Скачиваний:
98
Добавлен:
08.05.2021
Размер:
9.6 Mб
Скачать

Published on 03 April 2009. Downloaded on 11/18/2019 10:55:52 AM.

View Article Online / Journal Homepage / Table of Contents for this issue

PAPER

www.rsc.org/loc | Lab on a Chip

 

 

A novel multishear microdevice for studying cell mechanics

Lien Chau, Michael Doran and Justin Cooper-White*

Received 7th January 2009, Accepted 27th March 2009

First published as an Advance Article on the web 3rd April 2009

DOI: 10.1039/b823180j

Shear stresses are known to influence the morphology, and even the fate, of many cell types, including endothelial, smooth muscle, and osteoblast cells. This paper describes a novel shear device for the study of cell mechanics. Unlike all other published shear devices, such as parallel-plate flow chambers, where a single shear stress is evaluated for a single input flow rate, the described device enables the simultaneous evaluation of 10 different shear stresses ranging over two orders of magnitude (0.7–130 dynes cm 2, 0.07–13 Pa). Human umblical vein endothelial cells (HUVECs) were exposed to the shear stress profiles provided by the device over a 20 h perfusion period, and the secretion level of von Willebrand factor (vWF) was investigated. Confirming previous studies, increasing shear resulted in increased vWF secretion. Furthermore, changes in cell morphology, including cell and nuclear size (area) and perimeter with shear, were analysed. HUVECs under shear stresses ranging from 1–3 dynes cm 2 (0.1–0.3 Pa) showed similar vWF content, cell and nuclear size and perimeter to static cultures, while cells under shear stresses above 5 dynes cm 2 (0.5 Pa) showed significantly higher vWF secretion and were at least 30% smaller in cell size. We also note that cells exposed to perfusion rates resulting in a shear stress of 0.7 dynes cm 2 (0.07 Pa) showed significantly lower levels of vWF and were 35% smaller in size than those under static conditions. Overall, the results confirm the significant utility of this device to rapidly screen cellular responses to simultaneously imposed physiologically relevant ranges of shear stress.

Introduction

The inner surface of arteries and veins is lined with a continuous layer of endothelial cells. As blood flows across this surface, shear stresses are generated and can be sensed directly by the endothelial cells.1 Shear stress is known to regulate a number of endothelial cell functions. In particular, cell shape and orientation, and cytoskeleton protein production, secretion and organisation, are significantly influenced or modulated by flow.

A number of studies have looked at shear devices to investigate the effects of shear stress on endothelial cells.2–4 In a study conducted by Davies and colleagues,2 a shear device having a geometry similar to a cone-plate viscometer was utilised. The fluid shear stress was produced by the rotation of a shallow cone with respect to a stationary base plate. The combination of viscosity, cone angle, and rotational speed in the study produced shear stresses in the range of 1–15 dynes cm 2 (0.1–1.5 Pa). Cone- and-plate apparatus have also been used to study cells at constant shear stresses up to 50 dynes cm 2 (5 Pa).5,6 Another shear device is the parallel-plate flow chamber, which have been used by numerous groups to study endothelial cells at shear stresses of 8 dynes cm 2 (0.8 Pa),7 15.2 dynes cm 2 (1.52 Pa),8 and 10 and 25 dynes cm 2 (1 and 2.5 Pa).9 However, like the cone- and-plate apparatus, only one shear stress can be presented to cells in a parallel-plate flow chamber in each experiment, which is dependent upon the flow rate and the gap between two plates. Studying the effect of an entire physiologically relevant range of

Australian Institute for Bioengineering and Nanotechnology, The University of Queensland, St. Lucia, 4072, Australia. E-mail: j.cooperwhite@uq.edu.au; Fax: +61 7 3346 3973; Tel: +61 7 3346 3858

shear stresses on cells is thus costly in terms of time and resources.

Usami and colleagues designed a device that is able to generate a linear variant shear stress profile within the same flow field at a constant flow rate by introducing a tapered parallel-plate channel to the flow chamber device concept.4 The study investigated platelet adhesion versus shear stress with shear stress ranging from 0–43 dynes cm 2 (0–4.3 Pa). However, it was difficult to pinpoint shear effects with the device as the cells spread without a clear boundary between varying shear stress values.

Very recently, in line with our work, Gutierrez and colleagues recognised the need for a microfluidic multishear device to minimise sample volume.10 They devised a microfluidic device that had an array of 8 parallel channels covering a shear stress range from 0.5 to 50 dynes cm 2 (0.05–5 Pa). All flow chambers had a depth h ¼ 24 mm and a width w ¼ 200 mm. Gaver and Kute have shown that when the cell aspect ratio (CAR, g ¼ R/H, where R is the cell size and H is the channel height, dictating the size of the gap width between the channel wall and top of the cell) is greater than 0.25, cells will cause significant flow disruption and thus the stress, force, and torque can be significantly greater than predicted based on flow in a cell-free system.11 According to this criteria, for CAR < 0.25, the device from Gutierrez and colleagues can only be utilised for very small cells (<6 mm in height, which is lower than the mean height of a human umbilical vein endothelial cell (HUVEC) monolayer (7 mm),12 and that of many other cell types).13,14

We have developed a multishear microdevice that is simpler in design, yet enables four times greater cell sizes and spans twice the shear stress range achieved with all other shear devices. Our microdevice comprises of 10 channels, all with h ¼ 100 mm and

This journal is ª The Royal Society of Chemistry 2009

Lab Chip, 2009, 9, 1897–1902 | 1897

Published on 03 April 2009. Downloaded on 11/18/2019 10:55:52 AM.

View Article Online

w ¼ 200 mm to

study cells up to 24

m

m in height (e.g. oviductal

 

 

 

 

 

 

 

 

 

 

15

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

epithelial cells).

 

In addition, our device spans a shear stress

 

 

 

 

 

 

 

 

 

 

range of 0.7–130 dynes cm 2 (0.07–13 Pa), which also covers the

 

 

 

 

 

 

 

 

 

 

physiological shear stresses (1–70 dynes cm 2, 0.1–7 Pa)16,17 in

 

 

 

 

 

 

 

 

 

 

resting conditions. It exceeds this range to accommodate for

 

 

 

 

 

 

 

 

 

 

increased wall shear stress during exercise17 and high shear

 

 

 

 

 

 

 

 

 

 

pathologic states, e.g. thrombosis (70 to >100 dynes cm 2, 7 to

 

 

 

 

 

 

 

 

 

 

>10 Pa).16

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

This paper outlines the design and development of our novel

 

 

 

 

 

 

 

 

 

 

and simple multishear microdevice that can be used as a tool to

 

 

 

 

 

 

 

 

 

 

simultaneously investigate the effect of a wide range of shear

 

 

 

 

 

 

 

 

 

 

stresses on cell behaviour. The utility of this device was validated

 

 

 

 

 

 

 

 

 

 

by studying the effect of shear stress on the von Willebrand

 

 

 

 

 

 

 

 

 

 

factor (vWF, a large multimeric glycoprotein involved in platelet

Fig. 2 The microfluidic device with ten parallel channels.

adhesion and aggregation) production of HUVECs over a range

 

 

 

 

 

 

 

 

 

 

of 0.7–130 dynes cm 2 (0.07–13 Pa).

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

(where m is the viscosity, and L is the length of the channel). As

Experimental

 

 

 

 

 

 

 

all channels were designed to possess identical width and height,

 

 

 

 

 

 

 

the relationship between flow rate and resistance between two

 

 

 

 

 

 

 

 

 

 

 

Design and fabrication

 

 

 

 

 

channels was simplified to LAQA ¼ LBQB. Based on this equality,

The device has one inlet, one outlet and ten channels (Fig. 1). The

and the condition that the sum of flow rates in the individual

channels equal the main inlet flow rate into the device, the

flow rate within a microchannel is given by Q ¼ DP/R, where Q is

channel lengths were calculated to create a two order of magni-

the flow rate, DP is the pressure drop across the channel, and R is

tude difference in flow rates between channel 1 and channel 10,

the channel resistance. With a common reservoir outlet (Fig. 1

which in turn gave two orders of magnitude difference in shear

and Fig. 2), the pressure drop of each channel was designed to be

stress, according to tw ¼ 6Qm/(wh2), where tw is the shear stress,

equal (i.e. DPA ¼ DPB, where DPA and DPB are the pressure

and m is the viscosity of the media (0.001 N s m 2). The Reynolds

drops in channel A and channel B, respectively. Similarly, QARA

number (Re ¼ nDhr/m, where v is the velocity, Dh is the hydraulic

¼ QBRB, where QA and RA are the flow rate and resistance in

diameter (Dh

 

2wh/(w + h)), and r is the density of the media

channel A, and QB and RB are the flow rate and resistance in

 

3 ¼

 

 

 

 

 

 

 

channel B). The resistance of a rectangular microchannel with an

(998.2 kg m

 

)), and the entrance length (le ¼ (0.5 + 0.065Re)Dh)

aspect ratio of 2 (i.e. w z h, where w is the width and h is the

were also calculated

to ensure

laminar

flow and established

Poiseuille flow profiles in the channels.

 

 

height) can be found by:

 

 

 

 

 

 

 

 

To validate the device design, fluid flow profiles in the device

 

 

 

 

 

 

 

 

 

 

 

R

¼

(12 mL/wh3)[1

 

(h/w)((192/p5)PN

(1/n5)tanh(npw/

were obtained using micro particle image velocimetry (mPIV).

 

 

 

 

2h))] 1

 

n¼1,3,5

 

For these measurements, the fluid was seeded with red fluores-

 

 

 

 

 

 

 

 

 

 

cent particles of mean diameter 0.49 mm. A TSI PIVCAM 13 - 8

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

CCD (1280 1024 pixel resolution) camera synchronised with

 

 

 

 

 

 

 

 

 

 

 

a dual-head Nd:YAG pulse laser was used to obtain sequential

 

 

 

 

 

 

 

 

 

 

 

images. The time delay between two images was 40 ms. A total of

 

 

 

 

 

 

 

 

 

 

 

25 instantaneous velocity vector fields were extracted from the

 

 

 

 

 

 

 

 

 

 

 

particle images and ensemble-averaged to calculate the mean

 

 

 

 

 

 

 

 

 

 

 

local velocity. A 32 32 (pixel) interrogation window with

 

 

 

 

 

 

 

 

 

 

 

50–75% overlapping was used, corresponding to a spatial reso-

 

 

 

 

 

 

 

 

 

 

 

lution of 400 mm2. Velocity fields were then obtained using

 

 

 

 

 

 

 

 

 

 

 

Matlab, and wall and fluid shear stresses were derived from plots

 

 

 

 

 

 

 

 

 

 

 

of2 velocity versus

displacement (i.e. t

¼

dv

z/dx ¼ 2vmaxx/((w/

 

 

 

 

 

 

 

 

 

 

 

 

2

2

 

 

 

 

 

 

 

 

 

 

 

 

2) ) since vz ¼ vmax[1 (x /((w/2) ))]).

 

 

 

 

 

 

 

 

 

 

 

 

 

 

The device was fabricated using conventional microfabrication

 

 

 

 

 

 

 

 

 

 

 

techniques involving SU-8 photolithography and poly-

 

 

 

 

 

 

 

 

 

 

 

dimethylsiloxane (PDMS) soft lithography. Briefly, the designs

 

 

 

 

 

 

 

 

 

 

 

were generated using AutoCAD 2007 (Autodesk) and printed on

 

 

 

 

 

 

 

 

 

 

 

transparencies with 4000 dpi resolution (Graphic Skills, Bris-

 

 

 

 

 

 

 

 

 

 

 

bane, Australia). The transparencies were used as masks in

 

 

 

 

 

 

 

 

 

 

 

contact photolithography and a 365 nm light source (OAI, San

Fig. 1 Schematic of the multishear device containing ten channels with

Jose, CA, USA) was used to yield masters composed of a positive

relief of SU-8 photoresist on a silicon wafer.

varying lengths (H ¼ height; Z1 ¼ channel width; Z2 ¼ channel spacing;

The pattern created on the silicon wafers was replicated in

L1, 2, 3, 4, 5, 6, 7, 8, 9, 10

¼

length of channels). AA0

is the cross-sectional

view of this device.

 

 

 

 

 

 

poly-dimethylsiloxane (i.e. silicone elastomer). The silicone

 

 

 

 

 

 

 

 

 

 

 

 

 

 

1898

| Lab Chip, 2009, 9, 1897–1902

 

 

 

 

This journal is ª The Royal Society of Chemistry 2009

Published on 03 April 2009. Downloaded on 11/18/2019 10:55:52 AM.

View Article Online

elastomer (Sylgard 184, Dow Corning, Midland, MI, USA) was

average of three separate experiments. Student t-test was

mixed 10 : 1 (w/w) with a silicone elastomer curing agent. The

peformed and the results are represented as mean SE.

mixture was poured over the patterned wafer to completely cover

 

the pattern and placed in the oven for 40 min at 60 C to cure.

Immunofluorescence labeling

The PDMS replica of the network design was cut out and finally

 

bonded to a clean glass slide using an air plasma to produce

After exposure to shear, the cells in the shear devices were rinsed

a sealed microfludic network. In addition, a special mould

with PBS. They were then fixed with 4% paraformaldehye in PBS

featuring an inverted cone, to allow easy removal of air bubbles

solution for 10 min and permeabilised with 1% Triton-X100 in

in the device, was made for the reservoir outlet. The PDMS

PBS for 10 min. The device was incubated with normal goat

replica of the reservoir was bonded to the rest of the device after

serum for 30 min to prevent nonspecific binding. The von Wil-

being treated with air plasma (Fig. 2).

lebrand factor of the cells was labelled with indirect immuno-

 

fluorescence staining. Firstly, polyclonal rabbit anti-human vWF

Cells

antibody (DakoCytomation, Denmark) at a 1 : 400 dilution was

used to incubate the cells for 45 min, at room temperature

 

HUVECs were harvested from human umbilical cord veins.

( 20 C). Alexa Fluor 633 (lem ¼ 515 nm) conjugated goat anti-

Briefly, the cells were isolated by perfusing the veins with a digest

rabbit IgG (Molecule Probes, USA) was used to label the cells as

solution (0.2% w/v Collagenase Type 2 (Worthington) in Hank’s

the second antibody at a dilution of 1 : 250 for 30 min, at room

balanced salts solution (Sigma)) for 10 min at 37 C. The

temperature. Cells were thereafter rinsed with PBS. Phalloidine-

harvested cells were grown in medium RPMI 1640 (Gibco) with

rhodamine (2 U mL 1, lem ¼ 565 nm) was used to label the

10% fetal bovine serum (Gibco), 100 mg mL 1 endothelial cell

F-actin microfilaments of the cytoskeleton of the HUVECs at

growth supplement (ECGS, BD BioSciences), 50 mg mL 1

a dilution of 1 : 50 for 30 min. DNA staining was carried out

heparin (Sigma), and 100 mg mL 1 penicillin/streptomycin. Only

using Hoescht (33342, Invitrogen) for 10 min.

cells from passages 4 to 8 were used in our experiments. Cell

Images were acquired with a Qimaging CCD camera (Retiga

culture was conducted in a humidified 5% CO2 and 95% air

Exi FAST 1394) connected to an Olympus BX61 inverted

incubator at 37 C.

microscope. The fluorescent images obtained from the Qimaging

 

CCD camera were analysed with Invivo and Image-Pro Plus 6.1

Shear experiment

software to obtain mean fluorescence intensity (MFI) values.

Image J software was used to obtain cell and nuclear sizes (the

 

After bonding, the devices were sterilised by autoclaving and

area covered per cell and nucleus), perimeters (the length of

allowed to cool down overnight in a laminar flow hood. The

outline around a cell and nucleus), and circularities (a measure of

channels of the device were then coated with fibronectin (human

elongation, where a circularity value of 1.0 indicates a perfect

plasma, Sigma) via injection of a solution of 100 mg mL 1

circle, and decreasing values indicate increasing elongation).

fibronectin in PBS into the device and incubated for 1 h at room

 

temperature. The channels were then washed with PBS to remove

Results and discussion

unattached fibronectin. A cell suspension of 1 106 cells mL 1

 

was injected into the fibronectin coated channels and incubated

Measured and calculated superficial velocities and shear stresses

at 37 C for 6 h.

are shown in Fig. 4. As designed, the device delivers shear stresses

Microfluidic devices have the advantage of minimising sample

spanning over at least two orders of magnitude (0.7–130 dynes

volume, and are thus especially ideal when samples are scarce

cm 2, 0.07–13 Pa). This range of shear stresses exceeds the range

and expensive. However, under the most stringent condition of

of shear stresses deemed to be physiological (1–70 dynes cm 2,

no perfusion where all nutrient transport into the channels is via

0.1–7 Pa) within in vivo vasculature. It also exceeds the range of

diffusion, small dimensions are likely to impede nutrient

shear stresses achieved by former systems, and has the advantage

delivery.18,19 This is particularly evident during the cell attach-

of confined shear stresses in individual channels for accurate data

ment period. The cell attachment period was selected after cells

acquisition.4,9,10,18

were observed at 4, 6, 8, 12, and 24 h without perfusion in the

The HUVECs were evenly distributed in all channels after 6 h

microchannels. The cells appeared attached by 4 h. They spread

of cell attachment. While cell attachment is still evident in all

and retained an appearance that was similar to those cultured in

channels of the device (Fig. 3) after 20 h of perfusion, substantial

macroscopic conditions for up to 8 h. However, at 12 h, 40% of

reductions in cell density are noted in channel 1 (130 dynes cm 2,

the cells appeared to round up. By 24 h, more than 90% of the

13 Pa). This shear stress obviously exceeds the limit of HUVEC

cells detached. From this observation, the cell attachment period

adhesion. This highlights the versatility of this multishear

of 6 h was selected.

microdevice in detachment studies. It can be used to investigate

The shear device was connected to a peristaltic pump (403U/

the ability of endothelial cells to withstand exposure to a range of

VM4 Watson Marlow Pump, Sweden). Silicone tubing (20 cm,

shear stresses, which is important to the viability of pre-seeded

1500 mm ID, 1000 mm thick) was connected to the system for

vascular grafts, as seeded cells can be washed away once blood

bubble free oxygen supply. Perfusion was conducted at an inlet

flow is restored to the graft.20

flow rate of 20 mL h 1 (steadily ramped up from 0 mL h 1 in 1

Von Willebrand factor is a large multimeric plasma protein

min) on sub-confluent layers of HUVECs. The cell seeding inlet

produced by endothelial cells and it is involved in promoting

(Fig. 1) was not used during perfusion. The cells were then fixed

platelet adhesion to damaged vessel walls. According to the

and immunostained (Fig. 3). The data presented represents the

results (Fig. 5), the vWF content of HUVECs increases with

 

 

This journal is ª The Royal Society of Chemistry 2009

Lab Chip, 2009, 9, 1897–1902 | 1899

View Article Online

Published on 03 April 2009. Downloaded on 11/18/2019 10:55:52 AM.

Fig. 3 (A) HUVECs were immunostained for intracellular and extracellular von Willebrand factor (red), rhodamine phalloidon (green), and Hoechst (blue) after 20 h of perfusion. (Scale bar: 50 mm). (B) Red channel showing vWF content of HUVECs (Scale bar: 50 mm).

increasing shear. This finding is in line with a number of

into the circulating medium and subsequently extracellular vWF

studies.21,22 In addition, the vWF content tends to be highly

deposits are observed in the direction of flow.

concentrated around the nucleus at lower shear stress, and it is

According to Fig. 5, after 20 h of perfusion, HUVECs under

noted in the periphery of the intracellular space and the extra-

shear stresses ranging from 1–3 dynes cm 2 (0.1–0.3 Pa) produce

cellular matrix at higher shear stresses. Interestingly, under high

vWF at a level similar to the control (i.e. cells in the far corner of

shear, vWF appears to be more concentrated at the part of a cell

the reservoir in the same device where they experience zero to

that is furthest away from the direction of flow. This suggests

neglible fluid flow, similar to cells in culture flasks). Significantly

that endothelial cells, grown under static conditions and that are

higher vWF content is noted for cells under shear stresses above 5

suddenly exposed to shear stress, increase their vWF secretion

dynes cm 2 (0.5 Pa), suggesting that an increased vWF

 

 

1900 | Lab Chip, 2009, 9, 1897–1902

This journal is ª The Royal Society of Chemistry 2009

Published on 03 April 2009. Downloaded on 11/18/2019 10:55:52 AM.

View Article Online

Fig. 4 Measured and calculated superficial velocities and shear stresses for the multishear device at an inlet flow rate of 20 mL h 1. Standard error bars are smaller than symbols.

Fig. 5 Shear-dependent changes of vWF content of HUVECs after 20 h of perfusion. Data are the mean SE. **P < 0.05 flow versus static control.

production is necessary for these cells in terms of cell attachment. As for cells at 0.7 dynes cm 2 (0.07 Pa), the vWF content is significantly lower than the static control. This is interesting, considering that vWF is involved in cell adhesion. One explanation is that the flow rate associated with this shear stress is below the flow rate required for effective nutrient delivery and gas exchange within the microchannels. As aforementioned, microchannels hold very small volumes of medium ( 1 mL per channel) and perfusion is required for maintaining long-term cell viability.18,19 Therefore, it is likely that cells under 0.7 dynes cm 2 (0.07 Pa) were under an imposed quiescent-like state, resulting in reduced protein production. In order to confirm this, the cell morphology was analysed and further discussion is provided later in this paper.

In terms of morphology (cell and nuclear sizes, perimeters and circularities described in the Experimental section), HUVECs under shear stresses between 1–3 dynes cm 2 (0.1–0.3 Pa) closely resemble the static control, especially cells under 1.5 dynes cm 2 (0.15 Pa) (Fig. 3 and Fig. 6). Above 5 dynes cm 2 (0.5 Pa), cells are at least 30% smaller than the static control and display the general trend of decreasing cell size with increasing shear stress, yet the cell perimeter remains similar to the static control (Fig. 6A and Fig. 6B), except for cells at 130 dynes cm 2 (13 Pa), where cytoplasmic fragmentation is observed (Fig. 3). These cells appear to be retracting and blebbing, common signs of

Fig. 6 Quantitative analysis of cell morphology for HUVECs after 20 h of perfusion as a function of shear stress. (A) Cell and nuclear size. (B) Cell and nuclear perimeter. (C) Cell and nuclear circularity (a circularity value of 1.0 indicates a perfect circle. As the value approaches 0.0 it indicates an increasingly elongated polygon). Data are the mean SE. **P < 0.05 flow versus static control.

apoptosis.23,24 In particular, cells under 130 dynes cm 2 (13 Pa) have nuclear sizes that are 22% smaller than those in the static control, and extensive cell detachment is observed.

HUVECs under 0.7 dynes cm 2 (0.07 Pa) are 35% smaller than those in the static control (Fig. 6A). This further confirms our hypothesis that the flow rate corresponding to a shear stress of 0.7 dynes cm 2 (0.07 Pa) is below the ideal perfusion flow rate for maintaining HUVECs in microchannels, as mentioned earlier. Such information is particularly useful for vascular studies investigating the growth of endothelial cells in micronetworks, such as in vitro replicas of capillary beds, where perfusion is necessary for nutrient delivery. According to the results, flow rates associated with shear stresses from 1–3 dynes cm 2 (0.1–0.3 Pa) are most effective for growing HUVECs under perfusion in microchannels and micronetworks.

This journal is ª The Royal Society of Chemistry 2009

Lab Chip, 2009, 9, 1897–1902 | 1901

Published on 03 April 2009. Downloaded on 11/18/2019 10:55:52 AM.

View Article Online

On a final note, there is more cell elongation and actin filament

3

J. E. Moore, Jr., E. Burki, A. Suciu, S. Zhao, M. Burnier,

alignment in the direction of flow at higher shear levels (Fig. 3

 

H. R. Brunner and J. J. Meister, Ann. Biomed. Eng., 1994, 22, 416.

4

S. Usami, H. H. Chen, Y. Zhao, S. Chien and R. Skalak, Ann. Biomed.

and Fig. 6C). This agrees with findings in a number of studies.22,25

 

Eng., 1993, 21, 77.

In addition, in a small 2 h perfusion study, the cells were observed

 

5

C. Urbich, E. Dernbach, A. Reissner, M. Vasa, A. M. Zeiher and

to migrate downstream in the direction of flow. Future studies

 

S. Dimmeler, Arterioscler. Thromb. Vasc. Biol., 2002, 22, 69.

will incorporate

time lapse microscopy to investigate cell

6

J. Seebach, P. Dieterich, F. Luo, H. Schillers, D. Vestweber,

 

H. Oberleithner, H. J. Galla and H. J. Schnittler, Lab Invest., 2000,

behaviour with shear over time.

 

 

 

 

 

 

 

 

 

 

80, 1819.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

7

M. Morigi, C. Zoja, M. Figliuzzi, M. Foppolo, G. Micheletti,

Conclusions

 

 

 

 

 

 

 

M. Bontempelli, M. Saronni, G. Remuzzi and A. Remuzzi, Blood,

 

 

 

 

 

 

 

 

 

1995, 85, 1696.

This multishear

microdevice delivers

physiologically

relevant

8

C. G. Galbraith, R. Skalak and S. Chien, Cell Motil. Cytoskeleton,

shear stresses, spanning over two orders of magnitude for any

 

1998, 40, 317.

9

J. Y. Ji, H. Jing and S. L. Diamond, Circ. Res., 2003, 92, 279.

one flow rate, making it a very convenient and time saving tool

10

E. Gutierrez, B. G. Petrich, S. J. Shattil, M. H. Ginsberg,

for mapping or screening cell behaviour. In addition, it requires

 

A. Groisman and A. Kasirer-Friede, Lab Chip, 2008, 8, 1486.

minimal

sample

volume,

making

it

attractive

for

studies

11

D. P. Gaver, 3rd and S. M. Kute, Biophys. J., 1998, 75, 721.

12

G. Kaplanski, C. Farnarier, A. M. Benoliel, C. Foa, S. Kaplanski and

involving

scarce

and expensive samples. To date,

this device

 

P. Bongrand, J. Cell Sci., 1994, 107(Pt 9), 2449.

covers the widest range of shear stresses in comparison to other

 

13

W. Van Driessche, P. De Smet and G. Raskin, Pfluegers Arch., 1993,

published shear devices. We used it to study HUVECs and our

 

425, 164.

results showed that there was a shear-dependent regulation of

14

E. B. Hunziker and R. K. Schenk, J. Physiol., 1989, 414, 55.

15

H. G. Verhage, M. L. Bareither, R. C. Jaffe and M. Akbar, Am.

vWF, and cell elongation and actin filament alignment in the

 

J. Anat., 1979, 156, 505.

 

 

 

 

 

 

 

 

 

direction of shear.

 

 

 

 

 

16

A. M. Malek, S. L. Alper and S. Izumo, Jama, 1999, 282, 2035.

 

 

 

 

 

 

 

 

17

A. Gnasso, C. Carallo, C. Irace, M. S. De Franceschi, P. L. Mattioli,

Acknowledgements

 

 

 

 

 

 

C. Motti and C. Cortese, Atherosclerosis, 2001, 156, 171.

 

 

 

 

 

18

H. Lu, L. Y. Koo, W. M. Wang, D. A. Lauffenburger, L. G. Griffith

 

 

 

 

 

 

 

 

The authors wish to acknowledge Dr Nicholas Timmins and

 

and K. F. Jensen, Anal. Chem., 2004, 76, 5257.

19

G. M. Walker, H. C. Zeringue and D. J. Beebe, Lab Chip, 2004, 4, 91.

Dr Ian Aird from the Australian Institute for Bioengineering and

20

P. Feugier, R. A. Black, J. A. Hunt and T. V. How, Biomaterials,

Nanotechnology (AIBN) for donating HUVECs, and funding

 

2005, 26, 1457.

from Australian

Research

Council

(ARC) Discovery

Grants

21

M. Galbusera, C. Zoja, R. Donadelli, S. Paris, M. Morigi, A. Benigni,

 

M. Figliuzzi, G. Remuzzi and A. Remuzzi, Blood, 1997, 90, 1558.

Scheme.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

22

R. J. Sun, S. Muller, X. Wang, F. Y. Zhuang and J. F. Stoltz, Clin.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Hemorheol. Microcirc., 2000, 23, 1.

References

 

 

 

 

 

 

23

G. Hacker, Cell Tissue Res., 2000, 301, 5.

 

 

 

 

 

 

 

 

24

B. Levkau, B. Herren, H. Koyama, R. Ross and E. W. Raines, J. Exp.

1 D. M. Wootton and D. N. Ku, Annu. Rev. Biomed. Eng., 1999, 1, 299.

 

Med., 1998, 187, 579.

2 P. F. Davies, C. F. Dewey, Jr., S. R. Bussolari, E. J. Gordon and

25

C. F. Dewey, Jr., S. R. Bussolari, M. A. Gimbrone, Jr. and

M. A. Gimbrone, Jr., J. Clin. Invest., 1984, 73, 1121.

 

 

 

P. F. Davies, J. Biomech. Eng., 1981, 103, 177.

1902 | Lab Chip, 2009, 9, 1897–1902

This journal is ª The Royal Society of Chemistry 2009

Соседние файлы в папке статьи