review of shear flow affecting proteins

13
Innocent B. Bekard, 1 Peter Asimakis, 1 Joseph Bertolini, 2 Dave E. Dunstan 1 1 Department of Chemical and Biomolecular Engineering, University of Melbourne, Parkville, VIC 3010, Australia 2 Research and Development Department, CSL Biotherapies, 189-209 Camp Rd, Broadmeadows, VIC 3047, Australia Received 14 February 2011; revised 20 April 2011; accepted 26 April 2011 Published online 4 May 2011 in Wiley Online Library (wileyonlinelibrary.com). DOI 10.1002/bip.21646 This article was originally published online as an accepted preprint. The ‘‘Published Online’’ date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley. com INTRODUCTION P roteins are linear polymer chains of combinations of the 20 amino acid residues. These complex macro- molecules form an essential part of an array of bio- logical processes in living systems, such as structural components of cells, biological catalysts (enzymes), and chemical messengers (hormones). The functional prop- erties of proteins depend on their unique three-dimensional structure which is determined primarily by the amino acid sequence of individual chains. The energy threshold between the native and denatured state of a protein is low, thus the physicochemical environment is a critical determinant of protein structure and function. 1 Perturbation of the native protein conformation results in loss of catalytic activity or biological function, 2,3 and in many cases aggregation and amyloid fibril formation. 4–6 Structural perturbation of protein molecules occurs as a result of conditions of temperature, pH, and ionic strength as well as through the presence of denaturants (e.g., urea, guanidine), organic solvents, and surfactants. 7 In addition, structural destabilization can result from exposure to hydro- dynamic shear forces 8–10 originating from shaking, sonica- tion, mixing, vortexing, and flow through conduits. 7,11,12 The effect of fluid shear stress on protein structure is a subject of practical interest because it is a common pheno- menon in bioprocessing. 13 Proteins have been shown to exhibit conformational dynamics in shear flow. 14–17 Protein solutions are exposed to shear stresses during bioprocessing steps as a result of centrifugation, fractionation, pumping, and ultrafiltration. 13,18,19 Similarly, the shipping and handling of biotherapeutics and protein based bioprocessing reagents such as monoclonal antibodies, hormones, cyto- kines, and enzymes can result in significant agitation. For these reasons, an understanding of the effects of shear flow on protein stability, and a means of testing and measuring this effect, would allow the appropriate design and selection of manufacturing processes, conditions, and formulations which would ensure maximum yield and stability. It has been known for some time that considerable shear stress is generated by blood flow in the circulatory system. 20 However the pathophysiological implications of this have not been explored. Thus, this review will also examine relevant papers in this area and highlight implications to the patho- genesis of some diseases. Review The Effects of Shear Flow on Protein Structure and Function Correspondence to: Dave E. Dunstan; e-mail: [email protected] Innocent B. Bekard and Peter Asimakis contributed equally to this work. ABSTRACT: Protein molecules are subjected to potentially denaturing fluid shear forces during processing and in circulation in the body. These complex molecules, involved in numerous biological functions and reactions, can be significantly impaired by molecular damage. There have been many studies on the effects of hydrodynamic shear forces on protein structure and function. These studies are reviewed and the implications to bioprocessing and pathophysiology of certain diseases are discussed. # 2011 Wiley Periodicals, Inc. Biopolymers 95: 733–745, 2011. Keywords: shear flow; protein; unfolding; aggregation; amyloid fibril; physiology; bioprocessing V V C 2011 Wiley Periodicals, Inc. Biopolymers Volume 95 / Number 11 733

description

Journal article on shear flow

Transcript of review of shear flow affecting proteins

  • ReviewThe Effects of Shear Flow on Protein Structure and Function

    Innocent B. Bekard,1 Peter Asimakis,1 Joseph Bertolini,2 Dave E. Dunstan11 Department of Chemical and Biomolecular Engineering, University of Melbourne, Parkville, VIC 3010, Australia

    2 Research and Development Department, CSL Biotherapies, 189-209 Camp Rd, Broadmeadows, VIC 3047, Australia

    Received 14 February 2011; revised 20 April 2011; accepted 26 April 2011

    Published online 4 May 2011 in Wiley Online Library (wileyonlinelibrary.com). DOI 10.1002/bip.21646

    This article was originally published online as an accepted

    preprint. The Published Online date corresponds to the

    preprint version. You can request a copy of the preprint by

    emailing the Biopolymers editorial office at biopolymers@wiley.

    com

    INTRODUCTION

    Proteins are linear polymer chains of combinations of

    the 20 amino acid residues. These complex macro-

    molecules form an essential part of an array of bio-

    logical processes in living systems, such as structural

    components of cells, biological catalysts (enzymes),

    and chemical messengers (hormones). The functional prop-

    erties of proteins depend on their unique three-dimensional

    structure which is determined primarily by the amino acid

    sequence of individual chains. The energy threshold between

    the native and denatured state of a protein is low, thus the

    physicochemical environment is a critical determinant of

    protein structure and function.1 Perturbation of the native

    protein conformation results in loss of catalytic activity or

    biological function,2,3 and in many cases aggregation and

    amyloid bril formation.46

    Structural perturbation of protein molecules occurs as a

    result of conditions of temperature, pH, and ionic strength

    as well as through the presence of denaturants (e.g., urea,

    guanidine), organic solvents, and surfactants.7 In addition,

    structural destabilization can result from exposure to hydro-

    dynamic shear forces810 originating from shaking, sonica-

    tion, mixing, vortexing, and ow through conduits.7,11,12

    The effect of uid shear stress on protein structure is a

    subject of practical interest because it is a common pheno-

    menon in bioprocessing.13 Proteins have been shown to

    exhibit conformational dynamics in shear ow.1417 Protein

    solutions are exposed to shear stresses during bioprocessing

    steps as a result of centrifugation, fractionation, pumping,

    and ultraltration.13,18,19 Similarly, the shipping and

    handling of biotherapeutics and protein based bioprocessing

    reagents such as monoclonal antibodies, hormones, cyto-

    kines, and enzymes can result in signicant agitation. For

    these reasons, an understanding of the effects of shear ow

    on protein stability, and a means of testing and measuring

    this effect, would allow the appropriate design and selection

    of manufacturing processes, conditions, and formulations

    which would ensure maximum yield and stability.

    It has been known for some time that considerable shear

    stress is generated by blood ow in the circulatory system.20

    However the pathophysiological implications of this have not

    been explored. Thus, this review will also examine relevant

    papers in this area and highlight implications to the patho-

    genesis of some diseases.

    ReviewThe Effects of Shear Flow on Protein Structure and Function

    Correspondence to: Dave E. Dunstan; e-mail: [email protected]

    Innocent B. Bekard and Peter Asimakis contributed equally to this work.

    ABSTRACT:

    Protein molecules are subjected to potentially denaturing

    uid shear forces during processing and in circulation in

    the body. These complex molecules, involved in numerous

    biological functions and reactions, can be signicantly

    impaired by molecular damage. There have been many

    studies on the effects of hydrodynamic shear forces on

    protein structure and function. These studies are

    reviewed and the implications to bioprocessing and

    pathophysiology of certain diseases are discussed. # 2011

    Wiley Periodicals, Inc. Biopolymers 95: 733745, 2011.

    Keywords: shear ow; protein; unfolding; aggregation;

    amyloid bril; physiology; bioprocessing

    VVC 2011 Wiley Periodicals, Inc.

    Biopolymers Volume 95 / Number 11 733

    p4notewifiHighlight

    p4notewifiHighlight

  • Studying the Effect of Shear Flow on ProteinsExtensional Flow vs. Simple Shear Flow. The common ow

    elds applied in shear studies are extensional ow and simple

    shear ow. A homogeneous extensional ow, also called

    elongational ow or stretching ow, is characterized by a lin-

    ear velocity gradient of the form vy _cy along the directionof ow (Figure 1).21 The strain rate, _c @xy@y , is constant. Anexample is the ow of a polymer solution through a con-

    verging channel with hyperbolically-shaped walls. This

    should not be confused with well developed ow of an

    incompressible uid through a circular pipe of innite

    length, which is characterized by inhomogeneous shear.

    Here, the ow pattern is of the Poiseulle type, showing a

    parabolic velocity prole where both the shear rate and

    uid velocity varies as a function of the distance from the

    vessel boundary (Figure 1). The strain rate is given as

    _c 2y0a where y0 is the distance between the center of thepolymer and the ow axis and a is a measure of the curvatureof the ow prole.22

    Simple shear ow on the other hand describes uid ow

    characterized by a velocity gradient perpendicular to the ow

    eld (Figure 1). It is modeled as a linear superposition of

    rotational ow with vorticity x, and elongational ow with

    strain rate _c @vx@y .16 In the rotational component of the oweld, protein molecules experience whole body rotation

    with no hydrodynamic shear strain, hence maintain their

    structural integrity. However, protein molecules are sub-

    jected to stretching events when oriented in the extensional

    ow eld. The hydrodynamic shear stress resulting from

    these events is thought to destabilize the native protein

    structure, leading to aggregate formation.17 For Newtonian

    uids, the mathematical relationship between shear stress

    (s) and velocity gradient ( _c) in a given ow eld is expressedas s g _c.23 A tangential stress of this type is called shearing,where innitesimally thin layers of uid slide over each other

    in laminar ow. Shear stress has units of force per unit area

    (N/m2), and is a measure of the actual hydrodynamic drag

    forces acting on the protein solution under ow. The velocity

    gradient, also referred to as the shear rate, has units of inverse

    time (s21).

    It is noteworthy that the magnitudes of rotational and

    elongational components of simple shear ow are equal (|x|5 | _c|). This implies that the conformational dynamics ofprotein molecules in the ow eld are dictated by the ran-

    dom occurrence of both events. In practice, the rotational

    and elongational components of ow differ in magnitude.

    FIGURE 1 Schematics of the common velocity proles in ow. (A) Homogeneous extensional

    ow typically associated with cross-slot geometries. The shear rate is constant. (B) Well developed

    ow in a conduit. The velocity prole is parabolic, typical of Poiseuille ow. In addition, the uid

    velocity is maximal at the center of the tube, approaching zero at the boundaries. The converse is

    true for the uid shear rate. (C) Simple shear ow, showing a velocity gradient perpendicular to

    the direction of ow. In all cases (AC), the arrows show both the direction of ow as well as the

    velocity prole.

    734 Bekard et al.

    Biopolymers

  • Devices for Generating Shear Flow

    In many protein studies, shear ow is generated by subjecting

    protein solutions to ow patterns characterized by uniform or

    heterogeneous velocity gradients.18 Shear devices that create a

    near uniform velocity gradient through protein solutions pro-

    vide a well controlled and quantiable shear stress. Hence,

    such devices are ideal for studying the effects of shear forces on

    protein molecules. Conversely, heterogeneous velocity gra-

    dients, for example, resulting from stirring or shaking, provide

    poorly controlled shear conditions that are difcult to quantify.

    It is worth noting that the differences in experimental devices

    applied in shear studies, and the associated data interpretation,

    could be a major contributory factor to the inconsistent exper-

    imental results reported throughout the literature.

    Several studies that show the effect of shear ow on a number

    of proteins, along with the shear devices employed, are summar-

    ized in Table I. By virtue of the differences in experimental design,

    the shear (strain history) as well as the shear stress values, which

    are not given here, varies signicantly among the different experi-

    ments. The experimental ow devices used for protein studies

    fall into two main categories: capillary and rotational devices.

    Capillary/Microuidic Flow Devices

    Capillary ow is where a uid is forced through a conduit of

    known dimensions, by applying a given pressure difference

    between the inlet and outlet of the conduit.43 Here, the uid

    velocity prole is of the Poiseuille type (non-homogeneous

    shear ow) with the maximum shear rate at the uid-vessel

    boundary, reducing gradually towards the vessel center

    (Figure 1B). Very high shear rates, up to 105 s21, can be

    achieved but over short residence times on the order of

    milliseconds.41,44 A schematic of a typical capillary ow tube

    is shown in Figure 2. It has been reported that the short resi-

    dence time spent in the ow eld could contribute to

    incomplete protein stretching (unfolding).45 However, recir-

    culation of protein solutions through conduits have been

    successfully applied to increase residence times,32 offering a

    good model system for uid shear stress in ultraltration

    units and membranes.

    Rotational Flow Devices

    Cone and plate, parallel plate and concentric cylinder visc-

    ometers are three types of rotational ow devices commonly

    Table I Studies of the Shear-Stability of Protein Systems Using a Variety of Shear Devices

    Device Geometry Protein Shear Rate/Rotational Speed Result References

    Couette Insulin 200 s21 Unfolding and aggregation 14

    Poly-L-lysine [333 s21 Unfolding 24b-lactoglobulin 25 s21, 150 s21 Fibril formation 25, 26von Willebrand factor \3.0 3 103 s21 Unfolding 27Amyloid-b 50 s21 Fibril formation 28

    Taylor-Couette Lysozyme 750 rpm,[14 s21 Unfolding 29, 30Cytochrome-c 750 rpm No change 29

    Cone and plate a-amylase 120 s21 Loss of activity 31Fibrinogen 290 s21 Degradation 32

    Catalase 91.5 s21 Loss of activity 33

    Rennet 9.15 s21 Loss of activity 33

    Glycoprotein Ib and IIb-IIIa 8.23 103 s21 Unfolding 34von Willebrand factor 6.7 3 103 s21 Aggregation 35

    Concentric cylinder Urease 683 s21 Particle formation 36

    Catalase 683 s21 Particle formation 36

    Alcohol dehydrogenase 683 s21 Particle formation 37

    Deoxyribonuclease 1.5 3 103 rpm No change 38Growth hormone 1.5 3 103 rpm Unfolding and fragmentation 38Urease 48 s21 Loss of activity 39

    Capillary Catalase 67 s21 Loss of activity 40

    Catalase 4.6 3 104 s21 Minor loss of activity 36Cytochrome-c 2.0 3 105 s21 No change 41

    Microuidic cell von Willebrand factor [13 103 s21 Unfolding 42

    FIGURE 2 Schematic of a simple extrusion glass capillary vis-

    cometer. The arrows show the direction of ow.

    Effects of Shear Flow on Protein Structure and Function 735

    Biopolymers

  • used to impart shear stress on protein systems across a nar-

    row gap (Figure 3). All models consist of two parts where

    one rotates with the other being stationary. In the case of the

    concentric cylinder, also known as Couette or coaxial cylin-

    der ow-cell, there are several sorts that differ in basic design

    but operate on the same principles.46 However, a through-

    cell Couette arrangement is ideal for generating a simple

    shear ow, as the design minimizes any end effects.47 This

    type of device also curtails particle settling or migration away

    from the solid boundaries, thereby limiting variations of

    uid properties across the gap.48 In addition, a nominal

    shear rate can be calculated for each rotational speed. At high

    rotational speeds, large deviations from the ideal ow may

    occur due to the formation of vortices. It is recognized that

    vortices occur in extensional ow regions in uids, and

    should therefore be avoided if a simple shear ow is

    desired.49 The main advantage of rotational shear devices

    over that of the capillary design is that the entire test solution

    is sheared at a near constant rate. This permits the study of

    any time-dependent behavior of protein samples over an

    extended period of time. In the through-cell Couette

    arrangement, the shear rate can be regarded as constant only

    if the gap width is much less than the radius of the inner cyl-

    inder.50 In the cone-and-plate design, the solution is con-

    tained in the space between a cone of large apex angle and a

    plate with a at surface normal to the axis. For small angles

    between the cone and plate, the shear rate is considered con-

    stant for all radial positions.

    The Effects of Shear Flow on Protein SolutionsProtein Function. Preliminary studies on the relationship

    between shear ow and protein structure exploited the cata-

    lytic activity of enzyme solutions as a means of measuring

    protein integrity. Thus, a decrease in enzyme activity implied

    a structural deformation of the native enzyme molecules.

    These early studies were critical in dening the appropriate,

    and operation of, devices to allow the study of the behavior

    of proteins in shear ow, allowing the estimation of shear

    forces required to perturb protein function with insight into

    possible resulting structural changes.

    Studies on catalase, rennet, and carboxypeptidase solu-

    tions sheared in a narrow gap coaxial viscometer40 at shear

    rates between 9.15 and 1155 s21 for 90 min, showed a strong

    correlation between decrease in enzyme activity and shear

    (shear rate 3 time). A loss of enzyme activity occurred for

    shear greater than 104 although the rennet solutions partially

    regained activity upon the cessation of shearing. The

    observed enzyme inactivation was attributed to the breaking

    of the molecules tertiary structure when oriented appropri-

    ately in the shear eld. The authors discounted any signi-

    cant contribution from the airliquid interface in the

    viscometer, where protein molecules, by virtue of their

    amphipathic nature, may aggregate (causing inactivity). It is

    also possible that the loss in enzyme activity was due to a

    ow-induced limited residence time for enzyme-substrate

    interaction. A similar shear-effect was observed upon expo-

    sure of catalase and rennet solutions to shear rates in the

    region of 104 s21 in model ultraltration systems.33 Whereas

    the catalase solutions lost some activity during ultraltration,

    the rennet solutions did not suffer inactivation in the circula-

    tion lter systems. Further studies on plasma brinogen,

    sheared in a Weissenberg rheogoniometer (shear rates 290

    1155 s21) with a Couette attachment, revealed a tempera-

    ture-independent loss in its clottability.32

    Studies on the shear-stability of urease, using a coaxial cyl-

    inder viscometer, which allowed direct monitoring of urease

    catalyzed hydrolysis of urea during shear exposure, showed

    continuous decrease in the rate of urease activity with

    increasing shear rate (48, 288, 741, and 1717 s21).39 The par-

    tial deactivation of urease revealed both reversible and irre-

    versible elements. The reversible inactivation was ascribed to

    hydrodynamic distortion of the enzymes structure in ow,

    whereas permanent inactivation was attributed to the break-

    ing of the enzymes tertiary structure. The permanent loss in

    urease activity correlated with shear strain, with a critical

    shear value of 105, similar to that previously reported.40 A

    comparison of catalase degradation in streamline and turbu-

    lent ow regimes, upon mixing in a cylindrical container,

    showed a signicant inactivation in the latter where a vortex

    and air bubbles had formed in the solution.51 This suggested

    that the airliquid interface may have contributed signi-

    cantly to catalase degradation. It was also shown that shear

    stress in selected microltration membranes caused a signi-

    cant decrease in the catalytic activity of yeast alcohol dehy-

    drogenase (ADH), under solution conditions where the

    FIGURE 3 Schematics of the common rotational ow devices:

    (A) concentric cylinder, (B) cone and plate, and (C) parallel plate

    rheometers.

    736 Bekard et al.

    Biopolymers

  • enzyme was slightly unstable.52 This observation was attrib-

    uted to membrane-enzyme interactions resulting from a

    shear-induced deformation of the enzyme structure.

    In studies of the shear-stability of ADH solutions using a

    high shear concentric viscometer, a rotating disk reactor and

    several pumps, it was found that fully active tetrameric mole-

    cules of ADH showed negligible change in activity even for

    strain rates as high as 26, 000 s21 over 1 h in the high shear

    device.53 Turbulent regimes in the rotating-disk reactor

    (3600 rpm for 5 h) and up to about 1500 passes in the

    various pumps likewise had little effect on ADH activity.

    However, the introduction of a gas-liquid interface in the

    rotating-disk reactor resulted in a 40% decrease in the initial

    ADH activity after 5 h. Similarly, a signicant decrease in

    ADH activity was observed in both a gear and a Jabsco pump

    after 2000 passes.53 This report suggested that interfacial

    denaturation, and not shear per se, accounts for the defor-

    mation of proteins in solutions exposed to shear ow, and

    that the native conformation of globular proteins, when in a

    chemically stable environment, are stable in shear ows usu-

    ally encountered in practice.

    The importance of an airliquid interface in promoting

    shear-induced protein denaturation was shown in a study

    with progesterone 11a-hydroxylase complex, using anopen concentric cylinder viscometer and a horizontal vis-

    cometer closed at both ends.54 It was observed that the

    gas-liquid interface, in concert with high shear regimes,

    which replenished the protein at the interface, promoted

    rapid protein denaturation as previously reported.53 In

    addition, a wall surface effect on the enzyme complex was

    observed as the gap width of the viscometer was reduced

    from 0.475 to 0.15 mm. Furthermore, it was observed

    that viscous enzyme solutions showed a marked decrease

    in activity as a function of shear stress.54 These observa-

    tions lead the authors to propose that the applied shear

    stress resulted in enzyme denaturation via either the rup-

    ture of the enzyme membrane or the enhanced solubiliza-

    tion of component(s) of the enzyme complex by the ve-

    locity gradient in the agitated solution.

    The above studies clearly showed that shear could have a

    profound effect on the activity of enzymes, and by inference

    on protein structure. The shear-effect was more pronounced

    in the presence of an airliquid interface. A key limitation of

    the above studies was that any deduction of structural change

    was based on indirect evidence derived from enzyme activity.

    Indeed, whereas shear may not have caused signicant struc-

    tural damage, a decrease in enzyme activity could result from

    minor modications to an enzymes active site. In many of

    the experiments, the enzymes were subjected to intermittent

    shearing force to allow sampling. Therefore, signicant time

    would have elapsed between sampling of solutions and test-

    ing for enzyme activity. This would have allowed reversible

    structural changes to revert and thus go undetected. The

    advent of new experimental techniques, allowing real-time

    measurements of the solution conformation of proteins, is

    providing new insight into the shear-stability of proteins

    with respect to variations in the sensitivity of different pro-

    teins to shear, and the specic molecular changes occurring

    in response to shear exposure.14,55

    Protein Structure

    Human von Willebrand Factor (vWF), a blood plasma pro-

    tein with important functions in coagulation, was amongst

    the rst of protein systems, besides enzymes, to be studied

    for shear induced structural changes. vWF molecules, which

    are initially secreted by the endothelial cells as large, hyper-

    active multimers, with relative molecular mass up to 50 3106, circulate as a series of smaller, less active proteolytic frag-

    ments in the plasma as a result of the action of the metallo-

    protease ADAMTS-13.56,57 Interestingly, the proteolysis of

    large vWF multimers is not observed when normal plasma is

    incubated in vitro.58 This lead to the conclusion that the gen-

    eration of proteolytic fragments of vWF in circulation

    depends on haemodynamic shear stress.57 The size distribu-

    tion of vWF multimers in microcirculation is critical to its

    haemostatic potential.

    Tsai et al. investigated this theory by perfusing normal

    plasma through long capillary tubes (190 or 354 cm in

    length), at shear rates between 1508 and 4761 s21, followed

    by an examination of the size distribution of vWF using a

    combination of SDS agarose gel electrophoresis and densito-

    metric scanning analysis of autoradiographs.59 Relative to

    unsheared plasma, an increase in smaller vWF multimers was

    observed in the sheared samples at the expense of the larger

    multimers, especially in high shear regimes. These observa-

    tions lead to the conclusion that shear stress might be

    involved in the modulation of the size distribution of vWF in

    circulation by enhancing the proteolytic cleavage of large

    vWF multimers.59

    In a subsequent study by Siedlecki et al., using a combina-

    tion of atomic force microscopy and a rotating disk system,

    vWF molecules were observed (visually) to undergo a shear-

    induced conformational transition from a globular state to

    an extended chain conformation at a critical shear stress

    value[31.5 dyn/cm2 ( _c 4618.8 s21).60 The result was theexposure of intra-molecular domains of vWF. It was sug-

    gested that shear-induced structural changes in vWF may

    have an important role in platelet adhesion and thrombus

    formation in regions of high shear stress, as would develop at

    the site of a bleeding injury.60

    Effects of Shear Flow on Protein Structure and Function 737

    Biopolymers

    p4notewifiImage

    p4notewifiHighlight

    p4notewifiHighlight

    p4notewifiLine

  • To this point, it had been considered that clot formation

    was initiated through direct shear activation of platelets.58

    This was substantiated in a later study,61 which investigated

    the mechanics involved in the formation of reversible tether

    bonds between the platelet glycoprotein receptor Iba and theA1 domain of surface-immobilized vWF under hydro-

    dynamic shear ow. Puried platelets were perfused over

    plasma vWF-A1 coated surfaces in a parallel-plate ow

    chamber at shear stresses between 0.2 and 8 dyn/cm2. Platelet

    tethering was visualized using a Nikon microscope equipped

    with a high speed video camera, and the motion of the teth-

    ered platelets calculated using an analytical two-dimensional

    model. The results showed that hydrodynamic compressive

    forces inuence tether bond formation between platelets and

    vWF.61

    Recently, uorescence microscopy has been used to

    directly visualize individual uorescently labeled vWF multi-

    mers under hydrodynamic stress in a microuidic device.42

    vWF was uorescently labeled with Alexa uor 488, which

    was attached to the primary amines of the glycoprotein using

    tetrauorophenyl ester. It was found that shear rates greater

    than 1000 s21 triggered a reversible stretching of vWF in so-

    lution. Furthermore, real-time measurements of the confor-

    mational dynamics of vWF in response to laminar shear ow

    in a Couette ow-cell, using a combination of small angle

    neutron scattering (SANS) spectroscopy and quantitative

    modeling, showed an irreversible conformational change in

    vWF, under hydrodynamic shear stress, at shear rates 3000s21.27 The observed conformational change was found to

    involve the exposure of hydrophobic domains at shear rates

    [2300 s21, as indicated by an increase in the binding afnityof the hydrophobic dye 1,10-bis(anilino)-4-,40-bis(naphta-

    lene)-8,80-disulfonate (bis-ANS).27 The shear-effect was

    more pronounced with increasing shear rate.

    A more recent single-molecule investigation of the shear-

    stability of vWF has provided more insight into the relevance

    of shear stress in the homeostatic regulation of the size of cir-

    culating vWF multimers as well as the potentiation of vWF-

    mediated platelet aggregation.62 Specically, the study exam-

    ined the folded and unfolded states of a single A2 domain of

    vWF, which contains the site of cleavage for ADAMST-13, in

    the presence and absence of the metalloprotease. The A2

    domain of vWF was unfolded by directly applying force with

    laser tweezers. The results showed that only the unfolded A2

    domain is cleaved by ADAMST-13.62 These observations cor-

    roborated previous studies which found that shear-induced

    stretching of large vWF multimers results in the exposure of

    the scissile Tyr1605-Met1606 bond within the A2 domain, faci-

    litating cleavage by ADAMST-13 in normal plasma.56,57

    Smaller vWF multimers in microcirculation equally undergo

    stretching in regions of high shear, such as a bleeding arte-

    riole, which expose bonding sites in the A1 domain of vWF

    for complexation with platelets during haemostasis.57,60,63

    While studies of vWF provided clear evidence that shear

    ow can alter protein structure, it may be argued that differ-

    ent proteins experience different shear since the strain rates

    required to trigger structural alterations decreases with

    increasing molecular weight and solvent viscosity. For exam-

    ple, studies on the effect of high shear rates on ferric equine

    cytochrome c, a small globular protein of molecular weight

    12,384 Da, showed no signicant structural alterations when

    exposed to shear rates as high as 105 s21 in a capillary ow

    device.41 By applying a simple model, the authors suggested

    that a shear rate of 107 s21 would be required to denaturea small globular protein in water.41 However, the transient

    residence time associated with capillary ow devices may

    explain the hysteresis observed in such systems.45 More

    importantly, the results41 contradict the observed abrupt

    stretching of vWF ([20,000 kDa) in a microuidic device ata threshold shear rate of 103 s21.42 Clearly, the larger, more

    complex vWF would be expected to be more susceptible to

    shear stress relative to cytochrome c. However, it is notew-

    orthy that specic conformational characteristics such as

    a-helix and b-sheet composition, and intra-molecularinteractions (e.g., hydrophobic, electrostatic) and covalent

    bonds (e.g., disulde), would be an important determinant

    of protein stability.

    Thus, protein systems with comparable molecular weights

    may differ in their shear-stability due to differences in their

    solution properties. For example, previous studies on

    the effects of both high shear ([107) and high shear rate([105 s21), using concentric cylinder-based systems, onrecombinant human growth hormone (rhGH) (22 kDa)and recombinant human deoxyribonuclease (rhDNase)

    (31 kDa) showed that rhDNase was more resistant toshear-effects.38 For instance, whereas scanning microcalorim-

    etry and SDS-PAGE analysis of extensively sheared solutions

    showed changes in the melting temperature of rhGH and the

    presence of low molecular weight fragments, respectively, no

    such changes were observed for rhDNase. Interestingly, the

    observed structural changes in rhGH were not detected in

    both the near and the far-UV circular dichroism spectral pro-

    le of the sample.38 Further studies on both samples showed

    that rhGH denatured upon exposure to high shear in the

    presence of an airliquid interface whereas rhDNase

    remained relatively stable.64 The stability of rhDNase was

    attributed to its comparatively high surface tension and low

    foaming tendency in solution.

    Intriguingly, it has been shown that human and bovine

    albumin samples, which differ in only two amino acids,

    738 Bekard et al.

    Biopolymers

  • show different aggregation kinetics upon shaking at high

    shear rates ([8000 rpm).65 Multi-angle laser light-scatter-ing measurements showed that in addition to the mono-

    mer-dimer transition observed in both albumin samples

    upon shearing, human serum albumin also formed trimers

    especially at high shear rates. Since the protein concentra-

    tion and solventair interfaces were considered as xed

    factors, the observed differences in the degree of albumin

    aggregation was attributed to the minor variation in their

    primary structure.

    Protein Aggregation

    As has been discussed above, shear can result in protein

    denaturation and compromised function. It is also well

    established that denaturation of proteins can lead to aggrega-

    tion.66 Hence, the use of agitation (uncontrolled shear stress)

    to both induce and accelerate protein aggregation and amy-

    loid bril formation is common practice throughout the lit-

    erature.66 For example, under favorable solution conditions,

    various forms of agitation including shaking, sonication and

    stirring have been successfully applied to accelerate the aggre-

    gation of insulin,6771 amyloid-b,28,72 b2-microglobulin,73,74

    b-lactoglobulin,25 albumin,65 and the PDZ domain frommurine Protein Tyrosine Phosphatase Bas-like (PDZ2)75

    among others. In one particular study, a range of structurally

    diverse protein systems including BSA, myoglobin, lysozyme,

    Tm0979, SOD and hisactophin were shown to generate amy-

    loid like structures following sonication.76

    Shear effects on protein aggregation have also been stud-

    ied in uniform ow elds. For example, a shear-induced

    nucleation, without ow enhanced aggregation, of preheated

    b-lactoglobulin solutions (0.5 wt%) at low pH has beenreported.77 b-lactoglobulin solutions were exposed to Cou-ette ow in a rheometer, at a moderate shear rate of 200 s21

    over 5 h, while measuring the ow-induced birefringence.

    Thermal treatment of the b-lactoglobulin samples triggeredthe formation of metastable pre-brillar aggregates, without

    which bril formation was absent in the sheared samples.

    Shearing of b-lactoglobulin samples was done either continu-ously or in short pulses, both producing a similar degree of

    bril formation. These observations lead to the conclusion

    that brief mechanical perturbation of an aggregation prone

    protein solution was enough to enhance bril formation.77

    That is, shear ow only triggered the nucleation of bril for-

    mation but did not inuence the polymerization of preaggre-

    gates into mature brils, discounting orthokinetic coagula-

    tion under the experimental conditions. However, the brils

    formed from the continuous shear treatment showed a

    smaller variance in Gaussian length distribution relative to

    those from the pulse shear treatments, indicating a homoge-

    nizing effect of the continuous shear treatment. Evidently, a

    synergy of heat and agitation triggered rapid bril formation.

    Another study showed that the aggregation of preheated

    b-lactoglobulin solutions in Couette ow was shear rate de-pendent.25 In addition, if shearing was performed prior to

    thermal treatment, a rapid enhancement of brillar species

    still occurred, suggesting a shear-induced formation of the

    precursor nuclei required for bril development as previously

    reported.77 Interestingly, at shear rates[500 s21, bril degra-dation, possibly arising from the persistent stretching (exten-

    sional ow) events in the ow eld, was observed. A further

    study78 revealed that a sufciently high b-lactoglobulin con-centration ([3 wt%) was required to initiate bril formationwhen samples were heated and sheared simultaneously. In

    addition, the amount of brils formed increased as a function

    of shear rate up to 337 s21 beyond which an opposing shear

    effect, similar to that reported earlier,25 was observed. The

    paradoxical shear-effect was explained by a ow enhanced

    polymerization of the bril nucleus in low shear regimes,

    whereas in high shear regimes, the extensional component of

    the ow eld overwhelmed the nascent inter-b-strand hydro-gen bonds stabilizing the aggregating brils. Furthermore, the

    authors observed that the shear applied produced shorter

    brils and also enhanced, below a critical bril concentration,

    the viscosity of the resulting bril solutions.78

    More recently, a similar ow enhanced bril formation

    has been observed in aqueous solutions of bovine insulin,14

    thermally treated amyloid-b28 and b-lactoglobulin.26 Theshear-stability of bovine insulin in Couette ow was probed

    directly via novel circular dichroism and tyrosine autouor-

    escence measurements, in situ and in real time. For a given

    shear rate, helical segments of native insulin were observed to

    unfold as a function of time. Atomic force microscopy

    images of the sheared samples revealed the presence of bril-

    lar forms in relatively strong shear regimes (600 s21).14 The

    shear-effect on preheated amyloid-b samples28 exposed toCouette ow showed similar trends to that reported earlier.25

    For the b-lactoglobulin studies,26 the authors compared themorphology and mechanical properties of brils formed

    under heterogeneous (stirred) and uniform (Couette) ow

    conditions using atomic force microscopy. They found that

    b-lactoglobulin brils resulting from heterogeneous ow hadtwisted ribbon-like morphologies and higher mechanical

    strength relative to the beaded brils formed in Couette ow.

    Evidently, the intertwining of brils produced in the variable

    shear regime imparted additional mechanical stability to the

    overall bril structure. It was hypothesized that polymeriza-

    tion of bril seeds occurs in the ow eld, and in the case of

    stirred samples, mature brils anneal into twisted ribbons

    possibly via orthokinetic coagulation. Taking together, even

    Effects of Shear Flow on Protein Structure and Function 739

    Biopolymers

  • if in some cases the initial structural destabilization required

    for protein aggregation is not triggered by shear, its ability to

    accelerate the polymerization of metastable prebrillar aggre-

    gates into mature brils still has grave implications in bio-

    processing and physiology.

    Regarding the mechanism of shear-induced protein

    aggregation, there is a general consensus that mechanical

    perturbation of a protein molecule often results in a struc-

    tural destabilization of the native conformation, leading to

    the exposure of sequestered hydrophobic residues to the

    surrounding medium. Solvent-exposed hydrophobic groups

    nucleate via hydrophobic interactions and subsequently ag-

    gregate. Several studies show that the initial destabilization

    of the protein structure occurs at the solventair or sol-

    ventsolid interface, where surface tension forces unfold

    the native protein.36,37,64,79 The airwater interfacial force

    is estimated to be 140 pN,80 taking the airwater interface

    to be 2 nm in depth with a 0.07 N/m surface tension.81

    This value is comparable to the 150 pN observed to

    unfold globular proteins in atomic force microscopy stud-

    ies.82 In addition, as agitation increases the turnover of

    the airliquid interface, and thus the number of protein

    molecules interacting with this interface, the nucleation of

    solvent-exposed hydrophobic species via hydrophobic

    interactions, and hence the aggregation of the bulk protein

    solution, increases dramatically.72 In fact, the inuence of

    agitation can be so profound that, in the case of insulin,

    it was found to attenuate the effects of other parameters,

    such as protein concentration and ionic strength, that

    inuence the kinetics of amyloid bril assembly.68

    Molecular Models and Theoretical Aspects

    The complexity of protein molecules with respect to confor-

    mational heterogeneity as well as intra- and intermolecular

    interactions between amino acid residues, with the consequent

    differences in the exibility of secondary conformations and

    tertiary congurations between proteins, results in variations

    in their solution properties in shear ow. Indeed, differences in

    the levels of association in diverse protein systems dictate the

    different degrees of shielding, especially of hydrophobic resi-

    dues concealed in the protein matrix, from hydrodynamic drag

    during shear ow.83,84 Hence, it is difcult to obtain uniform

    data to allow the formulation of a theoretical model relating

    protein conformation to shear. For these reasons, dilute solu-

    tions of homopolymers, especially unbranched polymer

    chains, have been used as model systems for shear studies

    because of their inherent structural uniformity.

    By virtue of its conformational plasticity in response to

    temperature, pH, salt concentration and alcohol content in

    solution,24 poly-L-lysine is commonly used as a model system

    in protein studies.85 The homopolypeptide exists as a

    random coil at neutral pH, a-helix at high alkaline pH andb-sheet at temperatures[308C.86 For these reasons, poly-L-lysine is the ideal choice in probing the initial conformational

    transitions, under denaturing conditions, between the three

    basic protein secondary conformations namely: random coil,

    a-helix, and b-sheet.Experimental studies on the shear-stability of poly-L-ly-

    sine solutions (437 kDa), in an initial a-helical conformation,showed an increase in solution viscosity and turbidity with

    shearing time.87 A combination of Raman spectroscopy (col-

    lected in situ) and circular dichroism were employed for data

    acquisition in a Couette ow cell at 150 s21. In addition, a

    ow induced-gelation and an increase in the b-sheet contentof the samples was observed in solution concentrations[0.3g/dL. However, the circular dichroism data further revealed a

    reversal of the a-helix to b-sheet conformational transitionafter prolonged shearing ([50 min). The observed confor-mational change and subsequent aggregation of the sheared

    samples was thought to involve ow-enhanced hydrophobic

    interactions between individual poly-L-lysine molecules in

    solution.87 Further studies, via real-time circular birefrin-

    gence measurements, showed a reversible, shear-induced, he-

    lix-to-coil transition of poly-L-lysine in simple shear ow.24

    The conformational transition was observed at a critical

    strain rate of 300 s21, and the change was attributed to a

    shear-induced breakage of intramolecular hydrogen bonds in

    a-helical poly-L-lysine.More recently, the molecular-weight-dependence of the

    shear-induced unfolding of a-helical poly-L-lysine in Couetteow has been measured in real-time using circular dichroism

    spectroscopy.55 The authors observed that the shear-stability of

    the helices increased as a function of molecular weight. The hys-

    teresis observed in the heavy chains was attributed to the large

    network of hydrogen bonds (cohesive force) stabilizing the helix

    structure, and the associated hydrodynamic screening of helical

    segments from the drag in the ow eld. Interestingly, irrespec-

    tive of the molecular weight of poly-L-lysine, unfolding of the

    helical segments was found to occur at a critical strain ( _ctc)value of 105 (Figure 4), similar to that observed in globularproteins.39,40 For this reason, the authors proposed that the

    shear rate is not as critical as the duration of its application,

    which makes the idea of a critical shear rate arbitrary. In

    addition, the helix content, a, was found to show a powerlaw dependence with strain: a _ct1=2.

    The inherent monodispersity associated with lamda bacte-

    riophage DNA made it another ideal model for studies of

    polymer dynamics in hydrodynamic ows. For example, a

    novel real-time uorescence videomicrocopy of uorescently

    labeled single DNA molecules was developed to visualize the

    740 Bekard et al.

    Biopolymers

  • conformational dynamics of DNA chains in both elongation21

    and shear16 ows. The elongation and shear ow experiments

    were performed in a microfabricated ow cell and between

    two parallel glass plates (50 lm gap), respectively, at verylow shear rates ranging from 0.05 to 4.0 s21. It was found that,

    at relatively high shear rates, the sharp coil-stretch transitions

    of DNA chains observed in pure elongation ow was absent in

    shear ow.16,21 Individual chains in steady shear ow rather

    showed large, aperiodic temporal uctuations in conforma-

    tion, consistent with what had been predicted in theory.84 In

    addition, stretching of DNA chains in shear ow occurred at a

    strain value\50 (101),16 which is several orders of magni-tude lower than the 105 critical strain value observed forpoly-L-lysine55 as well as folded proteins.39,40

    Jaspe and Hagen41 suggests that coil-stretch transitions

    readily occur in homopolymer systems because the free

    energy cost of unfolding is negligible compared to that of a

    native protein, and this was purported to explain why an ex-

    traordinarily high shear rate (107 s21) is required to destabi-

    lize the latter. Another study88 estimated a destabilizing

    strain rate in the region of 108 s21 for collapsed polymers in

    planar shear ow.

    Classical theories on the shear-stability of polymer chains

    predict that a polymer chain would remain in a coil state in

    shear ow until a critical velocity gradient, which triggers an

    abrupt unwinding of the coil, is reached.86 In addition,

    abrupt conformational transitions are not expected in simple

    shear ow since the occurrence of extensional strain, hence

    polymer stretching, is random. It is hypothesized that at very

    high strain rates, where polymer chains spend a relative short

    time (residence time) in an extensional ow eld, no unfold-

    ing or incomplete chain extensions may occur.45 These pre-

    dictions were substantiated in later studies that simulated

    polymer dynamics in shear ow. For example, in molecular

    dynamics (MD) simulations of a grafted collapsed macro-

    molecule in strong uniform ow, it was observed that a exi-

    ble macromolecule remains compact below a critical shear

    value but shows a globule-stretch transition above this criti-

    cal shear value.89 The transition was reported to be more

    pronounced in strong extensional ow.

    Lemak et al.90 later compared the conformational instabil-

    ities of a minimalist b-barrel protein in both uniform andelongational ow elds using MD simulations. It was

    reported that whereas unfolding was abrupt in elongational

    ow, the uniform ow elds presented a multi-step unfold-

    ing process, with several intermediates, depending on the

    tethered terminus of the protein. In explaining this observa-

    tion, it was proposed that whereas every monomer (i.e.,

    amino acid residues) experienced a destabilizing hydrody-

    namic drag in elongational ow, only the terminal monomer

    experiences this force in uniform ow. Hence, in uniform

    ow, deformation of the protein follows an unzipping mech-

    anism where the b-strands are unfolded one at a time, lead-ing to the evolution of intermediate states. More importantly,

    it was evident that the nature of the applied external force

    dictated the preferred mechanism for protein unfolding.

    Szymczak and others observed similar ow effects on

    employing Brownian dynamics simulations to examine ow-

    induced stretching of integrin and ubiquitin.91,92 They noted

    that the hydrodynamic drag on a grafted protein chain in

    extensional ow increases along the chain length as the poly-

    mer unfolds. Therefore, the initiation of unfolding triggers a

    positive feedback mechanism which leads to the rapid unravel-

    ing of the entire chain to the exclusion of intermediate states.

    Indeed, it had previously been predicted that the uid drag

    acting on an initially coiled polymer increases as the molecule

    unfolds due to the decreased hydrodynamic interactions

    between its constituent monomers.84,93 In addition, the

    authors observed that whereas the grafted protein chains

    unfold through a series of intermediate states in uniform ow,

    consistent with that reported earlier,90 synthetic helices unrav-

    eled smoothly with no detectable intermediate states.91 This

    highlights the fact that proteins and homopolymers do indeed

    differ in their hydrodynamic properties in shear ow. Further-

    more, it was also noted that the array of metastable conforma-

    tions observed during protein unfolding in uniform ow was

    absent in mechanical pulling measurements in a force clamp

    apparatus such as the atomic force microscope.

    FIGURE 4 Change in the helix content of PLL as a function of

    shear strain. The gure shows that the a-helix-PLL structure (68.3kDa) is stable below a strain value of 105. For clarity, the shear rates

    plotted are: (&) 74 s21, (^) 302 s21, (*) 518 s21, and (l) 715s21. For clarity, error bars representing 6 standard deviation areshown only for the sample sheared at 518 s21.55

    Effects of Shear Flow on Protein Structure and Function 741

    Biopolymers

  • A subsequent study examined the inuence of hydrody-

    namic interactions, which facilitate the folding of protein

    molecules into compact native structures, on the mechani-

    cal stretching of proteins using ubiquitin as a model sys-

    tem.94 Mechanical stretching at constant speed, at constant

    force, and through uid ow was examined. It was found

    that hydrodynamic interactions (cohesive forces) between

    amino acid residues facilitated unfolding (stretching)

    when a constant speed or force was applied to grafted

    ubiquitin. This observation was explained on the basis

    that an amino acid residue, which is pulled away from

    the bulk protein, creates a ow which drags other residues

    with it. In contrast, these cohesive forces were found to

    counter uid forces by screening monomers concealed in

    the protein matrix from the drag in the ow eld, thereby

    hindering ubiquitin unfolding in both uniform and elon-

    gational ow elds.94 It has been reported that hydrody-

    namic interactions increase with increasing protein size

    and hence inuence the critical shear rate required to

    unfold a protein of radius R.88 However, the authors note

    that this prediction pertains to solutions with high solvent

    viscosity and low interfacial tension.

    Recently, Sulkowska and Cieplak applied the coarse-

    grained MD model to investigate the resistance of 7510 pro-

    teins to mechanical stretching at a constant speed.95 Pulling

    was done by the termini to determine the maximum force

    of resistance, Fmax, of the protein chains. They restricted

    their study to non-fragmented protein chains deposited in

    the Protein Data Bank by August, 2005 and between 40 and

    150 amino acid residues long. For specic sequential lengths

    of proteins, the average Fmax increased with increasing chain

    length. That is, longer protein chains showed better resist-

    ance relative to shorter chains. In relation to conforma-

    tional classes, a-proteins were found to be weak and gener-ated small Fmax values relative to b- and ab-proteins. Specif-ically, immunoglobulins and transport protein homologies

    were shown to yield larger Fmax values. However, the

    authors acknowledged that proteins with multi-a-domainsmay show a signicantly high resistance to mechanical

    stretching. In total, 134 proteins were considered to be

    strong with a threshold force of 170 pN. The authors pro-

    posed that the mechanical strength of the strongest proteins

    arises from a clamp consisting of long parallel b-strandswhich are stabilized by a network of intra- and intermolecu-

    lar interactions.

    Implications in Physiology and Bioprocessing

    Proteins are known to show a strong conformation-function

    relationship. Destabilization of the native protein structure

    can potentially lead to the evolution of dysfunctional, aggre-

    gation-prone species with undesirable effects.4 Reports in the

    literature demonstrate that shear, which is a common stress

    in both physiology20,58,96 and bioprocessing,13,18 can induce

    protein deformation.97 Therefore, it is important to analyze

    the relevance of shear stress to protein molecules under phys-

    iological ow and in bioprocessing.

    Pulsatile blood ow generates non-uniform shear stress

    throughout the arterial system during circulation.96 Arterial

    shear rates peak at 1640 s21, 96 reaching 104 s21, 58 in partly

    clogged coronary arterioles. The literature contains a signi-

    cant number of reports98107 (and references therein) linking

    the haemodynamic stress generated on arterial walls during

    microcirculation, specically in regions of relatively low

    shear stress, to the pathogenesis of atherosclerosis through

    the modulation of endothelial function.108 A characteristic

    hallmark of this disease state is the deposition of amyloid-

    like protein aggregates as plaque on arterial walls. Arterial

    amyloid-plaques may trigger vascular inammation, weaken

    the elasticity of blood vessels, and alter biochemical reactions

    such as lipid metabolism.109 Interestingly, it has been shown

    that molecular connement enhances the deformation of

    entangled polymers under squeeze ow,110 a condition which

    models pulsatile blood ow. Similarly, there are a consider-

    ably number of reports27,34,35,42,59,60,111113 citing haemody-

    namic shear stress as the primary facilitator of the proteolytic

    cleavage of large vWF multimers freshly released from endo-

    thelial cells as well as vWF-mediated platelet adhesion during

    haemostasis. The accumulation of large vWF multimers in

    the plasma may lead to disease states including thrombotic

    thrombocytopenic purpura, microvascular occlusion and

    atherosclerosis, due to the development of microvascular

    thrombi resulting from the complexation of the hyperactive

    vWF multimers and platelets at regions of low physiological

    shear.63,114 On the other hand, the smaller size of circulating

    vWF multimers, resulting from proteolysis, enhance their

    haemostatic potential in high shear regimes as occurs during

    arteriolar bleeding.62 The inuence of shear stress on the

    activity of platelets and vWF has been discussed in other

    review articles.57,58,63,115

    Others have linked the deterioration in cerebral microcir-

    culation (cerebral hypoperfusion) to the onset of protein

    conformational disorders such as Alzheimers disease.116118

    It is suggested that reduced cerebral blood ow may replace

    old age as a risk factor for Alzheimers disease.118 It also pro-

    posed that haemodynamic shear stress regulates insulin-like

    growth factor-I activity indirectly, and possibly its effect on

    vascular pathologies.119

    It is noteworthy that although protein conformational

    diseases and vascular disorders have received considerable

    attention in research, the role of physiological shear stress in

    742 Bekard et al.

    Biopolymers

  • the initiation and/or progression of these debilitating diseases

    remain to be fully understood.

    In the case of bioprocessing, shear stresses are present dur-

    ing the separation, purication, shipping, and handling of

    protein based products.13 For example, processing steps

    involving ultraltration reactors,120,121 stirred tanks,122,123

    homogenizers,124 and pumps53,125 induce high shear stresses.

    Simulated lobe pump studies have shown that high shear

    stress in these processing steps is sufcient to cause minor

    changes to protein structure.125 These structural changes

    may trigger the exposure of hitherto buried hydrophobic

    regions to the aqueous environment, initiating intra/inter-

    molecular hydrophobic interactions leading to aggregation.

    The simulation further suggests that shear stresses may have

    played a crucial role in the protein aggregation behavior

    observed in a previous lobe pump study.126 Indeed, protein

    aggregation and particle formation during processing can

    affect the efciency of protein recovery, efcacy, safety, and

    product shelf-life.11 Furthermore, therapeutic protein aggre-

    gates are associated with a range of clinical side effects

    including non-specic anti-complementary activity leading

    to anaphylactic shock.127,128 In addition, insoluble aggregates

    can cause immune responses, lodge in the lung capillary bed,

    and block blood vessels.129131

    The design of experiments to establish the effect of uid

    shear stress alone on protein structure and function has

    proved difcult. Existing shear devices generate other stresses

    including surface interactions and enhanced proteinair con-

    tacts, which complicates data interpretation.121,132,133 How-

    ever, it has been established that the shear-stability of a pro-

    tein molecule depends on: (i) its primary structure and mo-

    lecular weight (ii) the magnitude of shear strain and the

    duration of its application, and (iii) the viscosity of its sur-

    rounding medium.31,64,65 Therefore, these factors must be

    taken in consideration for a complete understanding of the

    shear-stability of protein systems. Advancing knowledge in

    this eld will be crucial for the design of processes, especially

    in the bioprocessing industry, to ensure protein stability,

    maintain functionality, and promote process yield.

    Future development of protein based products will

    increasingly require several areas of improvement, especially

    in-process-monitoring of the stability of protein solutions.

    This will require the development of experimental techniques

    to allow the measurement of shear stress on proteins of inter-

    est. This is imperative as several studies show that individual

    protein systems demonstrate unique ow properties in a

    given ow eld. Additionally, it will be essential to appreciate

    the shear ( _ct) thresholds beyond which particular proteinsystems destabilize in solution. This will inform the develop-

    ment of new processing equipment to improve efciency

    during production. It is noteworthy that knowledge from

    such studies could contribute to our understanding of the

    purported shear-induced pathogenesis of vascular and pro-

    tein conformational disorders in vivo.

    CONCLUSIONProtein unfolding and aggregation is an issue of concern

    both in vitro and in vivo, specically in bioprocessing and in

    the pathogenesis of disease. Shear rates less than 103 s21 have

    been found to signicantly alter the three-dimensional struc-

    ture of globular proteins, leading to aggregation and amyloid

    bril formation. The strain history ( _ct) is found to be an im-portant parameter in the shear-effect. Therefore, even in

    environments where ostensibly low shear is generated can

    lead to shear-induced protein structural change. Given the

    diversity of protein species, and the diversity of molecular

    sizes, conformations and nature of intramolecular interac-

    tions, the susceptibility of proteins to shear can vary. It is

    clear that high shear generating processes should be avoided

    in bioprocessing. Shear also exists in the circulatory system

    and could be a signicant contributor to the pathogenesis of

    certain diseases by facilitating the aggregation of aberrantly

    folded proteins.

    REFERENCES1. Petsko, G. A.; Ringe, D. Protein Structure and Function; New

    Science Press: London, 2004.

    2. Dobson, C. M. Trends in Biochemical Science 1999, 24,

    329332.

    3. Stefani, M.; Dobson, C. M. J Mol Med 2003, 81, 678699.

    4. Dobson, C. M. Nature 2003, 426, 884890.

    5. Chiti, F.; Dobson, C. M. Annu Rev Biochem 2006, 75,

    333366.

    6. Calamai, M.; Chiti, F.; Dobson, C. M. Biophys J 2005, 89,

    42014210.

    7. Wang, W. Int J Pharm 1999, 185, 129188.

    8. Bryant, Z.; Pande, V. S.; Rokhsar, D. S. Biophys J 2000, 78,

    584589.

    9. Li, P.-C.; Makarov, D. E. J Chem Phys 2003, 119, 92609268.

    10. Rohs, R.; Etchebest, C.; Lavery, R. Biophys J 1999, 76,

    27602768.

    11. Wang, W. Int J Pharm 2005, 289, 130.

    12. Morel, M.-H.; Redl, A.; Guilbert, S. Biomacromolecules 2002,

    3, 488497.

    13. Thomas, C. R. In Chemical Engineering Problems in Biotech-

    nology; Winkler, M. A., Ed.; Elsevier Applied Science: London,

    1991, pp 2393.

    14. Bekard, I.; Dunstan, D. E. J Phys Chem B Lett 2009, 113,

    84538457.

    15. Dunstan, D. E.; Hill, E. K.; Wei, Y. Macromolecules 2004, 37,

    16631665.

    16. Smith, D. E.; Babcock, H. P.; Chu, S. Science 1999, 283,

    17241727.

    Effects of Shear Flow on Protein Structure and Function 743

    Biopolymers

  • 17. Elias, C. B.; Joshi, J. B. Adv Biochem Eng Biotechnol 1997, 59,

    4772.

    18. Yim, S. S.; Shamlou, P. A. Adv Biochem Eng Biotechnol 2000,

    67, 83122.

    19. Lin, J.; Meyer, J. D.; Carpenter, J. F.; Manning, M. C. Pharm

    Res 2000, 17, 391396.

    20. Tangelder, G. J.; Slaaf, D. W.; Arts, T.; Reneman, R. S. Am J

    Physiol 1988, 254, H1059H1064.

    21. Perkins, T. T.; Smith, D. E.; Chu, S. Science 1997, 276,

    20162021.

    22. Danker, G.; Vlahovska, P. M.; Misbah, C. Phys Rev Lett 2009,

    102, 148102.

    23. Munson, B. R.; Young, D. F.; Okiishi, T. H.; Huebsch, W. W.

    Fundamentals of Fluid Mechanics; Wiley: Hoboken, New

    Jersey, 2009.

    24. Lee, A. T.; McHugh, A. J. Biopolymers 1999, 50, 589594.

    25. Hill, E. K.; Krebs, B.; Goodall, D. G.; Howlett, G. J.; Dunstan,

    D. E. Biomacromolecules 2006, 7, 1013.

    26. Dunstan, D. E.; Hamilton-Brown, P.; Asimakis, P.; Ducker, W.;

    Bertolini, J. Soft Matter 2009, 5, 50205028.

    27. Singh, I.; Themistou, E.; Porcar, L.; Neelamegham, S. Biophys J

    2009, 96, 23132320.

    28. Dunstan, D. E.; Hamilton-Brown, P.; Asimakis, P.; Ducker, W.

    A.; Bertolini, J. Protein Eng Des Sel 2009, 22, 741746.

    29. Imomoh, E.; Dusting, J.; Ashton, L.; Blanch, E.; Balabani, S.

    14th International Symposium on Applications of Laser Tech-

    niques to Fluid Mechanics, Lisbon, Portugal, 2008.

    30. Ashton, L.; Dusting, J.; Imomoh, E.; Balabani, S.; Blanch, E. W.

    Biophys J 2009, 96, 42314236.

    31. van der Veen, M. E.; van Iersel, D. G.; van der Goot, A. J.;

    Boom, R. M. Biotechnol Progr 2004, 20, 11401145.

    32. Charm, S. E.; Wong, B. L. Science 1970, 170, 466468.

    33. Charm, S. E.; Lai, C. J. Biotechnol Bioeng 1971, 13, 185202.

    34. Peterson, D. M.; Stathopulos, N. A.; Giorgio, T. D.; Hellums, J.

    D.; Moake, J. L. Blood 1987, 69, 625628.

    35. Ikeda, Y.; Handa, M.; Kawano, K.; Kamata, T.; Murata, M.;

    Araki, Y.; Anbo, H.; Kawai, Y.; Watanabe, K.; Itagaki, I.; Sakai,

    K.; Ruggeri, Z. M. J Clin Investig 1991, 87, 12341240.

    36. Thomas, C. R.; Dunnill, P. Biotechnol Bioeng 1979, 21, 2279

    2302.

    37. Thomas, C. R.; Nienow, A. W.; Dunnill, P. Biotechnol Bioeng

    1979, 21, 22632279.

    38. Maa, Y.; Hsu, C. C. Biotechnol Bioeng 1996, 51, 458465.

    39. Tirrell, M.; Middleman, S. Biotechnol Bioeng 1975, 17, 299303.

    40. Charm, S. E.; Wong, B. L. Biotechnol Bioeng 1970, 12,

    11031109.

    41. Jaspe, J.; Hagen, S. J. Biophys J 2006, 91, 34153424.

    42. Schneider, S. W.; Nuschele, S.; Wixforth, A.; Gorzelanny, C.;

    Alexander-Katz, A.; Netz, R. R.; Schneider, M. F. Proc Natl

    Acad Sci 2007, 104, 78997903.

    43. Chien, S.; Skalak, R. Biorheology 1981, 18, 307330.

    44. Bell, D. J.; Dunnill, P. Biotechnol Bioeng 1982, 24, 12711285.

    45. Cathey, C. A.; Fuller, G. G. J Non-Newtonian Fluid Mechanics

    1990, 34, 6388.

    46. Rodel, W. In Instrumentation and Sensors for the Food Indus-

    try; Kress-Rogers, E., Ed.; Butterworth-Heinemann: Oxford,

    1993, pp 375415.

    47. Hill, E. K.; Watson, R. L.; Dunstan, D. E. J Fluorescence 2005,

    15, 255266.

    48. Guyon, E.; Hulin, J. P.; Petit, L.; Mitescu, C. D. Physical Hydro-

    dynamics; Oxford University Press: New York, 2001.

    49. Taylor, G. I. Philosophical Transactions Royal Society London

    Series A 1923, 223, 289343.

    50. van Wazer, J. R.; Lyon, J. W.; Kim, K. Y.; Colwell, R. E. Viscosity

    and Flow Measurement; Interscience Publishers: New York,

    1963.

    51. Charm, S. E.; Wong, B. L. Biotechnol Bioeng 1978, 20,

    451453.

    52. Bowen, W. R.; Gan, Q. Biotechnol Bioeng 1992, 40, 491497.

    53. Virkar, P. D.; Narendranathan, T. J.; Hoare, M.; Dunnill, P.

    Biotechnol Bioeng 1981, 23, 425429.

    54. Talboys, B. L.; Dunnill, P. Biotechnol Bioeng 1985, 27,

    17301734.

    55. Bekard, I. B.; Barnham, K. J.; White, L. R.; Dunstan, D. E. Soft

    Matter 2011, 7, 203210.

    56. Dong, J.; Moake, J. L.; Nolasco, L.; Bernardo, A.; Arceneaux,

    W.; Shrimpton, C. N.; Schade, A. J.; McIntire, L. V.; Fujikawa,

    K.; Lopez, J. A. Blood 2002, 100, 4033.

    57. Sadler, J. E. Annu Rev Biochem 1998, 67, 395424.

    58. Kroll, M. H.; Hellums, J. D.; McIntire, L. V.; Schafer, A. I.;

    Moake, J. L. Blood 1996, 88, 15251541.

    59. Tsai, H. M.; Sussman, I. I.; Nagel, R. L. Blood 1994, 83,

    2171.

    60. Siedlecki, C. A.; Lestini, B. J.; Kottke-Marchant, K.; Eppell, S.

    J.; Wilson, D. L.; Marchant, R. E. Blood 1996, 88, 29392950.

    61. Mody, N. A.; Lomakin, O.; Doggett, T. A.; Diacovo, T. G.;

    King, M. R. Biophys J 2005, 88, 14321443.

    62. Zhang, X.; Halvorsen, K.; Zhang, C. Z.; Wong, W. P.; Springer,

    T. A. Science 2009, 324, 1330.

    63. Sadler, J. E. Annu Rev Med 2005, 56, 173191.

    64. Maa, Y.; Hsu, C. C. Biotechnol Bioeng 1997, 54, 503512.

    65. Oliva, A.; Santovena, A.; Farina, J.; Llabres, M. J Pharm

    Biomed Anal 2003, 33, 145155.

    66. Ramirez-Alvarado, M.; Merkel, J. S.; Regan, L. Proc Natl Acad

    Sci 2000, 97, 89798984.

    67. Ahmad, A.; Uversky, V. N.; Hong, D.; Fink, A. L. J Biol Chem

    2005, 280, 42669.

    68. Nielsen, L.; Khurana, R.; Coats, A.; Frokjaer, S.; Brange, J.;

    Vyas, S.; Uversky, V. N.; Fink, A. L. Biochemistry 2001, 40,

    60366046.

    69. Grainger, R. J.; Ko, S.; Koslov, E.; Prokop, A.; Tanner, R. D.;

    Loha, V. Appl Biochem Biotechnol 2000, 84, 761768.

    70. Sluzky, V.; Tamada, J. A.; Klibanov, A. M.; Langer, R. Proc Natl

    Acad Sci USA 1991, 88, 93779381.

    71. Sluzky, V.; Klibanov, A. M.; Langer, R. Biotechnol Bioeng 1992,

    40, 895903.

    72. Hamilton-Brown, P.; Bekard, I.; Ducker, W. A.; Dunstan, D. E.

    J Phys Chem B 2008, 112, 1624916252.

    73. Sasahara, K.; Yagi, H.; Sakai, M.; Naiki, H.; Goto, Y. Biochemis-

    try 2008, 47, 26502660.

    74. Ohhashi, Y.; Kihara, M.; Naiki, H.; Goto, Y. J Biol Chem 2005,

    280, 32843.

    75. Sicorello, A.; Torrassa, S.; Soldi, G.; Gianni, S.; Travaglini-Allo-

    catelli, C.; Taddei, N.; Relini, A.; Chiti, F. Biophys J 2009, 96,

    22892298.

    76. Stathopulos, P. B.; Scholz, G. A.; Hwang, Y. M.; Rumfeldt,

    J. A. O.; Lepock, J. R.; Meiering, E. M. Protein Sci 2004, 13,

    30173027.

    744 Bekard et al.

    Biopolymers

  • 77. Akkermans, C.; Venema, P.; Rogers, S. S.; van der Goot, A.

    J.; Boom, R. M.; van der Linden, E. Food Biophys 2006, 1,

    144150.

    78. Akkermans, C.; van der Goot, A. J.; Venema, P.; van der

    Linden, E.; Boom, R. M. Food Hydrocolloids 2008, 22, 1315

    1325.

    79. Alexander-Katz, A.; Netz, R. R. Europhys Lett (EPL) 2007, 80,

    18001.

    80. Bee, J. S.; Stevenson, J. L.; Mehta, B.; Svitel, J.; Pollastrini, J.;

    Platz, R.; Freund, E.; Carpenter, J. F.; Randolph, T. W. Biotech-

    nol Bioeng 2009, 103, 936.

    81. Adamson, A. W.; Gast, A. P.; NetLibrary, I. Physical Chemistry

    of Surfaces; Wiley: New York, 1997.

    82. Best, R. B.; Brockwell, D. J.; Toca-Herrera, J. L.; Blake, A. W.;

    Smith, D. A.; Radford, S. E.; Clarke, J. Anal Chim Acta 2003,

    479, 87105.

    83. Agarwal, U. S.; Mashelkar, R. A. J Chem Phys 1994, 100,

    60556061.

    84. De Gennes, P. G. J Chem Phys 1974, 60, 50305042.

    85. Greeneld, N. J.; Fasman, G. D. Biochemistry 1969, 8, 4108

    4116.

    86. JiJi, R. D.; Balakrishnan, G.; Hu, Y.; Spiro, T. G. Biochemistry

    2006, 45, 3441.

    87. Immaneni, A.; McHugh, A. J. Biopolymers 1998, 45, 239

    246.

    88. Alexander-Katz, A.; Netz, R. R. Macromolecules 2008, 41,

    33633374.

    89. Buguin, A.; Brochard-Wyart, F. Macromolecules 1996, 29,

    49374943.

    90. Lemak, A. S.; Lepock, J. R.; Chen, J. Z. Y. Proteins Struct Funct

    Genet 2003, 51, 224235.

    91. Szymczak, P.; Marek, C. J Chem Phys 2006, 125, 164903.

    92. Szymczak, P.; Cieplak, M. J Chem Phys 2007, 127, 155106.

    93. Agarwal, U. S.; Rohit, B.; Mashelkar, R. A. J Chem Phys 1998,

    108, 16101617.

    94. Szymczak, P.; Cieplak, M. J Phys Condensed Matter 2007, 19,

    285224285235.

    95. Sulkowska, J. I.; Cieplak, M. Biophys J 2008, 94, 613.

    96. Stroev, P. V.; Hoskins, P. R.; Easson, W. J. Atherosclerosis 2007,

    191, 276280.

    97. McNally, E.; Lockwod, C. In Protein Formulation and

    Delivery; McNally, E., Ed.; Marcel Dekker: New York, 2000,

    pp 111138.

    98. Ge, J.; Erbel, R.; Gorge, G.; Haude, M.; Meyer, J. Br Heart J

    1995, 73, 462465.

    99. Ku, D. N.; Giddens, D. P.; Zarins, C. K.; Glagov, S. Arterioscler

    Thromb Vasc Biol 1985, 5, 293302.

    100. Malek, A. M.; Alper, S. L.; Izumo, S. JAMA 1999, 282,

    20352042.

    101. Cunningham, K. S.; Gotlieb, A. I. Lab Invest 2004, 85, 9

    23.

    102. Krams, R.; Wentzel, J. J.; Oomen, J. A. F.; Vinke, R.; Schuurb-

    iers, J. C. H.; De Feyter, P. J.; Serruys, P. W.; Slager, C. J. Arte-

    rioscler Thromb Vasc Biol 1997, 17, 2061.

    103. Gibson, C. M.; Diaz, L.; Kandarpa, K.; Sacks, F. M.; Pasternak,

    R. C.; Sandor, T.; Feldman, C.; Stone, P. H. Arterioscler

    Thromb Vasc Biol 1993, 13, 310.

    104. Shaaban, A. M.; Duerinckx, A. J. Am J Roentgenol 2000, 174,

    1657.

    105. Vita, J. A.; Treasure, C. B.; Ganz, P.; Cox, D. A.; David Fish, R.;

    Selwyn, A. P. J Am Coll Cardiol 1989, 14, 11931199.

    106. Davies, P. F.; Remuzzi, A.; Gordon, E. J.; Dewey, C. F.;

    Gimbrone, M. A. Proc Natl Acad Sci USA 1986, 83, 2114.

    107. Irace, C.; Cortese, C.; Fiaschi, E.; Carallo, C.; Farinaro, E.;

    Gnasso, A. Stroke 2004, 35, 464.

    108. Jiang, Y.; Kohara, K.; Hiwada, K. Stroke 2000, 31, 23192324.

    109. Howlett, G. J.; Moore, K. J. Curr Opin Lipidol 2006, 17, 541.

    110. Rowland, H. D.; King, W. P.; Cross, G. L. W.; Pethica, J. B. ACS

    Nano 2008, 2, 419428.

    111. Alevriadou, B. R.; Moake, J. L.; Turner, N. A.; Ruggeri, Z. M.;

    Folie, B. J.; Phillips, M. D.; Schreiber, A. B.; Hrinda, M. E.;

    McIntire, L. V. Blood 1993, 81, 1263.

    112. Goto, S.; Salomon, D. R.; Ikeda, Y.; Ruggeri, Z. M. J Biol Chem

    1995, 270, 23352.

    113. Chow, T. W.; Hellums, J. D.; Moake, J. L.; Kroll, M. H. Blood

    1992, 80, 113.

    114. Tsai, H. M. Semin Thromb Hemost 2003, 29, 479488.

    115. Reininger, A. J. Haemophilia 2008, 14, 1126.

    116. de la Torre, J. C. Lancet Neurol 2004, 3, 184190.

    117. de La Torre, J. C. Neurol Res 2004, 26, 517524.

    118. Crawford, J. G. Med Hypotheses 1998, 50, 2536.

    119. Elhadj, S.; Akers, R. M.; Forsten-Williams, K. Ann Biomed Eng

    2003, 31, 163170.

    120. Ahrer, K.; Buchacher, A.; Iberer, G.; Jungbauer, A. J Membr Sci

    2006, 274, 108115.

    121. Kim, K. J.; Chen, V.; Fane, A. G. Biotechnol Bioeng 1993, 42,

    260265.

    122. Cutter, L. A. AIChE J 1966, 12, 3545.

    123. Kresta, S. M.; Wood, P. E. AIChE J 1991, 37, 448460.

    124. Floury, J.; Bellettre, J.; Legrand, J.; Desrumaux, A. Chem Eng

    Sci 2004, 59, 843853.

    125. Gomme, P. T.; Prakash, M.; Hunt, B.; Stokes, N.; Cleary, P.;

    Tatford, O. C.; Bertolini, J. Biotechnol Appl Biochem 2006, 43,

    113120.

    126. Gomme, P. T.; Hunt, B.; Tatford, O. C.; Johnston, A.; Bertolini,

    J. Biotechnol Appl Biochem 2006, 43, 103111.

    127. Thornton, C. A.; Ballow, M. Arch Neurol 1993, 50, 135136.

    128. Wang, Y.-C. J.; Hanson, M. A. J Parenteral Sci Technol 1988,

    42, S3S26.

    129. Katakam, M.; Banga, A. K. J Pharm Pharmacol 1995, 47,

    103107.

    130. Cleland, J. L.; Powell, M. F.; Shire, S. J. Crit Rev Ther Drug

    Carrier Syst 1993, 10, 307377.

    131. Dorris, G. G.; Bivins, B. A.; Rapp, R. P.; Weiss, D. L.; DeLuca,

    P. P.; Ravin, M. B. Anesth Analg 1977, 56, 422427.

    132. Meireles, M.; Aimar, P.; Sanchez, V. Biotechnol Bioeng 1991,

    38, 528534.

    133. Narendranathan, T. J.; Dunnill, P. Biotechnol Bioeng 1982, 24,

    21032107.

    Reviewing Editor: Stephen Blacklow

    Effects of Shear Flow on Protein Structure and Function 745

    Biopolymers