “Regulation of phospholipid asymmetry in the erythrocyte membrane,”

Curr. Opin. Hematol., vol. 15, no. 3, pp. 191–195, 2008.

[7] E. Soupene and F. A. Kuypers, “Identification of an erythroid ATP-dependent

aminophospholipid transporter,” Br. J. Haematol., vol. 133, no. 4, pp. 436–438, 2006.

Page 115: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

103

[8] D. Kamp and C. W. m Haest, “Evidence for a role of the multidrug resistance protein

(MRP) in the outward translocation of NBD-phospholipids in the erythrocyte membrane,”

Biochim. Biophys. Acta - Biomembr., vol. 1372, no. 1, pp. 91–101, 1998.

[9] F. Basse, J. G. Stout, P. J. Sims, and T. Wiedmer, “Isolation of an erythrocyte membrane

protein that mediates calcium-dependent transbilayer movement of phospholipid,” J. Biol.

Chem., vol. 271, no. 29, pp. 17205–17210, 1996.

[10] B. Verhoven, “Mechanisms of phosphatidylserine exposure, a phagocyte recognition

signal, on apoptotic T lymphocytes,” J. Exp. Med., vol. 182, no. 5, pp. 1597–1601, 1995.

[11] F. A. Kuypers, J. Yuan, R. A. Lewis, L. M. Snyder, C. R. Kiefer, A. Bunyaratvej, S.

Fucharoen, L. Ma, L. Styles, K. de Jong, and S. L. Schrier, “Membrane phospholipid

asymmetry in human thalassemia.,” Blood, vol. 91, no. 8, pp. 3044–51, 1998.

[12] B. L. Wood, D. F. Gibson, and J. F. Tait, “Increased erythrocyte phosphatidylserine

exposure in sickle cell disease: flow-cytometric measurement and clinical associations.,”

Blood, vol. 88, no. 5, pp. 1873–1880, 1996.

[13] Z. Yasin, S. Witting, M. B. Palascak, C. H. Joiner, D. L. Rucknagel, and R. S. Franco,

“Phosphatidylserine externalization in sickle red blood cells: Associations with cell age,

density, and hemoglobin F,” Blood, vol. 102, no. 1, pp. 365–370, 2003.

[14] B. N. Yamaja Setty, S. Kulkarni, and M. J. Stuart, “Role of erythrocyte phosphatidylserine

in sickle red cell-endothelial adhesion,” Blood, vol. 99, no. 5, pp. 1564–1571, 2002.

[15] S. Manno, Y. Takakuwa, and N. Mohandas, “Identification of a functional role for lipid

Page 116: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

104

asymmetry in biological membranes: Phosphatidylserine-skeletal protein interactions

modulate membrane stability,” Proc. Natl. Acad. Sci., vol. 99, no. 4, pp. 1943–1948, 2002.

[16] X. An, G. Debnath, X. Guo, S. Liu, S. E. Lux, A. Baines, W. Gratzer, and N. Mohandas,

“Identification and functional characterization of protein 4.1R and actin-binding sites in

erythrocyte β spectrin: Regulation of the interactions by phosphatidylinositol-4,5-

bisphosphate,” Biochemistry, vol. 44, no. 31, pp. 10681–10688, 2005.

[17] M. E. Reid and N. Mohandas, “Red blood cell blood group antigens: structure and

function.,” Semin. Hematol., vol. 41, no. 2, pp. 93–117, Apr. 2004.

[18] D. J. Anstee, “Blood group antigens defined by the amino acid sequences of red cell

surface proteins.,” Transfus. Med., vol. 5, no. 1, pp. 1–13, Mar. 1995.

[19] V. Nicolas, C. Le Van Kim, P. Gane, C. Birkenmeier, J. P. Cartron, Y. Colin, and I.

Mouro-Chanteloup, “Rh-RhAG/ankyrin-R, a new interaction site between the membrane

bilayer and the red cell skeleton, is impaired by Rhnull-associated mutation,” J. Biol.

Chem., vol. 278, no. 28, pp. 25526–25533, 2003.

[20] W. Dc, B. M. E. Reid, Y. Takakuwa, J. Conboy, G. Tchernia, and N. Mohandas,

“Glycophorin C content of human erythrocyte membrane is regulated by protein 4.1,” pp.

2229–2234, 2013.

[21] M. Salomao, X. Zhang, Y. Yang, S. Lee, J. H. Hartwig, J. A. Chasis, N. Mohandas, and X.

An, “Protein 4.1R-dependent multiprotein complex: New insights into the structural

organization of the red blood cell membrane,” Proc. Natl. Acad. Sci. U. S. A., vol. 105, no.

23, pp. 8026–8031, 2008.

Page 117: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

105

[22] A. A. Khan, T. Hanada, M. Mohseni, J. J. Jeong, L. Zeng, M. Gaetani, D. Li, B. C. Reed,

D. W. Speicher, and A. H. Chishti, “Dematin and adducin provide a novel link between

the spectrin cytoskeleton and human erythrocyte membrane by directly interacting with

glucose transporter-1,” J. Biol. Chem., vol. 283, no. 21, pp. 14600–14609, 2008.

[23] L. J. Bruce, R. Beckmann, M. L. Ribeiro, L. L. Peters, J. A. Chasis, J. Delaunay, N.

Mohandas, D. J. Anstee, and M. J. A. Tanner, “A band 3-based macrocomplex of integral

and peripheral proteins in the RBC membrane,” Blood, vol. 101, no. 10, pp. 4180–4188,

2003.

[24] V. Bennett and A. J. Baines, “Spectrin and Ankyrin-Based Pathways: Metazoan

Inventions for Integrating Cells Into Tissues,” Physiol. Rev., vol. 81, no. 3, pp. 1353–1392,

2001.

[25] N. Mohandas and X. An, “New insights into function of red cell membrane proteins

and their interaction with spectrin-based membrane skeleton,” Transfus. Clin. Biol., vol.

13, no. 1–2 SPEC. ISS., pp. 29–30, 2006.

[26] S. E. L. Iv, G. V. Bennett, D. Branton, L. Bruce, J. Delaunay, D. Discher, V. Fowler, P.

Gallagher, W. Gratzer, V. Marchesi, N. Mohandas, J. Morrow, and J. Palek, “Review

Article Anatomy of the red cell membrane skeleton : unanswered questions,” Blood, vol.

127, no. 2, pp. 187–200, 2016.

[27] C. Korsgren and S. E. Lux, “The carboxyterminal EF domain of erythroid ??-spectrin is

necessary for optimal spectrin-actin binding,” Blood, vol. 116, no. 14, pp. 2600–2607,

2010.

Page 118: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

106

[28] X. An, X. Guo, H. Sum, J. Morrow, W. Gratzer, and N. Mohandas, “Phosphatidylserine

Binding Sites in Erythroid Spectrin: Location and Implications for Membrane Stability,”

Biochemistry, vol. 43, no. 2, pp. 310–315, 2004.

[29] S.-C. Liu, P. Windisch, S. Kim, and J. Palek, “Oligomeric states of spectrin in normal

erythrocyte membranes: Biochemical and electron microscopic studies,” Cell, vol. 37, no.

2, pp. 587–594, Jun. 1984.

[30] X. An, M. Christine Lecomte, J. A. Chasis, N. Mohandas, and W. Gratzer, “Shear-

response of the spectrin dimer-tetramer equilibrium in the red blood cell membrane,” J.

Biol. Chem., vol. 277, no. 35, pp. 31796–31800, 2002.

[31] X. An, E. Gauthier, X. Zhang, X. Guo, D. J. Anstee, N. Mohandas, and J. A. Chasis,

“Adhesive activity of Lu glycoproteins is regulated by interaction with spectrin,” Blood,

vol. 112, no. 13, pp. 5212–5218, 2008.

[32] S. Huang, A. Amaladoss, M. Liu, H. Chen, R. Zhang, P. R. Preiser, M. Dao, and J. Han,

“In vivo splenic clearance correlates with in vitro deformability of red blood cells from

Plasmodium yoelii-infected mice,” Infect. Immun., vol. 82, no. 6, pp. 2532–2541, 2014.

[33] H. A. del Portillo, M. Ferrer, T. Brugat, L. Martin-Jaular, J. Langhorne, and M. V. G.

Lacerda, “The role of the spleen in malaria,” Cell. Microbiol., vol. 14, no. 3, pp. 343–355,

2012.

[34] H. Sarin, “Physiologic upper limits of pore size of different blood capillary types and

another perspective on the dual pore theory of microvascular permeability,” J. Angiogenes.

Res., vol. 2, no. 1, pp. 1–19, 2010.

Page 119: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

107

[35] D. T. Reid H, Barnes A, Lock P, Dormandy J, “A simple method for measuring

erythrocyte deformability,” J. Clin. Pathol., vol. 17, pp. 194–195, 1964.

[36] C. Hemorheology, “Effect of pore diameter and cell volume on erythrocyte filterability,”

no. c, pp. 449–461, 1985.

[37] C. Hemorheology, “Evaluation of Leucocyte Removal Methods for Studies of

Erythrocyte,” vol. 5, no. c, pp. 137–147, 1985.

[38] M. I. Gregersen, C. A. Bryant, W. E. Hammerle, S. Usami, and S. Chien, “Flow

Characteristics of Human Erythrocytes through Polycarbonate Sieves,” Science (80-. ).,

vol. 157, no. 3790, pp. 825–827, Aug. 1967.

[39] B. Turczynski, K. Michalska-Malecka, L. Slowinska, S. Szczesny, and W. Romaniuk,

“Correlations between the severity of retinopathy in diabetic patients and whole blood and

plasma viscosity,” Clin Hemorheol Microcirc, vol. 29, no. 2, pp. 129–137, 2003.

[40] J. D. Martin, J. N. Marhefka, K. B. Migler, and S. D. Hudson, “Interfacial rheology

through microfluidics,” Adv. Mater., vol. 23, no. 3, pp. 426–432, 2011.

[41] A. M. Forsyth, J. Wan, W. D. Ristenpart, and H. a. Stone, “The dynamic behavior of

chemically ‘stiffened’ red blood cells in microchannel flows,” Microvasc. Res., vol. 80, no.

1, pp. 37–43, 2010.

[42] E. S. Park, J. P. Yan, R. A. Ang, H. Lee, X. Deng, S. P. Duffy, K. Beja, M. Annala, P. C.

Black, K. N. Chi, A. W. Wyatt, and H. Ma, “Isolation and genome sequencing of

individual circulating tumor cells using hydrogel encapsulation and laser capture

Page 120: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

108

microdissection †,” Lab Chip, 2018.

[43] S. S. Shevkoplyas, T. Yoshida, S. C. Gifford, and M. W. Bitensky, “Direct measurement

of the impact of impaired erythrocyte deformability on microvascular network perfusion

in a microfluidic device,” Lab Chip, vol. 6, no. 7, p. 914, 2006.

[44] T. Ye, H. Li, and K. Y. Lam, “Modeling and simulation of microfluid effects on

deformation behavior of a red blood cell in a capillary.,” Microvasc. Res., vol. 80, no. 3,

pp. 453–63, Dec. 2010.

[45] Q. Guo, S. P. Duffy, K. Matthews, X. Deng, A. T. Santoso, E. Islamzada, and H. Ma,

“Deformability based sorting of red blood cells improves diagnostic sensitivity for malaria

caused by Plasmodium falciparum,” Lab Chip, vol. 16, pp. 645–654, 2016.

[46] D. K. Wood, a. Soriano, L. Mahadevan, J. M. Higgins, and S. N. Bhatia, “A Biophysical

Indicator of Vaso-occlusive Risk in Sickle Cell Disease,” Sci. Transl. Med., vol. 4, no. 123,

p. 123ra26-123ra26, 2012.

[47] J. Kim, H. Lee, and S. Shin, “Advances in the measurement of red blood cell

deformability: A brief review,” J. Cell. Biotechnol., vol. 1, no. 1, pp. 63–79, 2015.

[48] E. a Evans and P. L. La Celle, “Intrinsic material properties of the erythrocyte membrane

indicated by mechanical analysis of deformation.,” Blood, vol. 45, no. 1, pp. 29–43, 1975.

[49] R. M. Hochmuth, “Micropipette aspiration of living cells.,” J. Biomech., vol. 33, no. 1, pp.

15–22, Jan. 2000.

[50] E. Evans, N. Mohandas, and a. Leung, “Static and dynamic rigidities of normal and sickle

Page 121: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

109

erythrocytes. Major influence of cell hemoglobin concentration,” J. Clin. Invest., vol. 73,

no. 2, pp. 477–488, 1984.

[51] E. a Evans, “Bending elastic modulus of red blood cell membrane derived from buckling

instability in micropipet aspiration tests.,” Biophys. J., vol. 43, no. 1, pp. 27–30, 1983.

[52] Q. Guo, S. J. Reiling, P. Rohrbach, and H. Ma, “Microfluidic biomechanical assay for red

blood cells parasitized by Plasmodium falciparum,” Lab Chip, vol. 12, no. 6, p. 1143,

2012.

[53] Q. Guo, S. Park, and H. Ma, “Microfluidic micropipette aspiration for measuring the

deformability of single cells,” Lab Chip, vol. 12, no. 15, p. 2687, 2012.

[54] Y. Zheng, J. Nguyen, Y. Wei, and Y. Sun, “Recent advances in microfluidic techniques

for single-cell biophysical characterization.,” Lab Chip, vol. 13, no. 13, pp. 2464–83, Jul.

2013.

[55] G. Binnig and C. F. Quate, “Atomic Force Microscope,” Phys. Rev. Lett., vol. 56, no. 9,

pp. 930–933, 1986.

[56] A. L. Weisenhorn, S. Kasas, J. M. Solletti, M. Khorsandi, V. Gotzos, D. U. Roemer, and

G. P. Lorenzi, “Deformation and height anomaly of soft surfaces studied with an AFM,” p.

26, 1993.

[57] M. Heuberger, G. Dietler, L. Schlapbach, and et al., “Mapping the local Young’s modulus

by analysis of the elastic deformations occurring in atomic force microscopy,”

Nanotechnology, vol. 6, no. 1, pp. 12–23, 1995.

Page 122: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

110

[58] E. A-Hassan, W. F. Heinz, M. D. Antonik, N. P. D’Costa, S. Nageswaran, C.-A.

Schoenenberger, and J. H. Hoh, “Relative microelastic mapping of living cells by atomic

force microscopy,” Biophys. J., vol. 74, no. 3, pp. 1564–1578, 1998.

[59] R. Matzke, K. Jacobson, and M. Radmacher, “Direct, high-resolution measurement of

furrow stiffening during division of adherent cells,” Nat. Cell Biol., vol. 3, no. 6, pp. 607–

610, 2001.

[60] J. L. MacIaszek, B. Andemariam, and G. Lykotrafitis, “Microelasticity of red blood cells

in sickle cell disease,” J. Strain Anal. Eng. Des., vol. 46, no. 5, pp. 368–379, 2011.

[61] I. Dulińska, M. Targosz, W. Strojny, M. Lekka, P. Czuba, W. Balwierz, and M.

Szymoński, “Stiffness of normal and pathological erythrocytes studied by means of

atomic force microscopy.,” J. Biochem. Biophys. Methods, vol. 66, no. 1–3, pp. 1–11, Mar.

2006.

[62] M. Fornal, M. Lekka, G. Pyka-Fościak, K. Lebed, T. Grodzicki, B. Wizner, and J. Styczeń,

“Erythrocyte stiffness in diabetes mellitus studied with atomic force microscope,” Clin.

Hemorheol. Microcirc., vol. 35, no. 1–2, pp. 273–276, 2006.

[63] A. Ashkin, “Acceleration and Trapping of Particles by Radiation Pressure,” Phys. Rev.

Lett., vol. 24, no. 4, pp. 156–159, 1970.

[64] D. G. Grier, “A revolution in optical manipulation,” Nature, vol. 424, no. 6950, pp. 810–

816, 2003.

[65] S.-H. Lee and D. G. Grier, “Holographic microscopy of holographically trapped three-

Page 123: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

111

dimensional structures,” Opt. Express, vol. 15, no. 4, p. 1505, 2007.

[66] G. Lenormand and A. Richert, “A New Determination of the Shear Modulus of the

Human Erythrocyte Membrane Using Optical Tweezers,” Biophys. J., vol. 76, no.

February, pp. 1145–1151, 1999.

[67] M. Dao, C. T. Lim, and S. Suresh, “Mechanics of the human red blood cell deformed by

optical tweezers,” J. Mech. Phys. Solids, vol. 51, no. 11–12, pp. 2259–2280, 2003.

[68] C. T. Lim, M. Dao, S. Suresh, C. H. Sow, and K. T. Chew, “Large deformation of living

cells using laser traps,” Acta Mater., vol. 52, no. 7, pp. 1837–1845, 2004.

[69] M. Puig-de-Morales-Marinkovic, K. T. Turner, J. P. Butler, J. J. Fredberg, and S. Suresh,

“Viscoelasticity of the human red blood cell.,” Am. J. Physiol. Cell Physiol., vol. 293, no.

2, pp. C597-605, Aug. 2007.

[70] M. Marinkovic, M. Diez-Silva, I. Pantic, J. J. Fredberg, S. Suresh, and J. P. Butler,

“Febrile temperature leads to significant stiffening of Plasmodium falciparum parasitized

erythrocytes,” Am. J. Physiol. Physiol., vol. 296, no. 1, pp. C59–C64, 2009.

[71] R. Waugh and E. a Evans, “Thermoelasticity of red blood cell membrane.,” Biophys. J.,

vol. 26, no. 1, pp. 115–131, 1979.

[72] G. Popescu, Quantitative phase imaging of cells and tissues. McGraw-Hill, 2011.

[73] G. Popescu, T. Ikeda, C. A. Best, K. Badizadegan, R. R. Dasari, and M. S. Feld,

“Erythrocyte structure and dynamics quantified by Hilbert phase microscopy,” J. Biomed.

Opt., vol. 10, no. 6, p. 060503, 2005.

Page 124: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

112

[74] Y. Y. Park, C. a. Best, T. Kuriabova, M. L. M. Henle, M. S. Feld, A. J. Levine, and G.

Popescu, “Measurement of the nonlinear elasticity of red blood cell membranes,” Phys.

Rev. E, vol. 83, pp. 1–11, 2011.

[75] Y. Park, M. Diez-Silva, G. Popescu, G. Lykotrafitis, W. Choi, M. S. Feld, and S. Suresh,

“Refractive index maps and membrane dynamics of human red blood cells parasitized by

Plasmodium falciparum.,” Proc. Natl. Acad. Sci. U. S. A., vol. 105, no. 37, pp. 13730–

13735, 2008.

[76] S. Suresh, J. Spatz, J. P. Mills, a. Micoulet, M. Dao, C. T. Lim, M. Beil, and T.

Seufferlein, “Connections between single-cell biomechanics and human disease states:

Gastrointestinal cancer and malaria,” Acta Biomater., vol. 1, no. 1, pp. 15–30, 2005.

[77] J. Guck, S. Schinkinger, B. Lincoln, F. Wottawah, S. Ebert, M. Romeyke, D. Lenz, H. M.

Erickson, R. Ananthakrishnan, D. Mitchell, J. Käs, S. Ulvick, and C. Bilby, “Optical

deformability as an inherent cell marker for testing malignant transformation and

metastatic competence.,” Biophys. J., vol. 88, no. 5, pp. 3689–98, May 2005.

[78] J. D. Pajerowski, K. N. Dahl, F. L. Zhong, P. J. Sammak, and D. E. Discher, “Physical

plasticity of the nucleus in stem cell differentiation DEVELOPMENTAL,” vol. 1, pp. 0–5,

2007.

[79] O. Otto, P. Rosendahl, A. Mietke, S. Golfier, C. Herold, D. Klaue, S. Girardo, S. Pagliara,

A. Ekpenyong, A. Jacobi, M. Wobus, N. Töpfner, U. F. Keyser, J. Mansfeld, E. Fischer-

Friedrich, and J. Guck, “Real-time deformability cytometry: on-the-fly cell mechanical

phenotyping,” Nat. Methods, vol. 12, no. 3, pp. 199–202, Mar. 2015.

Page 125: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

113

[80] J. P. Shelby, J. White, K. Ganesan, P. K. Rathod, and D. T. Chiu, “A microfluidic model

for single-cell capillary obstruction by Plasmodium falciparum-infected erythrocytes.,”

Proc. Natl. Acad. Sci. U. S. A., vol. 100, no. 25, pp. 14618–22, Dec. 2003.

[81] Y. Zheng, E. Shojaei-Baghini, A. Azad, C. Wang, and Y. Sun, “High-throughput

biophysical measurement of human red blood cells,” Lab Chip, vol. 12, no. 14, p. 2560,

2012.

[82] M. J. Rosenbluth, W. a Lam, and D. a Fletcher, “Analyzing cell mechanics in hematologic

diseases with microfluidic biophysical flow cytometry.,” Lab Chip, vol. 8, no. 7, pp.

1062–1070, 2008.

[83] D. R. Gossett, H. T. K. Tse, S. a Lee, Y. Ying, A. G. Lindgren, O. O. Yang, J. Rao, A. T.

Clark, and D. Di Carlo, “Hydrodynamic stretching of single cells for large population

mechanical phenotyping.,” Proc. Natl. Acad. Sci. U. S. A., vol. 109, no. 20, pp. 7630–5,

May 2012.

[84] O. K. Baskurt, D. Gelmont, and H. J. Meiselman, “Red blood cell deformability in sepsis.,”

Am. J. Respir. Crit. Care Med., vol. 157, no. 2, pp. 421–427, 1998.

[85] O. K. Baskurt, D. Gelmont, and H. J. Meiselman, “Red blood cell deformability in sepsis.,”

Am. J. Respir. Crit. Care Med., vol. 157, no. 2, pp. 421–7, 1998.

[86] A. G. Moutzouri, A. T. Skoutelis, C. A. Gogos, Y. F. Missirlis, and G. M. Athanassiou,

“Red blood cell deformability in patients with sepsis: a marker for prognosis and

monitoring of severity.,” Clin. Hemorheol. Microcirc., vol. 36, no. 4, pp. 291–9, 2007.

Page 126: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

114

[87] J. M. B. Pöschl, C. Leray, P. Ruef, J. P. Cazenave, and O. Linderkamp, “Endotoxin

binding to erythrocyte membrane and erythrocyte deformability in human sepsis and in

vitro,” Crit. Care Med., vol. 31, no. 3, pp. 924–928, 2003.

[88] G. P. Zaloga, D. Washburn, K. W. Black, and R. Prielipp, “Human sepsis increases

lymphocyte intracellular calcium.,” Crit. Care Med., vol. 21, no. 2, pp. 196–202, Feb.

1993.

[89] Y. I. Cho, M. P. Mooney, and D. J. Cho, “Hemorheological disorders in diabetes mellitus,”

J. Diabetes Sci. Technol., vol. 2, no. 6, pp. 1130–1138, 2008.

[90] C. Le Devehat, T. Khodabandehlou, and M. Vimeux, “Impaired hemorheological

properties in diabetic patients with lower limb arterial ischaemia.,” Clin. Hemorheol.

Microcirc., vol. 25, no. 2, pp. 43–8, 2001.

[91] D. E. McMillan, N. G. Utterback, and J. La Puma, “Reduced erythrocyte deformability in

diabetes.,” Diabetes, vol. 27, no. 9, pp. 895–901, Sep. 1978.

[92] E. Cecchin, S. De Marchi, G. Panarello, and V. De Angelis, “Rheological abnormalities of

erythrocyte deformability and increased glycosylation of hemoglobin in the nephrotic

syndrome,” Am. J. Nephrol., vol. 7, no. 1, pp. 18–21, 1987.

[93] S. Zimny, F. Dessel, M. Ehren, M. Pfohl, and H. Schatz, “Early detection of

microcirculatory impairment in diabetic patients with foot at risk.,” Diabetes Care, vol. 24,

no. 10, pp. 1810–4, Oct. 2001.

[94] C. Le Devehat, T. Khodabandehlou, and M. Vimeux, “Relationship between

Page 127: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

115

hemorheological and microcirculatory abnormalities in diabetes mellitus.,” Diabete

Metab., vol. 20, no. 4, pp. 401–4.

[95] M. Brownlee and A. Cerami, “The Biochemistry of the Complications of Diabetes

Mellitus,” Annu. Rev. Biochem., vol. 50, no. 1, pp. 385–432, Jun. 1981.

[96] C. Watala, H. Witas, L. Olszowska, and W. Piasecki, “The association between

erythrocyte internal viscosity, protein non-enzymatic glycosylation and erythrocyte

membrane dynamic properties in juvenile diabetes mellitus.,” Int. J. Exp. Pathol., vol. 73,

no. 5, pp. 655–63, Oct. 1992.

[97] C. D. Brown, H. S. Ghali, Z. Zhao, L. L. Thomas, and E. A. Friedman, “Association of

reduced red blood cell deformability and diabetic nephropathy,” Kidney Int., vol. 67, no. 1,

pp. 295–300, 2005.

[98] S. Shin, Y. Ku, M.-S. Park, and J.-S. Suh, “Slit-flow ektacytometry: Laser diffraction in a

slit rheometer,” Cytom. Part B Clin. Cytom., vol. 65B, no. 1, pp. 6–13, May 2005.

[99] S. Shin, J. X. Hou, J. S. Suh, and M. Singh, “Validation and application of a microfluidic

ektacytometer (RheoScan-D) in measuring erythrocyte deformability.,” Clin. Hemorheol.

Microcirc., vol. 37, no. 4, pp. 319–28, Jan. 2007.

[100] N. Tangpukdee, C. Duangdee, P. Wilairatana, and S. Krudsood, “Malaria diagnosis: A

brief review,” Korean J. Parasitol., vol. 47, no. 2, pp. 93–102, 2009.

[101] A. M. Vaughan, A. S. I. Aly, and S. H. I. Kappe, “Malaria Parasite Pre-Erythrocytic Stage

Infection: Gliding and Hiding,” Cell Host Microbe, vol. 4, no. 3, pp. 209–218, Sep. 2008.

Page 128: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

116

[102] C. Coban, M. S. J. Lee, and K. J. Ishii, “Tissue-specific immunopathology during malaria

infection,” Nat. Rev. Immunol., vol. 18, no. 4, pp. 266–278, 2018.

[103] L. Tilley, M. W. A. Dixon, and K. Kirk, “The Plasmodium falciparum-infected red blood

cell,” Int. J. Biochem. Cell Biol., vol. 43, no. 6, pp. 839–842, 2011.

[104] L. Sharma and G. Shukla, “Placental Malaria: A New Insight into the Pathophysiology.,”

Front. Med., vol. 4, p. 117, 2017.

[105] P. F. Mens, E. C. Bojtor, and H. D. F. H. Schallig, “Molecular interactions in the placenta

during malaria infection,” Eur. J. Obstet. Gynecol. Reprod. Biol., vol. 152, no. 2, pp. 126–

132, Oct. 2010.

[106] M. Diez-Silva, M. Dao, J. Han, C.-T. Lim, and S. Suresh, “Shape and Biomechanical

Characteristics of Human Red Blood Cells in Health and Disease.,” MRS Bull., vol. 35, no.

5, pp. 382–388, 2010.

[107] B. M. Cooke, N. Mohandas, and R. L. Coppel, “The malaria-infected red blood cell:

structural and functional changes.,” Adv. Parasitol., vol. 50, pp. 1–86, 2001.

[108] X. Pei, X. An, X. Guo, M. Tarnawski, R. Coppel, and N. Mohandas, “Structural and

functional studies of interaction between Plasmodium falciparum knob-associated

histidine-rich protein (KAHRP) and erythrocyte spectrin,” J. Biol. Chem., vol. 280, no. 35,

pp. 31166–31171, 2005.

[109] H. Cranston, C. Boylan, G. Carroll, S. Sutera, Williamson, I. Gluzman, and D. Krogstad,

“Plasmodium falciparum maturation abolishes physiologic red cell deformability,”

Page 129: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

117

Science (80-. )., vol. 223, no. 4634, pp. 400–403, Jan. 1984.

[110] R. Ruangveerayuth, K. T. Chotivanich, P. A. Kager, K. Silamut, B. J. Angus, M. R.

Hardeman, A. M. Dondorp, N. J. White, and J. Vreeken, “Prognostic Significance of

Reduced Red Blood Cell Deformability in Severe Falciparum Malaria,” Am. J. Trop. Med.

Hyg., vol. 57, no. 5, pp. 507–511, Nov. 1997.

[111] F. K. Glenister, “Contribution of parasite proteins to altered mechanical properties of

malaria-infected red blood cells,” Blood, vol. 99, no. 3, pp. 1060–1063, Feb. 2002.

[112] M. Paulitschke and G. B. Nash, “Membrane rigidity of red blood cells parasitized by

different strains of Plasmodium falciparum,” vol. 122, no. 5, pp. 581–589, 1993.

[113] J. P. Mills, M. Diez-Silva, D. J. Quinn, M. Dao, M. J. Lang, K. S. W. Tan, C. T. Lim, G.

Milon, P. H. David, O. Mercereau-Puijalon, S. Bonnefoy, and S. Suresh, “Effect of

plasmodial RESA protein on deformability of human red blood cells harboring

Plasmodium falciparum,” Proc. Natl. Acad. Sci., vol. 104, no. 22, pp. 9213–9217, May

2007.

[114] D. P. Kwiatkowski, “How Malaria Has Affected the Human Genome and What Human

Genetics Can Teach Us about Malaria,” Am. J. Hum. Genet., vol. 77, no. 2, pp. 171–192,

2005.

[115] M.-H. Odièvre, E. Verger, A. C. Silva-Pinto, and J. Elion, “Pathophysiological insights in

sickle cell disease.,” Indian J. Med. Res., vol. 134, no. 4, pp. 532–7, Oct. 2011.

[116] G. a Barabino, M. O. Platt, and D. K. Kaul, “Sickle cell biomechanics.,” Annu. Rev.

Page 130: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

118

Biomed. Eng., vol. 12, pp. 345–67, Aug. 2010.

[117] S. K. Ballas and E. D. Smith, “Red blood cell changes during the evolution of the sickle

cell painful crisis.,” Blood, vol. 79, no. 8, pp. 2154–63, Apr. 1992.

[118] J. M. Higgins, D. T. Eddington, S. N. Bhatia, and L. Mahadevan, “Sickle cell

vasoocclusion and rescue in a microfluidic device.,” Proc. Natl. Acad. Sci. U. S. A., vol.

104, no. 51, pp. 20496–500, Dec. 2007.

[119] H. F. Bunn, “The triumph of good over evil: protection by the sickle gene against malaria.,”

Blood, vol. 121, no. 1, pp. 20–5, Jan. 2013.

[120] J. A. Kark, D. M. Posey, H. R. Schumacher, and C. J. Ruehle, “Sickle-Cell Trait as a Risk

Factor for Sudden Death in Physical Training,” N. Engl. J. Med., vol. 317, no. 13, pp.

781–787, Sep. 1987.

[121] M. J. Murray and P. Evans, “Sudden exertional death in a soldier with sickle cell trait.,”

Mil. Med., vol. 161, no. 5, pp. 303–5, May 1996.

[122] K. K. Kerle and K. D. Nishimura, “Exertional collapse and sudden death associated with

sickle cell trait.,” Mil. Med., vol. 161, no. 12, pp. 766–7, Dec. 1996.

[123] R. L. Hughes and J. Feig, “Sickle Cell Trait–Related Exertional Deaths: Observations at

Autopsy and Review of the Literature,” Mil. Med., vol. 180, no. 8, pp. e929–e932, 2015.

[124] J. M. Higgins, D. T. Eddington, S. N. Bhatia, and L. Mahadevan, “Sickle cell

vasoocclusion and rescue in a microfluidic device.,” Proc. Natl. Acad. Sci. U. S. A., vol.

104, no. 51, pp. 20496–20500, 2007.

Page 131: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

119

[125] J. A. Reisz, M. J. Wither, M. Dzieciatkowska, T. Nemkov, A. Issaian, T. Yoshida, A. J.

Dunham, R. C. Hill, K. C. Hansen, and A. D’Alessandro, “Oxidative modifications of

glyceraldehyde 3-phosphate dehydrogenase regulate metabolic reprogramming of stored

red blood cells.,” Blood, vol. 128, no. 12, pp. e32-42, Sep. 2016.

[126] J. R. Hess, “RBC storage lesions,” Blood, vol. 128, no. 12, pp. 1544–1545, 2016.

[127] L. J. Dumont and P. James, “Evaluation of proposed FDA criteria for the evaluation of

radiolabeled red cell recovery trials.”

[128] S. Design, “Scientific problems in the regulation of red blood cell products,” Transfusion,

vol. 52, no. August, pp. 1827–1835, 2012.

[129] C. G. Koch, L. Li, D. I. Sessler, P. Figueroa, G. a Hoeltge, T. Mihaljevic, and E. H.

Blackstone, “Duration of red-cell storage and complications after cardiac surgery.,” N.

Engl. J. Med., vol. 358, no. 12, pp. 1229–1239, 2008.

[130] M. Luten, B. Roerdinkholder-stoelwinder, P. Nicolaas, M. Schaap, and W. J. De Grip,

“Survival of red blood cells after transfusion: a comparison between red cells concentrates

of different storage periods,” Transfusion, vol. 48, no. July, pp. 1478–1485, 2008.

[131] T. L. Berezina, S. B. Zaets, C. Morgan, C. R. Spillert, M. Kamiyama, Z. Spolarics, E. a

Deitch, and G. W. Machiedo, “Influence of storage on red blood cell rheological

properties.,” J. Surg. Res., vol. 102, no. 1, pp. 6–12, 2002.

[132] P. E. Marik and J. William, “Effect of Stored-Blood Transfusion on Oxygen Delivery in

Patients With Sepsis,” JAMA, vol. 269, no. 23, pp. 3024–3029, 1993.

Page 132: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

120

[133] B. Red, “Efficacy of red blood cell transfusion in the critically ill: A systematic review of

the literature*,” Crit Care Med, vol. 36, no. 9, pp. 2667–2674, 2008.

[134] C. C. Division and H. Division, “Storage time of erythrocyte concentrates in critically ill

patients,” Blood Transfus., vol. 2, pp. 15–22, 2004.

[135] L. B. Hinshaw, “Sepsis/septic shock: participation of the microcirculation: an abbreviated

review.,” Crit. Care Med., vol. 24, no. 6, pp. 1072–1078, 1996.

[136] M. K. Athar, N. Puri, and D. R. Gerber, “Anemia and Blood Transfusions in Critically Ill

Patients,” J. Blood Transfus., vol. 2012, pp. 1–7, 2012.

[137] A. Tinmouth, D. Fergusson, I. C. Yee, and P. C. Hébert, “Clinical consequences of red

cell storage in the critically ill,” Transfusion, vol. 46, no. 11, pp. 2014–2027, 2006.

[138] L. Van De Watering, “Red cell storage and prognosis,” Vox Sang., vol. 100, no. 1, pp. 36–

45, 2011.

[139] D. H. Fernandes Da Cunha, A. M. Nunes Dos Santos, B. I. Kopelman, K. N. Areco, R.

Guinsburg, C. De Ara??jo Peres, A. K. Chiba, S. T. Kuwano, C. C. N. Terzian, and J. O.

Bordin, “Transfusions of CPDA-1 red blood cells stored for up to 28 days decrease donor

exposures in very low-birth-weight premature infants,” Transfus. Med., vol. 15, no. 6, pp.

467–473, 2005.

[140] K. Gustaf Edgren, Mads Kamper-Jørgensen, Sandra Eloranta, M. M. Rostgaard, Brian

Custer, Henrik Ullum, Edward L Murphy, Michael P Busch, Marie Reilly, and Henrik

Hjalgrim, and Olof Nyrén, “Duration of red blood cell storage and survival of transfused

Page 133: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

121

patients,” Transfusion, vol. 50, no. 6, pp. 1185–1195, 2011.

[141] D. Health, “The Effect of Older Blood on Mortality, Need for ICU Care, and the Length

of ICU Stay After Major Trauma,” pp. 781–785, 2003.

[142] F. Gauvin, P. C. Spinella, J. Lacroix, G. Choker, T. Ducruet, O. Karam, P. C. Hébert, J. S.

Hutchison, H. A. Hume, and M. Tucci, “Association between length of storage of

transfused red blood cells and multiple organ dysfunction syndrome in pediatric intensive

care patients,” Transfusion, vol. 50, no. 9, pp. 1902–1913, 2010.

[143] C. G. Koch, L. Li, D. I. Sessler, P. Figueroa, G. A. Hoeltge, T. Mihaljevic, and E. H.

Blackstone, “Duration of Red-Cell Storage and Complications after Cardiac Surgery,” N.

Engl. J. Med., vol. 358, no. 12, pp. 1229–1239, 2008.

[144] A. Tinmouth, D. J. Cook, J. C. Marshall, L. Clayton, M. Sc, L. Mcintyre, J. Callum, A. F.

Turgeon, M. A. Blajchman, T. S. Walsh, S. J. Stanworth, H. Campbell, D. Phil, G.

Capellier, P. Tiberghien, L. Bardiaux, L. Van De Watering, N. J. Van Der Meer, E. Sabri,

M. Sc, D. Vo, and B. Eng, “Age of Transfused Blood in Critically Ill Adults,” N. Engl. J.

Med., vol. 372, no. 15, pp. 1410–1418, 2015.

[145] C. Aubron, G. Syres, A. Nichol, M. Bailey, J. Board, G. Magrin, L. Murray, J. Presneill, J.

Sutton, S. Vallance, S. Morrison, R. Bellomo, and D. J. Cooper, “A pilot feasibility trial of

allocation of freshest available red blood cells versus standard care in critically ill patients,”

Transfusion, vol. 52, no. 6, pp. 1196–1202, 2012.

[146] D. J. Kor, R. Kashyap, R. B. Weiskopf, G. A. Wilson, C. M. Van Buskirk, J. L. Winters,

M. Malinchoc, R. D. Hubmayr, and O. Gajic, “Fresh red blood cell transfusion and short-

Page 134: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

122

term pulmonary, immunologic, and coagulation status: A randomized clinical trial,” Am. J.

Respir. Crit. Care Med., vol. 185, no. 8, pp. 842–850, 2012.

[147] N. M. Heddle, R. J. Cook, D. M. Arnold, M. A. Crowther, T. E. Warkentin, K. E. Webert,

J. Hirsh, R. L. Barty, Y. Liu, C. Lester, and J. W. Eikelboom, “The effect of blood storage

duration on in-hospital mortality: A randomized controlled pilot feasibility trial,”

Transfusion, vol. 52, no. 6, pp. 1203–1212, 2012.

[148] P. C. Hébert, I. Chin-Yee, D. Fergusson, M. Blajchman, R. Martineau, J. Clinch, and B.

Olberg, “A pilot trial evaluating the clinical effects of prolonged storage of red cells,”

Anesth. Analg., vol. 100, no. 5, pp. 1433–1438, 2005.

[149] T. S. Walsh, F. McArdle, S. A. McLellan, C. Maciver, M. Maginnis, R. J. Prescott, and D.

B. McClelland, “Does the storage time of transfused red blood cells influence regional or

global indexes of tissue oxygenation in anemic critically ill patients?,” Crit. Care Med.,

vol. 32, no. 2, pp. 364–371, 2004.

[150] D. A. Fergusson, P. Hébert, D. L. Hogan, L. LeBel, N. Rouvinez-Bouali, J. A. Smyth, K.

Sankaran, A. Tinmouth, M. A. Blajchman, L. Kovacs, C. Lachance, S. Lee, C. R. Walker,

B. Hutton, R. Ducharme, K. Balchin, T. Ramsay, J. C. Ford, A. Kakadekar, K. Ramesh,

and S. Shapiro, “Effect of fresh red blood cell transfusions on clinical outcomes in

premature, very low-birth-weight infants: The ARIPI randomized trial,” JAMA - J. Am.

Med. Assoc., vol. 308, no. 14, pp. 1443–1451, 2012.

[151] Steiner, “Addressing the Question of the Effect of RBC Storage on Clinical Outcomes:

The Red Cell Storage Duration Study (RECESS) (Section 7),” vol. 43, no. 1, pp. 107–116,

Page 135: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

123

2011.

[152] S. M. Frank, B. Abazyan, M. Ono, C. W. Hogue, D. B. Cohen, D. E. Berkowitz, P. M.

Ness, and V. M. Barodka, “Decreased erythrocyte deformability after transfusion and the

effects of erythrocyte storage duration,” Anesth. Analg., vol. 116, no. 5, pp. 975–981,

2013.

[153] S. Henkelman, M. J. Dijkstra-Tiekstra, J. De Wildt-Eggen, R. Graaff, G. Rakhorst, and W.

Van Oeveren, “Is red blood cell rheology preserved during routine blood bank storage?,”

Transfusion, vol. 50, no. 4, pp. 941–948, 2010.

[154] J. Czerwinska, S. M. Wolf, H. Mohammadi, and S. Jeney, “Red Blood Cell Aging During

Storage, Studied Using Optical Tweezers Experiment,” Cell. Mol. Bioeng., vol. 8, no. 2,

pp. 258–266, 2015.

[155] S. Huang, H. W. Hou, T. Kanias, J. T. Sertorio, H. Chen, D. Sinchar, M. T. Gladwin, and

J. Han, “Towards microfluidic-based depletion of stiff and fragile human red cells that

accumulate during blood storage,” Lab Chip, vol. 15, no. 2, pp. 448–458, 2015.

[156] K. G. Harmon, I. M. Asif, D. Klossner, and J. a. Drezner, “Incidence of Sudden Cardiac

Death in National Collegiate Athletic Association Athletes,” Circulation, vol. 123, no. 15,

pp. 1594–1600, 2011.

[157] B. J. Maron, J. J. Doerer, T. S. Haas, D. M. Tierney, and F. O. Mueller, “Sudden deaths in

young competitive athletes analysis of 1866 deaths in the united states, 1980-2006,”

Circulation, vol. 119, no. 8, pp. 1085–1092, 2009.

Page 136: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

124

[158] D. M. Hyslop, “Sudden death in young athletes.,” N. Engl. J. Med., vol. 329, no. 23, p.

1738, 1993.

[159] K. G. Harmon, J. a. Drezner, D. Klossner, and I. M. Asif, “Sickle cell trait associated with

a RR of death of 37 times in national collegiate athletic association football athletes: a

database with 2 million athlete-years as the denominator,” Br. J. Sports Med., vol. 46, no.

5, pp. 325–330, 2012.

[160] P. Connes, H. Reid, M. D. Hardy-Dessources, E. Morrison, and O. Hue, “Physiological

responses of sickle cell trait carriers during exercise,” Sport. Med., vol. 38, no. 11, pp.

931–946, 2008.

[161] G. Tsaras, A. Owusu-Ansah, F. O. Boateng, and Y. Amoateng-Adjepong, “Complications

Associated with Sickle Cell Trait: A Brief Narrative Review,” Am. J. Med., vol. 122, no. 6,

pp. 507–512, 2009.

[162] A. a Thompson, “Sickle cell trait testing and athletic participation: a solution in search of

a problem?,” ASH Educ. Progr. B., vol. 2013, no. 1, pp. 632–637, 2013.

[163] R. S. O. A. Ashley-Koch, Q. Yang, “sickle hemoglobin (Hb S) Allele and Sickle Cell

Disease: A HuGE Review,” Amerian J. Epidemiol., vol. 160, no. 9, pp. 825–841, 2004.

[164] B. L. Mitchell, “Sickle cell trait and sudden death--bringing it home.,” J. Natl. Med.

Assoc., vol. 99, no. 3, pp. 300–305, 2007.

[165] B. a. Tarini, M. A. Brooks, and D. G. Bundy, “A policy impact analysis of the mandatory

NCAA sickle cell trait screening program,” Health Serv. Res., vol. 47, no. 1 PART 2, pp.

Page 137: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

125

446–461, 2012.

[166] J. Tripette, M. D. Hardy-Dessources, M. Romana, O. Hue, M. Diaw, A. Samb, S. Diop,

and P. Connes, “Exercise-related complications in sickle cell trait,” Clin. Hemorheol.

Microcirc., vol. 55, no. 1, pp. 29–37, 2013.

[167] J. B. Osnes and L. Hermansen, “Acid-base balance after maximal exercise of short

duration.,” J. Appl. Physiol., vol. 32, no. 1, pp. 59–63, 1972.

[168] L. Hermansen and J. B. Osnes, “Blood and muscle pH after maximal exercise in man.,” J.

Appl. Physiol., vol. 32, no. 3, pp. 304–308, 1972.

[169] C. Booth, B. Inusa, and S. K. Obaro, “Infection in sickle cell disease: A review,” Int. J.

Infect. Dis., vol. 14, no. 1, pp. 2–12, 2010.

[170] P. Connes, F. Sara, M. D. Hardy-Dessources, L. Marlin, F. Etienne, L. Larifla, C. Saint-

Martin, and O. Hue, “Effects of short supramaximal exercise on hemorheology in sickle

cell trait carriers,” Eur. J. Appl. Physiol., vol. 97, no. 2, pp. 143–150, 2006.

[171] K. D. Tachev, J. K. Angarska, K. D. Danov, and P. A. Kralchevsky, “Erythrocyte

attachment to substrates : determination of membrane tension and adhesion energy,” vol.

19, pp. 61–80, 2000.

[172] T. C. Merkel, V. I. Bondar, K. Nagai, B. D. Freeman, and I. Pinnau, “Gas sorption,

diffusion, and permeation in poly (dimethylsiloxane),” J. Polym. Sci. Part B Polym. Phys.,

vol. 38, no. 3, pp. 415–434, 2000.

[173] Y. Zheng, M. Cachia, J. Ge, Z. Xu, C. Wang, and Y. Sun, “Mechanical differences of

Page 138: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

126

sickle cell trait (SCT) and normal red blood cells,” Lab Chip, 2015.

[174] R. S. Reneman and A. P. G. Hoeks, “Wall shear stress as measured in vivo: consequences

for the design of the arterial system,” Med. Biol. Eng. Comput., vol. 46, no. 5, pp. 499–

507, 2008.

[175] E. a Evans, “New membrane concept applied to the analysis of fluid shear- and

micropipette-deformed red blood cells.,” Biophys. J., vol. 13, no. 9, pp. 941–954, 1973.

[176] S. Braunmüller, L. Schmid, E. Sackmann, and T. Franke, “Hydrodynamic deformation

reveals two coupled modes/time scales of red blood cell relaxation,” Soft Matter, vol. 8,

no. 44, p. 11240, 2012.

[177] X. Liu, Z. Y. Tang, Z. Zeng, X. Chen, W. J. Yao, Z. Y. Yan, Y. Shi, H. X. Shan, D. G.

Sun, D. Q. He, and Z. Y. Wen, “The measurement of shear modulus and membrane

surface viscosity of RBC membrane with Ektacytometry: A new technique,” Math. Biosci.,

vol. 209, no. 1, pp. 190–204, 2007.

[178] J. L. Maciaszek and G. Lykotrafitis, “Sickle cell trait human erythrocytes are significantly

stiffer than normal,” J. Biomech., vol. 44, no. 4, pp. 657–661, 2011.

[179] B. Schauf, U. Lang, P. Stute, S. Schneider, K. Dietz, B. Aydeniz, and D. Wallwiener,

“Reduced red blood cell deformability, an indicator for high fetal or maternal risk, is

found in preeclampsia and IUGR.,” Hypertens. pregnancy, vol. 21, no. 2, pp. 147–60, Jan.

2002.

[180] O. K. Baskurt and H. J. Meiselman, “Blood Rheology and Hemodynamics,” Semin.

Page 139: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

127

Thromb. Hemost., vol. 29, no. 5, pp. 435–450, 2003.

[181] T. S. Hakim, “Erythrocyte deformability and segmental pulmonary vascular resistance:

osmolarity and heat treatment.,” J. Appl. Physiol., vol. 65, no. 4, pp. 1634–41, 1988.

[182] L. Vincent, L. Féasson, S. Oyono-Enguéllé, V. Banimbek, C. Denis, C. Guarneri, E.

Aufradet, G. Monchanin, C. Martin, D. Gozal, M. Dohbobga, D. Wouassi, M. Garet, P.

Thiriet, and L. Messonnier, “Remodeling of skeletal muscle microvasculature in sickle

cell trait and alpha-thalassemia.,” Am. J. Physiol. Heart Circ. Physiol., vol. 298, no. 2, pp.

H375-84, 2010.

[183] D. Kuzman, T. Žnidarčič, M. Gros, and S. Vrhovec, “Effect of pH on red blood cell

deformability,” Pflügers Arch., pp. 193–194, 2000.

[184] M. J. Simmonds, P. Connes, and S. Sabapathy, “Exercise-Induced Blood Lactate Increase

Does Not Change Red Blood Cell Deformability in Cyclists,” PLoS One, vol. 8, no. 8,

2013.

[185] N. Nemeth, I. Miko, A. Furka, F. Kiss, I. Furka, A. Koller, and M. Szilasi, “Concerning

the importance of changes in hemorheological parameters caused by acid-base and blood

gas alterations in experimental surgical models,” Clin. Hemorheol. Microcirc., vol. 51, no.

1, pp. 43–50, 2012.

[186] F. Liu, H. Mizukami, S. Sarnaik, and A. Ostafin, “Calcium-dependent human erythrocyte

cytoskeleton stability analysis through atomic force microscopy,” J. Struct. Biol., vol. 150,

no. 2, pp. 200–210, 2005.

Page 140: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

128

[187] P. S. Low, “Structure and function of the cytoplasmic domain of band 3: center of

erythrocyte membrane-peripheral protein interactions.,” Biochim. Biophys. Acta, vol. 864,

no. 2, pp. 145–167, 1986.

[188] F. Sara, P. Connes, O. Hue, M. Montout-Hedreville, M. Etienne-Julan, and M.-D. Hardy-

Dessources, “Faster lactate transport across red blood cell membrane in sickle cell trait

carriers.,” J. Appl. Physiol., vol. 100, no. 2, pp. 427–432, 2006.

[189] M. L. Wahl, J. a Owen, R. Burd, R. a Herlands, S. S. Nogami, U. Rodeck, D. Berd, D. B.

Leeper, and C. S. Owen, “Regulation of intracellular pH in human melanoma: potential

therapeutic implications.,” Mol. Cancer Ther., vol. 1, no. 8, pp. 617–628, 2002.

[190] P. Connes, F. Sara, M.-D. Hardy-Dessources, M. Etienne-Julan, and O. Hue, “Does higher

red blood cell (RBC) lactate transporter activity explain impaired RBC deformability in

sickle cell trait?,” Jpn. J. Physiol., vol. 55, no. 6, pp. 385–387, 2005.

[191] R. A. Robergs, F. Ghiasvand, and D. Parker, “Biochemistry of exercise-induced metabolic

acidosis.,” Am. J. Physiol. Regul. Integr. Comp. Physiol., vol. 287, no. 3, pp. R502-16,

Sep. 2004.

[192] J. L. MacIaszek, B. Andemariam, G. Huber, and G. Lykotrafitis, “Epinephrine modulates

BCAM/Lu and ICAM-4 expression on the sickle cell trait red blood cell membrane,”

Biophys. J., vol. 102, no. 5, pp. 1137–1143, 2012.

[193] C. Aubron, A. Nichol, D. J. Cooper, and R. Bellomo, “Age of red blood cells and

transfusion in critically ill patients.,” Ann. Intensive Care, vol. 3, no. 1, p. 2, 2013.

Page 141: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

129

[194] N. John, “A Review of Clinical Profile in Sickle Cell Traits,” Oman Med. J., vol. 25, no. 1,

pp. 3–8, 2010.

[195] P. C. Spinella, C. L. Carroll, I. Staff, R. Gross, J. M. Quay, L. Keibel, C. E. Wade, and J.

B. Holcomb, “Duration of red blood cell storage is associated with increased incidence of

deep vein thrombosis and in hospital mortality in patients with traumatic injuries,” Crit.

Care, vol. 13, no. 5, pp. 1–11, 2009.

[196] J. A. Weinberg, G. Mcgwin, R. L. Griffin, V. Q. Huynh, S. A. Cherry, M. B. Marques, D.

A. Reiff, J. D. Kerby, and L. W. Rue, “Age of Transfused Blood : An Independent

Predictor of mortality depite universal leukoreduction,” J Trauma., vol. 65, no. 2, pp.

279–284, 2007.

[197] J. A. Weinberg, G. Mcgwin, M. B. Marques, and S. A. Cherry, “Transfusions in the Less

Severely Injured : Does Age of Transfused Blood Affect Outcomes ?,” J Trauma., vol. 65,

pp. 794–798, 2008.

[198] J. A. Weinberg, G. M. Jr, M. J. Vandromme, M. B. Marques, S. M. Melton, D. A. Reiff,

and J. D. Kerby, “Duration of Red Cell Storage Influences Mortality After Trauma,” J

Trauma., vol. 69, no. 6, pp. 1427–1432, 2011.

[199] P. L. La Celle, “Alteration of deformability of the erythrocyte membrane in stored blood.,”

Transfusion, vol. 9, no. 5, pp. 238–245, 1969.

[200] R. R. Huruta, M. L. Barjas-Castro, S. T. Saad, F. F. Costa, a Fontes, L. C. Barbosa, and C.

L. Cesar, “Mechanical properties of stored red blood cells using optical tweezers.,” Blood,

vol. 92, no. 8, pp. 2975–2977, 1998.

Page 142: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

130

[201] Y. Zheng, J. Chen, T. Cui, N. Shehata, C. Wang, and Y. Sun, “Characterization of red

blood cell deformability change during blood storage.,” Lab Chip, vol. 14, no. 3, pp. 577–

83, 2014.

[202] B. S. Karon, C. M. Van Buskirk, E. A. Jaben, J. D. Hoyer, and D. D. Thomas, “Temporal

sequence of major biochemical events during Blood Bank storage of packed red blood

cells,” Blood Transfus., vol. 10, no. 4, pp. 453–461, 2012.

[203] Z. Xu, Y. Zheng, X. Wang, N. Shehata, C. Wang, S. Xie, and Y. Sun, “Stiffening of sickle

cell trait red blood cells under simulated strenuous exercise conditions,” Microsystems

Nanoeng., vol. 2, p. 16061, 2016.

[204] D. Arora, M. Behr, and M. Pasquali, “A tensor-based measure for estimating blood

damage,” Artif. Organs, vol. 28, no. 11, pp. 1002–1015, 2004.

[205] L. Gu, W. A. Smith, and G. P. Chatzimavroudis, “Mechanical fragility calibration of red

blood cells,” ASAIO J., vol. 51, no. 3, pp. 194–201, 2005.

[206] W. L. Haberman and R. M. Sayre, “Motion of rigid and fluid spheres in stationary and

moving liquids inside cylindrical tubes.,” David Taylor Model Basin Report, Washingt. D.

C., U. S. Navy Dept., 1958.

[207] R. P. Chhabra, Bubbles, Drops and Particles in Non-Newtonian Fluids. 2007.

[208] M. R. Hardeman, G. a J. Besselink, I. Ebbing, D. De Korte, C. Ince, and A. J. Verhoeven,

“Laser-assisted optical rotational cell analyzer measurements reveal early changes in

human RBC deformability induced by photodynamic treatment,” Transfusion, vol. 43, no.

Page 143: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

131

November, pp. 1533–1537, 2003.

[209] F. Li, C. U. Chan, and C. D. Ohl, “Yield Strength of Human Erythrocyte Membranes to

Impulsive Stretching,” Biophysj, vol. 105, no. 4, pp. 872–879, 2013.

[210] M. Singh and J. F. Stoltz, “Influence of temperature variation from 5 degrees C to 37

degrees C on aggregation and deformability of erythrocytes.,” Clin. Hemorheol.

Microcirc., vol. 26, no. 1, pp. 1–7, 2002.

[211] T. Lecklin, S. Egginton, and G. B. Nash, “Effect of temperature on the resistance of

individual red blood cells to flow through capillary-sized apertures.,” Pflügers Arch. Eur.

J. Physiol., vol. 432, no. 5, pp. 753–9, 1996.

[212] M. W. Rampling and P. Whittingstall, “The effect of temperature on the viscosity

characteristics of erythrocyte suspensions,” Ciln. Hermorheol., vol. 7, no. c, pp. 745–755,

1987.

[213] C. Aponte-Santamaria, R. Briones, A. D. Schenk, T. Walz, and B. L. de Groot,

“Molecular driving forces defining lipid positions around aquaporin-0,” Proc. Natl. Acad.

Sci., vol. 109, no. 25, pp. 9887–9892, 2012.

[214] A. Nans, N. Mohandas, and D. L. Stokes, “Native ultrastructure of the red cell

cytoskeleton by cryo-electron tomography,” Biophys. J., vol. 101, no. 10, pp. 2341–2350,

2011.

[215] G. Y. Huh, S. B. Glantz, S. Je, J. S. Morrow, and J. H. Kim, “Calpain proteolysis of alpha

II-spectrin in the normal adult human brain.,” Neurosci. Lett., vol. 316, no. 1, pp. 41–44,

Page 144: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

132

2001.

[216] T. Franco and P. S. Low, “Erythrocyte adducin: A structural regulator of the red blood cell

membrane,” Transfus Clin Biol, vol. 17, no. 3, pp. 87–94, 2010.

[217] E. van den Akker, T. J. Satchwell, R. C. Williamson, and A. M. Toye, “Band 3

multiprotein complexes in the red cell membrane; of mice and men,” Blood Cells, Mol.

Dis., vol. 45, no. 1, pp. 1–8, 2010.

[218] P. Lane, G. Hao, and S. S. Gross, “S-nitrosylation is emerging as a specific and

fundamental posttranslational protein modification: head-to-head comparison with O-

phosphorylation.,” Sci. STKE, vol. 2001, no. 86, p. re1, 2001.

[219] A. Czogalla and A. F. Sikorski, “Do we already know how spectrin attracts ankyrin?,”

Cell. Mol. Life Sci., vol. 67, no. 16, pp. 2679–2683, 2010.

[220] N. S. Gov and S. A. Safran, “Red Blood Cell Membrane Fluctuations and Shape

Controlled by ATP-Induced Cytoskeletal Defects,” Biophys. J., vol. 88, no. 3, pp. 1859–

1874, 2005.

[221] J. D. Reynolds, G. S. Ahearn, M. Angelo, J. Zhang, F. Cobb, and J. S. Stamler, “S-

nitrosohemoglobin deficiency: a mechanism for loss of physiological activity in banked

blood.,” Proc. Natl. Acad. Sci. U. S. A., vol. 104, no. 43, pp. 17058–17062, 2007.

[222] A. D’Alessandro, A. G. Kriebardis, S. Rinalducci, M. H. Antonelou, K. C. Hansen, I. S.

Papassideri, and L. Zolla, “An update on red blood cell storage lesions, as gleaned through

biochemistry and omics technologies,” Transfusion, vol. 55, no. 1, pp. 205–219, 2015.

Page 145: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

133

[223] D. De Korte, M. Kleine, H. G. H. Korsten, and A. J. Verhoeven, “Prolonged maintenance

of 2,3-diphosphoglycerate acid and adenosine triphosphate in red blood cells during

storage,” Transfusion, vol. 48, no. 6, pp. 1081–1089, 2008.

[224] D. A. Riccio, H. Zhu, M. W. Foster, B. Huang, C. H. L., G. M. Palmer, and T. J.

McMahon, “Renitrosylation of banked human red blood cells improves deformability and

reduced adhesivity,” Transfusion, vol. 116, no. 8, pp. 1477–1490, 2016.

[225] Y. Park, C. a. Best, T. Auth, N. S. Gov, S. a. Safran, G. Popescu, S. Suresh, and M. S.

Feld, “Metabolic remodeling of the human red blood cell membrane,” Proc. Natl. Acad.

Sci., vol. 107, no. 4, pp. 1289–1294, 2010.

[226] F. Suhr, J. Brenig, R. Müller, H. Behrens, W. Bloch, and M. Grau, “Moderate Exercise

Promotes Human RBC-NOS Activity, NO Production and Deformability through Akt

Kinase Pathway,” PLoS One, vol. 7, no. 9, pp. 1–11, 2012.

[227] M. Grau, S. Pauly, J. Ali, K. Walpurgis, M. Thevis, W. Bloch, and F. Suhr, “RBC-NOS-

Dependent S-Nitrosylation of Cytoskeletal Proteins Improves RBC Deformability,” PLoS

One, vol. 8, no. 2, pp. 1–10, 2013.

[228] M. E. Steiner, P. M. Ness, S. F. Assmann, D. J. Triulzi, S. R. Sloan, M. Delaney, S.

Granger, E. Bennett-Guerrero, M. a. Blajchman, V. Scavo, J. L. Carson, J. H. Levy, G.

Whitman, P. D’Andrea, S. Pulkrabek, T. L. Ortel, L. Bornikova, T. Raife, K. E. Puca, R.

M. Kaufman, G. a. Nuttall, P. P. Young, S. Youssef, R. Engelman, P. E. Greilich, R. Miles,

C. D. Josephson, A. Bracey, R. Cooke, J. McCullough, R. Hunsaker, L. Uhl, J. G.

McFarland, Y. Park, M. M. Cushing, C. T. Klodell, R. Karanam, P. R. Roberts, C. Dyke,

Page 146: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

134

E. a. Hod, and C. P. Stowell, “Effects of Red-Cell Storage Duration on Patients

Undergoing Cardiac Surgery,” N. Engl. J. Med., vol. 372, no. 15, pp. 1419–1429, 2015.

[229] J. C. a Cluitmans, M. R. Hardeman, S. Dinkla, R. Brock, and G. J. C. G. M. Bosman,

“Red blood cell deformability during storage: towards functional proteomics and

metabolomics in the Blood Bank.,” Blood Transfus., vol. 10 Suppl 2, pp. 12–18, 2012.

[230] S. Huang, H. W. Hou, T. Kanias, J. T. Sertorio, H. Chen, D. Sinchar, M. T. Gladwin, and

J. Han, “Towards microfluidic-based depletion of stiff and fragile human red cells that

accumulate during blood storage,” Lab Chip, vol. 15, no. 2, pp. 448–458, 2015.

[231] WHO, “Global Database on Blood Safety, Summary Report 2011,” no. june, pp. 1–9,

2011.

[232] A. H. Barbee I. Whitaker, Srijana Rajbhandary, “The 2013 AABB Blood Collection,

Utilization, and Patient Blood Management Survey Report,” Aabb, p. 88, 2015.

[233] D. Wang, J. Sun, S. B. Solomon, H. G. Klein, and C. Natanson, “Transfusion of older

stored blood and risk of death: A meta-analysis,” Transfusion, vol. 52, no. 6, pp. 1184–

1195, 2012.

[234] K. J. Wardrop, T. J. Owen, and K. M. Meyers, “Evaluation of an additive solution for

preservation of canine red blood cells,” J. Vet., vol. 8, no. 4, pp. 253–257, 1993.

[235] R. S. Franco, “Measurement of red cell lifespan and aging,” Transfus. Med. Hemotherapy,

vol. 39, no. 5, pp. 302–307, 2012.

[236] G. Zallen, P. J. Offner, E. E. Moore, J. Blackwell, D. J. Ciesla, J. Gabriel, C. Denny, and

Page 147: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

135

C. C. Silliman, “Age of transfused blood is an independent risk factor for postinjury

multiple organ failure,” Am. J. Surg., vol. 178, no. 6, pp. 570–572, 1999.

[237] P. J. Offner, “Increased Rate of Infection Associated With Transfusion of Old Blood After

Severe Injury,” Arch. Surg., vol. 137, no. 6, p. 711, 2002.

[238] S. R. Leal-noval, D. Ph, I. Jara-lópez, D. Ph, J. L. García-garmendia, and D. Ph,

“Influence of Erythrocyte Concentrate Storage Time on Postsurgical Morbidity in Cardiac

Surgery Patients,” no. 4, pp. 815–822, 2003.

[239] P. Bärtsch and J. S. R. Gibbs, “Effect of altitude on the heart and the lungs,” Circulation,

vol. 116, no. 19, pp. 2191–2202, 2007.

[240] J. D. Anderson and B. Honigman, “The effect of altitude-induced hypoxia on heart disease:

do acute, intermittent, and chronic exposures provide cardioprotection?,” High Alt. Med.

Biol., vol. 12, no. 1, pp. 45–55, 2011.

[241] K. Sun, Y. Zhang, A. D’Alessandro, T. Nemkov, A. Song, H. Wu, H. Liu, M. Adebiyi, A.

Huang, Y. E. Wen, M. V. Bogdanov, A. Vila, J. O’Brien, R. E. Kellems, W. Dowhan, A.

W. Subudhi, S. Jameson-Van Houten, C. G. Julian, A. T. Lovering, M. Safo, K. C.

Hansen, R. C. Roach, and Y. Xia, “Sphingosine-1-phosphate promotes erythrocyte

glycolysis and oxygen release for adaptation to high-altitude hypoxia,” Nat. Commun., vol.

7, no. May, pp. 1–13, 2016.

[242] C. H. Wang and a S. Popel, “Effect of red blood cell shape on oxygen transport in

capillaries.,” Math. Biosci., vol. 116, no. 1, pp. 89–110, 1993.

Page 148: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

136

[243] E. Bennett-Guerrero, T. H. Veldman, A. Doctor, M. J. Telen, T. L. Ortel, T. S. Reid, M. a

Mulherin, H. Zhu, R. D. Buck, R. M. Califf, and T. J. McMahon, “Evolution of adverse

changes in stored RBCs.,” Proc. Natl. Acad. Sci. U. S. A., vol. 104, no. 43, pp. 17063–

17068, 2007.

[244] T. Betz, M. Lenz, J.-F. Joanny, and C. Sykes, “ATP-dependent mechanics of red blood

cells.,” Proc. Natl. Acad. Sci. U. S. A., vol. 106, no. 36, pp. 15320–15325, 2009.

[245] K. D. Tachev, J. K. Angarska, K. D. Danov, and P. a. Kralchevsky, “Erythrocyte

attachment to substrates: Determination of membrane tension and adhesion energy,”

Colloids Surfaces B Biointerfaces, vol. 19, no. 1, pp. 61–80, 2000.

[246] E. a Evans, “A new material concept for the red cell membrane.,” Biophys. J., vol. 13, no.

9, pp. 926–940, 1973.

[247] K. K. Kalsi and J. González-Alonso, “Temperature-dependent release of ATP from human

erythrocytes: Mechanism for the control of local tissue perfusion,” Exp. Physiol., vol. 97,

no. 3, pp. 419–432, 2012.

[248] I. Mustafa, A. Al Marwani, K. Mamdouh Nasr, N. Abdulla Kano, and T. Hadwan, “Time

Dependent Assessment of Morphological Changes: Leukodepleted Packed Red Blood

Cells Stored in SAGM,” Biomed Res. Int., vol. 2016, 2016.

[249] H. Park, S. Lee, M. Ji, K. Kim, Y. Son, S. Jang, and Y. Park, “Measuring cell surface area

and deformability of individual human red blood cells over blood storage using

quantitative phase imaging,” Sci. Rep., vol. 6, no. September, pp. 1–10, 2016.

Page 149: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

137

[250] A. Hategan, R. Law, S. Kahn, and D. E. Discher, “Adhesively-tensed cell membranes:

lysis kinetics and atomic force microscopy probing.,” Biophys. J., vol. 85, no. 4, pp. 2746–

59, Oct. 2003.

[251] J. R. Hess, “Red cell storage,” J. Proteomics, vol. 73, no. 3, pp. 368–373, Jan. 2010.

[252] R. Agrawal, T. Smart, J. Nobre-Cardoso, C. Richards, R. Bhatnagar, A. Tufail, D. Shima,

P. H. Jones, and C. Pavesio, “Assessment of red blood cell deformability in type 2

diabetes mellitus and diabetic retinopathy by dual optical tweezers stretching technique.,”

Sci. Rep., vol. 6, p. 15873, Mar. 2016.

[253] R. L. S. Donald M. Mock1, 2, Nell I. Matthews1, Shan Zhu1, Ronald G. Strauss3, 4 and

and J. A. W. Demet Nalbant4, Gretchen A. Cress4, “Red blood cell (RBC) survival

determined in humans using RBCs labeled at multiple biotin densities,” Biophys. Chem.,

vol. 257, no. 5, pp. 2432–2437, 2005.

[254] william h Crosby, “Normal Functions of the spleen relative to red blood cells: a review,”

Blood, 1958.

[255] T. R. L. Klei, S. M. Meinderts, T. K. van den Berg, and R. van Bruggen, “From the cradle

to the grave: The role of macrophages in erythropoiesis and erythrophagocytosis,” Front.

Immunol., vol. 8, no. FEB, 2017.

[256] M. H. A. M. Fens, G. Storm, R. C. M. Pelgrim, A. Ultee, A. T. Byrne, C. A. Gaillard, W.

W. van Solinge, and R. M. Schiffelers, “Erythrophagocytosis by angiogenic endothelial

cells is enhanced by loss of erythrocyte deformability,” Exp. Hematol., vol. 38, no. 4, pp.

282–291, 2010.

Page 150: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

138

[257] N. G. Sosale, T. Rouhiparkouhi, A. M. Bradshaw, R. Dimova, R. Lipowsky, and D. E.

Discher, “Cell rigidity and shape override CD47 ’ s ’ Self ’ signaling in phagocytosis by

Hyperactivating Myosin-II,” Blood, vol. 125, no. 3, pp. 542–553, 2014.

[258] N. Psychogios, D. D. Hau, J. Peng, A. C. Guo, R. Mandal, S. Bouatra, I. Sinelnikov, R.

Krishnamurthy, R. Eisner, B. Gautam, N. Young, J. Xia, C. Knox, E. Dong, P. Huang, Z.

Hollander, T. L. Pedersen, S. R. Smith, F. Bamforth, R. Greiner, B. McManus, J. W.

Newman, T. Goodfriend, and D. S. Wishart, “The Human Serum Metabolome,” PLoS

One, vol. 6, no. 2, p. e16957, Feb. 2011.

[259] S. Xia, G. Chen, B. Wang, Y. Yin, Z. Sun, J. Zhao, P. Li, L. Zhao, and H. Zhou,

“Addition of Sodium Pyruvate to Stored Red Blood Cells Attenuates Liver Injury in a

Murine Transfusion Model,” Mediators Inflamm., vol. 2016, 2016.

[260] G. R. Zubieta-Calleja, G. Zubieta-Castillo, P.-E. Paulev, and L. Zubieta-Calleja, “Non-

invasive measurement of circulation time using pulse oximetry during breath holding in

chronic hypoxia.,” J. Physiol. Pharmacol., vol. 56 Suppl 4, no. 4, pp. 251–6, 2005.

[261] L. Shi, T.-W. Pan, and R. Glowinski, “Deformation of a single red blood cell in bounded

Poiseuille flows,” Phys. Rev. E, vol. 85, p. 16307, 2012.

[262] T. W. Pan and T. Wang, “Dynamical simulation of red blood cell rheology in

microvessels,” Int. J. Numer. Anal. Model., vol. 6, no. 3, pp. 455–473, 2009.

[263] A. M. Klein, L. Mazutis, I. Akartuna, N. Tallapragada, A. Veres, V. Li, L. Peshkin, D. A.

Weitz, and M. W. Kirschner, “Droplet barcoding for single-cell transcriptomics applied to

embryonic stem cells.,” Cell, vol. 161, no. 5, pp. 1187–1201, May 2015.

Page 151: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

139

[264] J. Clausell-Tormos, D. Lieber, J.-C. Baret, A. El-Harrak, O. J. Miller, L. Frenz, J.

Blouwolff, K. J. Humphry, S. Köster, H. Duan, C. Holtze, D. A. Weitz, A. D. Griffiths,

and C. A. Merten, “Droplet-Based Microfluidic Platforms for the Encapsulation and

Screening of Mammalian Cells and Multicellular Organisms,” Chem. Biol., vol. 15, no. 5,

pp. 427–437, May 2008.

[265] J. R. Hess, “An update on solutions for red cell storage,” Vox Sang., vol. 91, no. 1, pp. 13–

19, 2006.

[266] A. Doctor, R. Platt, M. L. Sheram, A. Eischeid, T. McMahon, T. Maxey, J. Doherty, M.

Axelrod, J. Kline, M. Gurka, A. Gow, and B. Gaston, “Hemoglobin conformation couples

erythrocyte S-nitrosothiol content to O2 gradients.,” Proc. Natl. Acad. Sci. U. S. A., vol.

102, no. 16, pp. 5709–14, Apr. 2005.

[267] S. Tuvia, a Moses, N. Gulayev, S. Levin, and R. Korenstein, “Beta-adrenergic agonists

regulate cell membrane fluctuations of human erythrocytes.,” J. Physiol., vol. 516 ( Pt 3,

pp. 781–92, 1999.

[268] S. Biagini, C. S. Dale, J. M. Real, E. S. Moreira, C. R. R. Carvalho, G. P. P. Schettino, S.

Wendel, and L. C. P. Azevedo, “Short-term effects of stored homologous red blood cell

transfusion on cardiorespiratory function and inflammation: an experimental study in a

hypovolemia model.,” Brazilian J. Med. Biol. Res. = Rev. Bras. Pesqui. medicas e Biol.,

vol. 51, no. 1, p. e6258, Nov. 2017.

[269] S. Muenster, A. Beloiartsev, B. Yu, E. Du, S. Abidi, M. Dao, G. Fabry, J. A. Graw, M.

Wepler, R. Malhotra, B. O. Fernandez, M. Feelisch, K. D. Bloch, D. B. Bloch, and W. M.

Page 152: Mechanical Characterization of Red Blood Cells Using ...€¦ · relationship between RBC mechanical properties and diseases. The work described in this thesis focuses on characterizing

140

Zapol, “Exposure of Stored Packed Erythrocytes to Nitric Oxide Prevents Transfusion-

associated Pulmonary Hypertension,” Anesthesiology, vol. 125, no. 5, pp. 952–963, Nov.

2016.

[270] D. M. Mock, N. I. Matthews, S. Zhu, R. G. Strauss, R. L. Schmidt, D. Nalbant, G. A.

Cress, and J. A. Widness, “Red blood cell (RBC) survival determined in humans using

RBCs labeled at multiple biotin densities.,” Transfusion, vol. 51, no. 5, pp. 1047–57, May

2011.

[271] M.-C. Giarratana, L. Kobari, H. Ne Lapillonne, D. Chalmers, L. Kiger, M. C. Marden, H.

Wajcman, and L. Douay, “Ex vivo generation of fully mature human red blood cells from

hematopoietic stem cells,” Nat. Biotechnol., vol. 23, 2004.

[272] K. Miharada, T. Hiroyama, K. Sudo, T. Nagasawa, and Y. Nakamura, “Efficient

enucleation of erythroblasts differentiated in vitro from hematopoietic stem and progenitor

cells,” Nat. Biotechnol., vol. 24, no. 10, pp. 1255–1256, 2006.