To the editor:

The topical application of ferric chloride (FeCl3) to the vasculature is one of the most commonly used experimental approaches to induce thrombosis. The method was first described by Kurz and colleagues1  and has subsequently been proven to be a highly effective and reliable approach to elucidate the role of platelet receptors, ligands, and activation pathways in promoting thrombosis.2-7  It has also shed new light on the role of coagulation proteases in regulating thrombin generation and thrombus growth8-11  and has been used in the context of thrombolysis,12-17  although the model appears to have significant limitations in this regard. Despite its widespread use, the precise mechanism by which FeCl3 induces thrombosis remains controversial. A recent study in Blood by Ciciliano and colleagues has provided further insight into the molecular mechanisms underlying FeCl3-induced thrombosis, suggesting an important role for charge-dependent aggregation effects of FeCl3 on blood cells and plasma proteins.18 

It has long been assumed that the major effects of FeCl3 are limited to the vasculature. FeCl3 ions have been localized to the endothelium with uptake of iron through endothelial pinocytic processes. Several studies have described ferric ion-rich membrane-enclosed bodies which transmigrate into the endothelium, followed by exocytosis into the vessel lumen.19,20  It was assumed that this accumulation of iron and generation of reactive oxygen species produce endothelial toxicity and denudation, leading to the exposure of subendothelial elements that promote thrombus formation.1,21  However, a number of studies have revealed minor endothelial denudation and collagen exposure following FeCl3 treatment,20,22,23  raising the possibility that FeCl3 has effects beyond the vessel wall.

The first demonstration that red blood cells (RBCs) may play an important role in promoting FeCl3-induced thrombosis was derived from our in vitro studies using isolated blood cell components perfused through mouse aortae ex vivo.23  Surprisingly, endothelial cells exposed to FeCl3 alone exhibited minor levels of injury. However, in the presence of whole blood or isolated RBCs, FeCl3-induced red cell hemolysis and hemoglobin oxidation promoted extensive vascular injury and thrombosis.23  Elegant electron microscopy studies by Barr and colleagues confirmed extensive red cell accumulation on the endothelium following FeCl3 exposure in vivo, with platelets rapidly recruited to accumulated red cell-derived structures.22  Eckly and colleagues revealed surface expression of tissue factor on the ferric ion-rich spherical bodies, which they attributed as the primary mediator of platelet adhesion and fibrin formation.20 

The plot thickens further. Ciciliano and colleagues used microfluidic devices coated with endothelial cells to dissect the effects of FeCl3 on individual blood cell and plasma components.18  These studies demonstrated concentration-dependent effects of FeCl3 on protein and blood cell aggregation, independent of effects on the endothelium. This aggregation effect was principally attributed to colloidal chemistry, whereby cells and proteins adhere and aggregate as a result of their charge. The authors have proposed that this physiochemical effect of FeCl3 on blood cells is the primary instigator driving blood cell adhesion to the endothelium. They argue that this mechanism then facilitates the “secondary” phase of FeCl3 injury, with red cell aggregates and damaged endothelium providing a reactive surface for the accumulation of platelets and initiation of blood coagulation, necessary for stable thrombus formation.19-21  However, it remains to be determined to what extent oxidative damage vs physicochemical effects predominate to induce RBC aggregation and thrombosis.23  In this context, aluminum chloride (AlCl3), which carries a similar charge to Fe3+ and was used as an additional means of evidence to support the role of colloidal chemistry in cellular and protein aggregation, is also well known to cause oxidative damage, including lipid peroxidation and RBC hemolysis.24  This is in contrast to chromium chloride (CrCl3), which was also used in this study as a negative control; however, this molecule has antioxidant properties.25  Whether the effects of AlCl3 can be offset by antioxidants, as was demonstrated for FeCl3 in our own ex vivo studies using isolated aorta,23  will be important to determine. It is also interesting to note that the oxidative effects of FeCl3 we observed in isolated aorta were initiated with concentrations of FeCl3 lower than that observed to induce macroscopic precipitation of plasma proteins.23 

Collectively, the studies highlighted above unequivocally demonstrate that the effects of FeCl3 on vascular injury and thrombosis are multifaceted and far more complex than originally envisioned. It will be a challenge to define a precise, unified mechanism of FeCl3-induced thrombosis because both the physicochemical and pro-oxidant effects of FeCl3 on blood cells and the vasculature are highly dependent on FeCl3 concentration and exposure time. Whether simplified in vitro models that use microfluidics and 3-dimensional printing approaches can accurately recapitulate the complex changes operating in the vasculature of a living animal remains to be seen. The dynamic interface between blood cells, subendothelial elements, and vascular reactivity is central to the thrombotic response, and therefore it will be important to demonstrate that the thrombosis mechanisms operating in microfluidic devices are similar to those occurring in isolated vessel segments. Nonetheless, it is likely that FeCl3 as an inducer of experimental thrombosis is here to stay, due to its simplicity, widespread availability, and ease of use. However, as noted by numerous authors,13,18,20-23  caution should be used when interpreting data from such a model, and particularly when attempting to draw generalizing conclusions about the proposed mechanisms regulating coagulation and blood cell interactions with the vessel wall during thrombus initiation.

It is notable that considerable effort has been made to avoid experimental bias in in vivo studies, by limiting genetic variability, sex differences, and the impact of diet, pathogens, and age-related changes in the mouse. However, it could be argued that we do not apply the same level of rigor to our experimental thrombosis models, all too regularly accepting the findings from a single in vivo thrombosis model. This, of course, is not unique to our field. However, an improved understanding of the molecular mechanisms promoting thrombosis in specific models and increased recognition of the pitfalls and limitations of each of our models, coupled with a requirement to report data using several distinct thrombosis models, should help enhance the veracity of our experimental findings and reduce future controversy. Only time will tell.

Acknowledgments: This work was supported by the National Health and Medical Research Council (NHMRC).

Contribution: S.M.S. and S.P.J. wrote the paper.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Shaun P. Jackson, Heart Research Institute and Charles Perkins Centre, The University of Sydney, Level 3, Building D17, Orphan School Creek Rd, Camperdown, NSW 2006, Australia; e-mail: shaun.jackson@sydney.edu.au.

1
Kurz
 
KD
Main
 
BW
Sandusky
 
GE
Rat model of arterial thrombosis induced by ferric chloride.
Thromb Res
1990
, vol. 
60
 
4
(pg. 
269
-
280
)
2
André
 
P
Prasad
 
KS
Denis
 
CV
, et al. 
CD40L stabilizes arterial thrombi by a beta3 integrin--dependent mechanism.
Nat Med
2002
, vol. 
8
 
3
(pg. 
247
-
252
)
3
Bergmeier
 
W
Piffath
 
CL
Goerge
 
T
, et al. 
The role of platelet adhesion receptor GPIbalpha far exceeds that of its main ligand, von Willebrand factor, in arterial thrombosis.
Proc Natl Acad Sci USA
2006
, vol. 
103
 
45
(pg. 
16900
-
16905
)
4
Chauhan
 
AK
Kisucka
 
J
Lamb
 
CB
Bergmeier
 
W
Wagner
 
DD
von Willebrand factor and factor VIII are independently required to form stable occlusive thrombi in injured veins.
Blood
2007
, vol. 
109
 
6
(pg. 
2424
-
2429
)
5
Marx
 
I
Christophe
 
OD
Lenting
 
PJ
, et al. 
Altered thrombus formation in von Willebrand factor-deficient mice expressing von Willebrand factor variants with defective binding to collagen or GPIIbIIIa.
Blood
2008
, vol. 
112
 
3
(pg. 
603
-
609
)
6
Ni
 
H
Denis
 
CV
Subbarao
 
S
, et al. 
Persistence of platelet thrombus formation in arterioles of mice lacking both von Willebrand factor and fibrinogen.
J Clin Invest
2000
, vol. 
106
 
3
(pg. 
385
-
392
)
7
Ni
 
H
Yuen
 
PS
Papalia
 
JM
, et al. 
Plasma fibronectin promotes thrombus growth and stability in injured arterioles.
Proc Natl Acad Sci USA
2003
, vol. 
100
 
5
(pg. 
2415
-
2419
)
8
Cheng
 
Q
Tucker
 
EI
Pine
 
MS
, et al. 
A role for factor XIIa-mediated factor XI activation in thrombus formation in vivo.
Blood
2010
, vol. 
116
 
19
(pg. 
3981
-
3989
)
9
Moller
 
F
Tranholm
 
M
A ferric chloride induced arterial injury model used as haemostatic effect model.
Haemophilia
2010
 
16(1):e216-e222
10
Renné
 
T
Pozgajová
 
M
Grüner
 
S
, et al. 
Defective thrombus formation in mice lacking coagulation factor XII.
J Exp Med
2005
, vol. 
202
 
2
(pg. 
271
-
281
)
11
Wang
 
X
Cheng
 
Q
Xu
 
L
, et al. 
Effects of factor IX or factor XI deficiency on ferric chloride-induced carotid artery occlusion in mice.
J Thromb Haemost
2005
, vol. 
3
 
4
(pg. 
695
-
702
)
12
Karatas
 
H
Erdener
 
SE
Gursoy-Ozdemir
 
Y
, et al. 
Thrombotic distal middle cerebral artery occlusion produced by topical FeCl(3) application: a novel model suitable for intravital microscopy and thrombolysis studies.
J Cereb Blood Flow Metab
2011
 
31(6):1452-1460
13
Machlus
 
KR
Cardenas
 
JC
Church
 
FC
Wolberg
 
AS
Causal relationship between hyperfibrinogenemia, thrombosis, and resistance to thrombolysis in mice.
Blood
2011
, vol. 
117
 
18
(pg. 
4953
-
4963
)
14
Sheffield
 
WP
Eltringham-Smith
 
LJ
Gataiance
 
S
Bhakta
 
V
A plasmin-activatable thrombin inhibitor reduces experimental thrombosis and assists experimental thrombolysis in murine models.
J Thromb Thrombolysis
2015
, vol. 
39
 
4
(pg. 
443
-
451
)
15
Wang
 
X
Palasubramaniam
 
J
Gkanatsas
 
Y
, et al. 
Towards effective and safe thrombolysis and thromboprophylaxis: preclinical testing of a novel antibody-targeted recombinant plasminogen activator directed against activated platelets.
Circ Res
2014
, vol. 
114
 
7
(pg. 
1083
-
1093
)
16
Zhu
 
Y
Carmeliet
 
P
Fay
 
WP
Plasminogen activator inhibitor-1 is a major determinant of arterial thrombolysis resistance.
Circulation
1999
, vol. 
99
 
23
(pg. 
3050
-
3055
)
17
Kim
 
YD
Nam
 
HS
Kim
 
SH
, et al. 
Time-dependent thrombus resolution after tissue-type plasminogen activator in patients with stroke and mice.
Stroke
2015
, vol. 
46
 
7
(pg. 
1877
-
1882
)
18
Ciciliano
 
JC
Sakurai
 
Y
Myers
 
DR
, et al. 
Resolving the multifaceted mechanisms of the ferric chloride thrombosis model using an interdisciplinary microfluidic approach.
Blood
2015
, vol. 
126
 
6
(pg. 
817
-
824
)
19
Tseng
 
MT
Dozier
 
A
Haribabu
 
B
Graham
 
UM
Transendothelial migration of ferric ion in FeCl3 injured murine common carotid artery.
Thromb Res
2006
, vol. 
118
 
2
(pg. 
275
-
280
)
20
Eckly
 
A
Hechler
 
B
Freund
 
M
, et al. 
Mechanisms underlying FeCl3-induced arterial thrombosis.
J Thromb Haemost
2011
, vol. 
9
 
4
(pg. 
779
-
789
)
21
Li
 
W
McIntyre
 
TM
Silverstein
 
RL
Ferric chloride-induced murine carotid arterial injury: A model of redox pathology.
Redox Biol
2013
, vol. 
1
 (pg. 
50
-
55
)
22
Barr
 
JD
Chauhan
 
AK
Schaeffer
 
GV
Hansen
 
JK
Motto
 
DG
Red blood cells mediate the onset of thrombosis in the ferric chloride murine model.
Blood
2013
, vol. 
121
 
18
(pg. 
3733
-
3741
)
23
Woollard
 
KJ
Sturgeon
 
S
Chin-Dusting
 
JP
Salem
 
HH
Jackson
 
SP
Erythrocyte hemolysis and hemoglobin oxidation promote ferric chloride-induced vascular injury.
J Biol Chem
2009
, vol. 
284
 
19
(pg. 
13110
-
13118
)
24
Lukyanenko
 
LM
Skarabahatava
 
AS
Slobozhanina
 
EI
Kovaliova
 
SA
Falcioni
 
ML
Falcioni
 
G
In vitro effect of AlCl3 on human erythrocytes: changes in membrane morphology and functionality.
J Trace Elem Med Biol
2013
 
27(2):160-167
25
Jain
 
SK
Patel
 
P
Rogier
 
K
Jain
 
SK
Trivalent chromium inhibits protein glycosylation and lipid peroxidation in high glucose-treated erythrocytes.
Antioxid Redox Signal
2006
, vol. 
8
 
1-2
(pg. 
238
-
241
)
Sign in via your Institution