Download PDF
Review  |  Open Access  |  14 Apr 2015

Inflammation and intracranial aneurysms: mechanisms of initiation, growth, and rupture

Views: 6505 |  Downloads: 1064 |  Cited:   3
Neurosciences 2015;2:68-76.
Author Information
Article Notes
Cite This Article

Abstract

Outcomes following aneurysmal subarachnoid hemorrhage remain poor in many patients, despite advances in microsurgical and endovascular management. Consequently, considerable effort has been placed in determining the mechanisms of aneurysm formation, growth, and rupture. Various environmental and genetic factors are implicated as key components in the aneurysm pathogenesis. Currently, sufficient evidence exists to incriminate the inflammatory response as the common pathway leading to aneurysm generation and rupture. Central to this model is the interaction between the vessel wall and inflammatory cells. Dysfunction of the endothelium and vascular smooth muscle cells (VSMCs) promotes a chronic pathological inflammatory response that progressively weakens the vessel wall. We review the literature pertaining to the cellular and chemical mechanisms of inflammation that contribute to aneurysm development. Hemodynamic stress and alterations in blood flow are discussed regarding their role in promoting chronic inflammation. Endothelial cell and VSMC dysfunction are examined concerning vascular remodeling. The contribution of inflammatory cytokines, especially tumor necrosis factor-α is illustrated. Inflammatory cell infiltration, particularly macrophage-mediated deterioration of vascular integrity, is reviewed. We discuss the inflammation as a means to determine aneurysms at greatest risk of rupture. Finally, future therapeutic implications of pharmacologic modulation of the inflammation are discussed.

Keywords

Aneurysm, endothelium, inflammation, subarachnoid hemorrhage, vascular smooth muscle cells

Introduction

Intracranial aneurysms and subarachnoid hemorrhage (SAH) represent significant disease entities, with ruptured aneurysms accounting for severe disability and death in approximately 27,000 Americans each year.[1] Among the general population, approximately 4-5% of individuals harbor an unruptured aneurysm.[2] However, with the evolution of sophisticated imaging modalities, such as magnetic resonance imaging (MRI), magnetic resonance angiography, and computed tomographic angiography, unruptured incidental aneurysms are continually discovered with increasing frequency. Microsurgical and endovascular obliteration remains the mainstays of treatment, yet both are associated with a significant risk of morbidity and mortality.[3,4] In addition, a significant number of unruptured aneurysms are treated preemptively to avoid the catastrophic sequelae associated with SAH. Despite the prevalence of these lesions, the often-devastating nature of the disease, and the risks associated with treatment, relatively little is known about the aneurysm pathogenesis and natural history. As a result, there has been a significant effort to define the mechanisms underlying aneurysm formation and growth. Although the evolution of aneurysms from initiation to rupture is undoubtedly multifactorial in nature, the inflammatory response appears to play a critical role in the pathogenesis of these lesions.

Further understanding of the relationship between the inflammatory response and aneurysm evolution may have important clinical implications in the future. Identification of patients prone to pathologic inflammatory states could allow detection of a population more likely to suffer aneurysm rupture. Additionally, the genetic and cellular processes mediating inflammation represent attractive targets for possible pharmacologic intervention. We review the current literature pertaining to the role of inflammation in the generation of aneurysm formation and rupture. The various vascular inflammatory stimuli are discussed, with special attention paid to hemodynamic stress and alterations in blood flow. Endothelial celland vascular smooth muscle cell (VSMC) dysfunction are detailed and examined in relation to vascular remodeling. The contribution of multiple cytokines to a sustained pathologic inflammatory state and their influence on aneurysm formation are outlined. Special emphasis is placed on a key inflammatory mediator, tumor necrosis factor-α (TNF-α). The contribution of inflammatory cell infiltration, particularly the potential of macrophage-mediated rupture, is detailed. Finally, future therapeutic implications of pharmacologic modulation of the inflammatory response are discussed.

Intracranial aneurysms and the inflammatory response

Multiple studies have identified various risk factors for aneurysmal expansion and rupture.[5] Genetics plays an important role, with approximately 10% of SAH patients having two or more family members also affected by unruptured or ruptured aneurysms.[6,7] The propensity for SAH within specific ethnic groups, particularly the Finish and Japanese populations, further highlights the contribution of genetics.[8,9] Gender also appears to play a role, as women appear to more frequently develop intracranial aneurysms and perhaps suffer from ruptured aneurysms more often than men.[10,11] Environmental factors, particularly smoking, have been clearly linked to a higher incidence of SAH.[11] Additional studies have linked binge drinking to aneurysm rupture.[12,13] Chronically uncontrolled hypertension clearly correlates with aneurysm formation in animal models and clinical studies have identified hypertension as a risk factor for aneurysmal SAH.[10,11,14-16]

Due to the variability in contributing risk factors, attempts have been made to identify a unifying underlying pathophysiologic mechanism that promotes aneurysm formation and rupture.[5] There is a tremendous mounting body of evidence that the inflammatory response represents a common endpoint that drives aneurysm evolution, which is succinctly summarized as initiation of development, growth, and potential rupture.[17] Hemodynamic stress and disruption of blood flow, oxidative stress, injurious environmental elements (i.e. cigarette smoking and cocaine), and pro-inflammatory genetic alterations all initiate a sustained and pathologic inflammatory response.[18,19] As a result, the intracranial vasculature is subjected to endothelial cell dysfunction, elevated inflammatory cell infiltration, detrimental changes within the tunica media, and exposure to increased concentrations of proteases. These processes lead to weakening, dilation, and remodeling of the vessel walls, which are key components in aneurysm formation and rupture.

Endothelial cell dysfunction, vascular remodeling, and inflammation

Vascular remodeling is a complex process that is driven, in large part, by hemodynamic stresses along vessel walls.[18] The propensity for aneurysms to form at vessel branch points and the association of aneurysms with environmental stimuli known to disrupt vascular integrity (smoking, hypertension) highlights the role of abnormal blood flow and shear stress in aneurysm formation.[17] High wall shear stress has been shown to initiate activation of the inflammatory response.[20-22] Central to this process is the endothelial cells, which act as an interface between blood flow and the vessel wall.[17] Through the process of mechanotransduction, these cells respond to the mechanical stimuli of shear, stretch, and flow by altering their physical structure and initiating biologic signaling.[23-25] Multiple mechanical sensors have been identified, including, ion channels, integrins, cell adhesion molecules, and G protein-coupled receptors at the apical and basal surfaces of endothelial cells.[26-29] Activation of these sensors initiates intracellular cascades that result in a sustained inflammatory response [Table 1].

Table 1

Summary of the role of inflammatory mediators in aneurysm growth and rupture

Endothelial cellsVascular smooth muscle cellsMacrophages
Role in vascular homeostasisInterface between blood flow and vessel wallContractile state maintains vessel wall integrityM2 subset is antiinflammatory
MechanotransductionPlay a role in vascular repair
Respond to shear stress and flow
Associated cell signaling molecules (secreted by or possess receptors for)VEGFNF-κBMMPs
bFBFMMPsElastases
SDF-1IL-1βMCP-1
COX-2MCP-1NF-κB
PGE2TNF-αTNF-α
NF-κB
TNF-α
Role in aneurysm formationMigration grants inflammatory cells increased access to vessel wallSecretory phenotype leads to inflammatory statePropagate inflammatory response
Proliferation/neovascularization leads to increased inflammatory cell access to vessel wallErratic migrationMay play significant role in progression to rupture
Complex cell signaling leading to chronic inflammatory stateApoptosisRelease elastases and MMPs involved in vessel wall degradation
Weakening of vessel wall

Aoki et al.[30] demonstrated a direct link between shear stress and activation of the inflammatory cascade in a rat model. Cyclooxygenase-2 (COX-2) activity, which is induced by hemodynamic force, generates prostaglandin E2 (PGE2), leading to activation of the pro-inflammatory mediator nuclear factor-κB (NF-κB).[30] Inflammation is then sustained through a positive feedback loop containing PGE2 and NF-κB, creating an environment in which vessel wall degradation can occur. Inhibition of COX-2 or loss of the PGE2 Rp suppressed NF-κB-mediated chronic inflammation and was associated with a decreased incidence of cerebral aneurysms. Additional support for the role of NF-κB-mediated pathways in inflammation-induced aneurysm formation is found in diminished aneurysm formation and growth in rats treated with statins, which inhibit NF-κB activity.[31-33]

Evidence also exists that implicates angiogenesis secondary to inflammation-driven endothelial cell dysfunction as a significant contributor to aneurysm formation. The presence of angiogenic growth factors, including, vascular endothelial growth factor (VEGF) and basic fibroblast growth factor, within aneurysm walls has been well documented within the literature.[34] In pathologic states VEGF induces changes in the endothelium leading to increased permeability at intercellular junctions and activation of pathways resulting in the breakdown of the tunica media and extracellular matrix.[35,36] Importantly, VEGF may also initiate the genesis of new capillary tubes, microvascular sprouting, and maturation of proliferating vessels.[37-39] bFGF targets endothelial cells, fibrocytes, and myocytes and mediates vascular wall maturation during angiogenesis.[40-43] Hoh et al.[44] recently reported on the expression of stromal cell-derived factor-1 (SDF-1), a chemokine with pro-angiogenic and pro-inflammatory properties, in the walls of human and murine intracranial aneurysms. SDF-1 promoted endothelial cell migration and proliferation, as well as capillary tube formation in in vitro studies.

Vessel proliferation within aneurysm walls is a proposed mechanism by which inflammatory cells gain increased access to the underlying tunica media, thereby accelerating degradation of this layer. The vaso vasorum is typically not present in the intracranial vasculature, with the exception of the proximal intracranial carotid and vertebral arteries.[45] However, multiple case reports have described the presence of an extensive vaso vasorum with the walls of intracranial aneurysms.[44,46,47] Neovascularization in these cases was also associated with inflammatory cell invasion on histopathologic examination.

Inflammation-driven degradation of vascular smooth muscle and the extracellular matrix

Under normal physiologic conditions VSMCs, the primary cellular component of the tunica media, remain in a contractile state, maintaining the integrity of the vessel wall. In pathologic conditions, such as those that arise in the setting of increased hemodynamic stress, endothelial dysfunction, as well as direct VSMC injury, leads to disruption of the tunica media and extracellular matrix.[48,49] Central to this process is the VSMC transition from a contractile to a secretory phenotype defined by a loss of markers of contractility and expression of pro-inflammatory cytokines and matrix metalloproteinases (MMPs).[17,50-54] Alterations in VSMC phenotype have been reported in the setting of atherosclerotic lesions, in which these cells upregulate the production of NF-κB, secrete cytokines, release MMPs, and migrate into the intima, where proliferation results in vessel stenosis.[55,56] Interestingly, in atherosclerotic lesions where inflammation leads to VSMC migration and proliferation, intracranial aneurysm walls are defined by VSMC erratic migration and apoptosis.[48] As aneurysm formation progresses, substantial thinning of the tunica media and cellular loss is observed.[48,57,58] Ruptured aneurysms are more frequently found to have hypocellular and hyalinized walls when compared to unruptured aneurysms, highlighting the progressive nature of wall destruction.[59,60]

Multiple inflammatory cascades appear to be involved in VSMC dysfunction and death [Table 1]. Interleukin-1β (IL-1β) is a pro-inflammatory cytokine that initiates a number of deleterious effects within the VSMCs and extracellular matrix.[48] IL-1β plays an important role in recruiting inflammatory cells to atherosclerotic lesions and areas of vessel injury.[61,62] This cytokine also activates NF-κB, which is responsible for inducing inflammatory cascades, promoting pro-inflammatory gene expression, and mediating downregulation of procollagen synthesis within the tunica media.[48,63] IL-1β also directly induces apoptosis of VSMCs, thereby promoting thinning of the aneurysm wall.[48] Further supporting the role of IL-1β in VSMC degradation is the impairment of aneurysm growth in IL-1β deficient mice.[64]

Ets-1 is a transcription factor primarily activated in VSMCs residing in the tunica media.[65] Multiple studies have shown Ets-1 to play a role in the regulation of vascular inflammation, pathologic remodeling, and angiogenesis.[66-69] Aoki et al.[65] demonstrated upregulation and activation of Ets-1 in intracerebral aneurysm VSMCs, strongly implicating this inflammatory transcription factor in aneurysm evolution.

Vascular smooth muscle cells in a secretory state generate increased monocyte chemoattractant protein-1 (MCP-1), a chemokine involved in the attraction and migration of monocytes and macrophages to areas of damaged tunica media.[70] MCP-1 upregulation has been demonstrated in the early stages of cerebral aneurysm formation in rat models, and MCP-1 blockade resulted in decreased macrophage infiltration and aneurysm progression.[71,72] Chalouhi et al.[73] reported high levels of MCP-1 in the lumen of unruptured cerebral aneuryms, implicating this chemokine in early aneurysm formation.

Normal maintenance of the extracellular matrix is dependent on the balance between MMPs and their inhibitors, tissue inhibitors of matrix metalloproteinases (TIMPs).[52] Imbalance in this system results in excessive breakdown of the collagen and elastin of the extracellular matrix and resultant vessel wall weakening, a key contributor to aneurysm development and rupture. MMPs and TIMPs have been linked to the development of atherosclerotic lesions and the genesis of abdominal aortic aneurysms.[12,74-80] Immunohistochemistry and western blotting have demonstrated the presence of MMPs within the walls of intracranial aneurysms.[81,82] In SAH patients, elevated serum MMP-9 levels have been documented, with normalization occurring by postbleed day 12.[83] Cigarette smoke, a well-established, inflammatory stimulus for aneurysm initiation and growth, has been demonstrated to induce macrophage differentiation and increased release of MMP-2/9.[84] In a comparison of smokers and nonsmokers, the carotids of smokers demonstrated elevated levels of MMPs and a decreased concentration of TIMPs and elastin compared to nonsmokers.[85]

Laboratory investigations also support derangement of MMP and TIMP interactions as a key component in the aneurysm pathogenesis. In a rat model, Aoki et al.[51] demonstrated increased levels of MMP-2 and MMP-9 in aneurysm walls, with increasing expression as aneurysms progressed. Ali et al.[86] reported MMP stimulation by cigarette smoke extract in rat cerebral VSMCs in vitro and in carotid VSCMs in vivo.

TUMOR NECROSIS FACTOR-α

Tumor necrosis factor-α has emerged as a potential key contributor in the generation, growth, and eventual rupture of intracranial aneurysms. Multiple studies have demonstrated increased levels of TNF-α within the cerebral circulation in response to injury or ischemia.[87,88] Elevated levels of TNF-α mRNA have been demonstrated in intracranial aneurysms with reverse transcription-polymerase chain reaction.[89] Plasma TNF-α is also increased in aneurysm patients while additional authors have reported an association between single-nucleotide polymorphisms in the TNF-α gene and an increased risk of aneurysm incidence and rupture.[90-94]

Tumor necrosis factor-α is a well-established mediator of inflammation and apoptosis, working through multiple receptors to modulate various cellular responses to injury. Tumor necrosis factor-α receptor 1 (TNFR1) primarily binds the soluble form of TNF-α and plays a central role in TNF-α-induced cellular signaling.[95,96] Importantly, TNFR1 contains a death domain that plays a role in TNF-α-mediated apoptosis.[97,98] A second receptor, TNFR2, is activated by membrane-bound TNF-α and primarily found on endothelial and immunomodulatory cells. Activation of TNFR1 signals apoptosis through the activation of the Fas-associated death domain (FADD), which in turn, binds and activates pro-caspase 8. Ultimately, this pathway leads to the activation of multiple proteases that result in apoptosis. Jayaraman et al.[2] postulate that a complex interaction between TNF-α and both receptors induces and promotes aneurysm growth.Initially, endothelial activation is driven by TNFR2 and membrane-bound TNF-α. Activated endothelial cells generate high concentrations of soluble TNF-α, which interacts with the TNFR1, leading to endothelial cell dysfunction and death. Endothelial cell death, in turn, creates increased vascular permeability, thereby creating multiple pathways for macrophage infiltration, MMP generation, and loss of the VSMC layer. An additional receptor, tTNF-Rp55, initiates apoptosis through the induction of caspases.[89] The TNF-Rp55 also induces the recruitment and activation of caspases through interaction with the FADD protein.[99] Documented increased expression of the FADD protein in human aneurysms also supports TNF-α-mediated apoptosis as a causative factor in aneurysm formation.[89]

Tumor necrosis factor-α also appears to play an important role in the pathologic VSMC changes observed in aneurysm formation and rupture, driving these cells to change from a contractile to a synthetic and pro-inflammatory phenotype. Furthermore, the secretory phenotype of VSMCs participates in the secretion of TNF-α when exposed to inflammatory stimuli.[100,101] Ali et al.[102] demonstrated TNF-α to suppress expression of contractile genes and induce the expression of pro-inflammatory genes in VSMCs in vivo and in vitro. In the same study, TNF-α induced increased expression of Kruppel-like transcription factor 4, a known regulator of VSMC differentiation.[102] Additional studies have identified cigarette smoke, a known potent pro-inflammatory stimulus, to induce similar phenotypic changes in VSMCs, including increased secretion of TNF-α.[100] Clearly the data shows TNF-α secretion to be involved in a positive feedback loop with VSMCs, with TNF-α stimulating phenotypic change and thus driving its own secretion from these cells. The fact that VSMCs appear to first exhibit pathologic migration, followed by disappearance prior to rupture suggests TNF-α first induces phenotypic change that is followed by apoptosis in the presence of chronically sustained levels of the cytokine.[103]

Inflammatory cell migration and infiltration

Endothelial cell dysfunction and apoptosis increase the permeability of vessel walls, allowing for enhanced binding and transmigration of inflammatory cells into the underlying VSMC layer. T-cells have been demonstrated within aneurysm walls where they respond to antigen presentation by monocytes and macrophages.[104] In a rat model, mast cells were significantly increased within the walls of forming cerebral aneurysms.[105] Inhibition of mast cell degranulation diminished aneurysm size, prevented thinning of the tunica media, blocked NF-κB activation, and decreased expression of MCP-1, MMP, and IL-1β. An evaluation of human intracerebral aneurysms found mast cells to be significantly increased in ruptured compared to unruptured aneurysms.[106] These findings suggest mast cells play an important role in aneurysm growth and rupture.

Monocytes and macrophages appear to be essential to aneurysm formation and rupture, a finding that has been repeatedly demonstrated in animal and clinical studies. These cells secrete MMPs and elastases which are responsible for degradation of the extracellular matrix and the internal elastic lamina.[107] Kanematsu et al.[108] observed macrophage depletion and inhibition of MCP-1 to be associated with a reduced incidence of intracranial aneurysms. Aoki et al.[72] also observed a significant decrease in cerebral aneurysm formation, macrophage accumulation, and expression of MMP-2 and MMP-9 in MCP-1 deficient mice. In a murine model, Ruzevick et al.[109] observed the haptoglobin 2-2 (Hp2-2) genotype, which is linked to a pro-inflammatory state, to be associated with significantly larger aneurysms and a greater number of macrophages within the aneurysm walls. Histopathological examination of ruptured and unruptured human intracranial aneurysms has demonstrated macrophage infiltration within aneurysmal walls.[59,60,110] Chalouhi et al.[73] found high plasma concentrations of MCP-1 within the lumens of intracranial aneurysms, suggesting active recruitment of macrophages and other inflammatory cells.

There is mounting evidence that macrophage invasion may be a causal factor in rupture. Frosen et al.[60] found more prominent macrophage infiltration in ruptured aneurysms compared to that found in unruptured aneurysms. Of particular interest was the infiltration of macrophages observed within the first 12 h after rupture, suggesting that macrophages may induce the rupture. MRI investigations of aneurysms with ferumoxytol (AMAG Pharmaceuticals, Lexington, Massachusetts, USA), a superparamagnetic iron oxide particle cleared by macrophages, have also linked macrophage infiltration with rupture.[111-113] In a study of 48 unruptured aneurysms, all aneurysms that showed early uptake of ferumoxytol on MRI ruptured within 6 months.[114] Among aneurysms demonstrating late uptake, there were no ruptures or increase in size during the follow-up period. Immunohistochemical analysis found greater levels of inflammation in aneurysms with early uptake of ferumoxytol. The authors propose that early uptake of ferumoxytol is associated with more prominent inflammation, increased macrophage infiltration, and thus, a greater risk of rupture. Follow-up investigations utilizing aspirin as an anti-inflammatory agent found a decrease in aneurysm wall signal intensity on ferumoxytol MRI and a diminished number of macrophages on immunostaining.[111] Hasan et al.[106] examined the M1 (pro-inflammatory) and M2 (anti-inflammatory) subsets of macrophages in a population of ruptured and unruptured clipped aneurysms. While M1 and M2 macrophages were observed in equal proportions in unruptured aneurysms, M1 macrophages were found in a significantly greater proportion in ruptured aneurysms.

Therapeutic implications

Current treatment options for ruptured and unruptured cerebral aneurysms include microsurgical and endovascular obliteration. These interventions are associated with a significant morbidity and mortality, the risk that is magnified in cases of asymptomatic unruptured aneurysms. Thus, study of the mechanisms underlying aneurysm evolution is critical to establish a better understanding of which aneurysms are more likely to rupture. Additionally, this data may provide insight into the biological pathways best-suited for pharmacological intervention, stabilization of the aneurysm wall, and prevention of rupture. The inflammatory process appears to possess multiple targets ideally suited for such pharmacological treatment.

The wide range of pro-inflammatory pathways induced by TNF-α makes this cytokine an attractive target for blockade.[2] Preservation of endothelial function, maintenance of the VSMC contractile phenotype, and inhibition of inflammatory cell migration are all potential benefits of TNF-α antagonism. MMP-2 and MMP-9 blockade has proven to limit aneurysm development in animal models, possibly identifying an additional mechanism with clinical relevance.[51] Modification of VSMC phenotype continues to be an area of interest, with both cytokine and transcription factor-mediated changes representing possible therapeutic targets.[102] Macrophages represent an additional cellular target for pharmacologic intervention. Hasan et al.[113] has already shown aspirin to potentially decrease the risk of aneurysm rupture, a clinical benefit possibly derived from diminished macrophage infiltration of the aneurysm wall. Nakakoa et al.[115] has demonstrated gene expression profiles in ruptured aneurysms that indicate the macrophage-mediated inflammation as a key contributor to aneurysm rupture. Future studies may determine these identified genes as diagnostic markers for aneurysms prone to rupture or potential therapeutic targets. Importantly, in all of these areas the therapeutic benefit must be weighed against the risks of altering what, in many instances, is a natural and protective biological response.[2]

Conclusion

Cerebral aneurysms remain a significant health and socioeconomic problem, with the majority of ruptures leading to death or severe disability. Microsurgical and endovascular interventions represent the two modalities of treatment; however, both are associated with risk. An understanding of the mechanisms leading to aneurysm development is necessary to identify markers predictive of growth and rupture, as well as targets for pharmacologic intervention. Through multiple pathways, the inflammatory response appears to drive the initiation of aneurysm formation, potentiate continued growth, and contribute to eventual rupture. Animal and human studies highlight the complex interaction between the endothelium, VSMCs, and inflammatory cells. Phenotypic change, cytokine production, apoptosis, and cellular migration all partake in aneurysm evolution. Delineation of those at greatest risk for rupture would afford improved patient selection with regards to aneurysm obliteration. Furthermore, the definition of these various processes may hold promising pharmacological therapeutic implications in the future.

Financial support and sponsorship

Nil.

Conflicts of interest

There are no conflicts of interest.

REFERENCES

1. Schievink WI. Intracranial aneurysms. N Engl J Med 1997;336:28-40.

2. Jayaraman T, Paget A, Shin YS, Li X, Mayer J, Chaudhry H, Niimi Y, Silane M, Berenstein A. TNF-alpha-mediated inflammation in cerebral aneurysms: a potential link to growth and rupture. Vasc Health Risk Manag 2008;4:805-17.

3. Stegmayr B, Eriksson M, Asplund K. Declining mortality from subarachnoid hemorrhage: changes in incidence and case fatality from 1985 through 2000. Stroke 2004;35:2059-63.

4. Nieuwkamp DJ, Setz LE, Algra A, Linn FH, de Rooij NK, Rinkel GJ. Changes in case fatality of aneurysmal subarachnoid haemorrhage over time, according to age, sex, and region: a meta-analysis. Lancet Neurol 2009;8:635-42.

5. Frösen J, Tulamo R, Paetau A, Laaksamo E, Korja M, Laakso A, Niemela M, Hernesniemi J. Saccular intracranial aneurysm: pathology and mechanisms. Acta Neuropathol 2012;123:773-86.

6. Ronkainen A, Hernesniemi J, Ryynänen M. Familial subarachnoid hemorrhage in east Finland, 1977-1990. Neurosurgery 1993;33:787-96.

7. Schievink WI, Schaid DJ, Michels VV, Piepgras DG. Familial aneurysmal subarachnoid hemorrhage: a community-based study. J Neurosurg 1995;83:426-9.

8. Juvela S, Porras M, Poussa K. Natural history of unruptured intracranial aneurysms: probability of and risk factors for aneurysm rupture. J Neurosurg 2000;93:379-87.

9. Yasui N, Suzuki A, Nishimura H, Suzuki K, Abe T. Long-term follow-up study of unruptured intracranial aneurysms. Neurosurgery 1997;40:1155-9.

10. Isaksen J, Egge A, Waterloo K, Romner B, Ingebrigtsen T. Risk factors for aneurysmal subarachnoid haemorrhage: the Tromsø study. J Neurol Neurosurg Psychiatry 2002;73:185-7.

11. Sandvei MS, Romundstad PR, Müller TB, Vatten L, Vik A. Risk factors for aneurysmal subarachnoid hemorrhage in a prospective population study: the HUNT study in Norway. Stroke 2009;40:1958-62.

12. Lijnen HR. Metalloproteinases in development and progression of vascular disease. Pathophysiol Haemost Thromb 2003;33:275-81.

13. Juvela S, Hillbom M, Numminen H, Koskinen P. Cigarette smoking and alcohol consumption as risk factors for aneurysmal subarachnoid hemorrhage. Stroke 1993;24:639-46.

14. Hashimoto N, Handa H, Hazama F. Experimentally induced cerebral aneurysms in rats. Surg Neurol 1978;10:3-8.

15. Hashimoto N, Kim C, Kikuchi H, Kojima M, Kang Y, Hazama F. Experimental induction of cerebral aneurysms in monkeys. J Neurosurg 1987;67:903-5.

16. Morimoto M, Miyamoto S, Mizoguchi A, Kume N, Kita T, Hashimoto N. Mouse model of cerebral aneurysm: experimental induction by renal hypertension and local hemodynamic changes. Stroke 2002;33:1911-5.

17. Turjman AS, Turjman F, Edelman ER. Role of fluid dynamics and inflammation in intracranial aneurysm formation. Circulation 2014;129:373-82.

18. Penn DL, Komotar RJ, Sander Connolly E. Hemodynamic mechanisms underlying cerebral aneurysm pathogenesis. J Clin Neurosci 2011;18:1435-8.

19. Chalouhi N, Ali MS, Starke RM, Jabbour PM, Tjoumakaris SI, Gonzalez LF, Rosenwasser RH, Koch WJ, Dumont AS. Cigarette smoke and inflammation: role in cerebral aneurysm formation and rupture. Mediators Inflamm 2012;2012:271582.

20. Chiu JJ, Wang DL, Chien S, Skalak R, Usami S. Effects of disturbed flow on endothelial cells. J Biomech Eng 1998;120:2-8.

21. Nagel T, Resnick N, Dewey CF Jr, Gimbrone MA, Jr. Vascular endothelial cells respond to spatial gradients in fluid shear stress by enhanced activation of transcription factors. Arterioscler Thromb Vasc Biol 1999;19:1825-34.

22. Tardy Y, Resnick N, Nagel T, Gimbrone MA Jr, Dewey CF, Jr. Shear stress gradients remodel endothelial monolayers in vitro via a cell proliferation-migration-loss cycle. Arterioscler Thromb Vasc Biol 1997;17:3102-6.

23. Burridge K, Chrzanowska-Wodnicka M. Focal adhesions, contractility, and signaling. Annu Rev Cell Dev Biol 1996;12:463-518.

24. Wang N, Tytell JD, Ingber DE. Mechanotransduction at a distance: mechanically coupling the extracellular matrix with the nucleus. Nat Rev Mol Cell Biol 2009;10:75-82.

25. Takeichi M. The cadherins: cell-cell adhesion molecules controlling animal morphogenesis. Development 1988;102:639-55.

26. Olesen SP, Clapham DE, Davies PF. Haemodynamic shear stress activates a K+ current in vascular endothelial cells. Nature 1988;331:168-70.

27. Tzima E, del Pozo MA, Shattil SJ, Chien S, Schwartz MA. Activation of integrins in endothelial cells by fluid shear stress mediates Rho-dependent cytoskeletal alignment. EMBO J 2001;20:4639-47.

28. Chachisvilis M, Zhang YL, Frangos JA. G protein-coupled receptors sense fluid shear stress in endothelial cells. Proc Natl Acad Sci U S A 2006;103:15463-8.

29. Tzima E, Irani-Tehrani M, Kiosses WB, Dejana E, Schultz DA, Engelhardt B, Cao G, DeLisser H, Schwartz MA. A mechanosensory complex that mediates the endothelial cell response to fluid shear stress. Nature 2005;437:426-31.

30. Aoki T, Nishimura M, Matsuoka T, Yamamoto K, Furuyashiki T, Kataoka H, Kitaoka S, Ishibashi R, Ishibazawa A, Miyamoto S, Morishita R, Ando J, Hashimoto N, Nozaki K, Narumiya S. PGE(2) -EP(2) signalling in endothelium is activated by haemodynamic stress and induces cerebral aneurysm through an amplifying loop via NF-κB. Br J Pharmacol 2011;163:1237-49.

31. Liao JK, Laufs U. Pleiotropic effects of statins. Annu Rev Pharmacol Toxicol 2005;45:89-118.

32. Aoki T, Kataoka H, Ishibashi R, Nozaki K, Hashimoto N. Simvastatin suppresses the progression of experimentally induced cerebral aneurysms in rats. Stroke 2008;39:1276-85.

33. Aoki T, Kataoka H, Ishibashi R, Nakagami H, Nozaki K, Morishita R, Hashimoto N. Pitavastatin suppresses formation and progression of cerebral aneurysms through inhibition of the nuclear factor kappaB pathway. Neurosurgery 2009;64:357-65.

34. Kilic T, Sohrabifar M, Kurtkaya O, Yildirim O, Elmaci I, Günel M, Pamir MN. Expression of structural proteins and angiogenic factors in normal arterial and unruptured and ruptured aneurysm walls. Neurosurgery 2005;57:997-1007.

35. Ray JM, Stetler-Stevenson WG. The role of matrix metalloproteases and their inhibitors in tumour invasion, metastasis and angiogenesis. Eur Respir J 1994;7:2062-72.

36. Unemori EN, Ferrara N, Bauer EA, Amento EP. Vascular endothelial growth factor induces interstitial collagenase expression in human endothelial cells. J Cell Physiol 1992;153:557-62.

37. Dvorak HF, Brown LF, Detmar M, Dvorak AM. Vascular permeability factor/vascular endothelial growth factor, microvascular hyperpermeability, and angiogenesis. Am J Pathol 1995;146:1029-39.

38. Ferrara N, Houck KA, Jakeman LB, Winer J, Leung DW. The vascular endothelial growth factor family of polypeptides. J Cell Biochem 1991;47:211-8.

39. Fong GH, Rossant J, Gertsenstein M, Breitman ML. Role of the Flt-1 receptor tyrosine kinase in regulating the assembly of vascular endothelium. Nature 1995;376:66-70.

40. Basilico C, Moscatelli D. The FGF family of growth factors and oncogenes. Adv Cancer Res 1992;59:115-65.

41. Bobik A, Campbell JH. Vascular derived growth factors: cell biology, pathophysiology, and pharmacology. Pharmacol Rev 1993;45:1-42.

42. Partanen J, Vainikka S, Alitalo K. Structural and functional specificity of FGF receptors. Philos Trans R Soc Lond B Biol Sci 1993;340:297-303.

43. Pepper MS, Ferrara N, Orci L, Montesano R. Potent synergism between vascular endothelial growth factor and basic fibroblast growth factor in the induction of angiogenesis in vitro. Biochem Biophys Res Commun 1992;189:824-31.

44. Hoh BL, Hosaka K, Downes DP, Nowicki KW, Wilmer EN, Velat GJ, Scott EW. Stromal cell-derived factor-1 promoted angiogenesis and inflammatory cell infiltration in aneurysm walls. J Neurosurg 2014;120:73-86.

45. Aydin F. Do human intracranial arteries lack vasa vasorum? A comparative immunohistochemical study of intracranial and systemic arteries. Acta Neuropathol 1998;96:22-8.

46. Numagami Y, Ezura M, Takahashi A, Yoshimoto T. Antegrade recanalization of completely embolized internal carotid artery after treatment of a giant intracavernous aneurysm: a case report. Surg Neurol 1999;52:611-6.

47. Iihara K, Murao K, Sakai N, Soeda A, Ishibashi-Ueda H, Yutani C, Yamada N, Nagata I. Continued growth of and increased symptoms from a thrombosed giant aneurysm of the vertebral artery after complete endovascular occlusion and trapping: the role of vasa vasorum. Case report. J Neurosurg 2003;98:407-13.

48. Aoki T, Kataoka H, Ishibashi R, Nozaki K, Morishita R, Hashimoto N. Reduced collagen biosynthesis is the hallmark of cerebral aneurysm: contribution of interleukin-1beta and nuclear factor-kappaB. Arterioscler Thromb Vasc Biol 2009;29:1080-6.

49. Chiu JJ, Chen LJ, Chang SF, Lee PL, Lee CI, Tsai MC, Lee DY, Hsieh HP, Usami S, Chien S. Shear stress inhibits smooth muscle cell-induced inflammatory gene expression in endothelial cells: role of NF-kappaB. Arterioscler Thromb Vasc Biol 2005;25:963-9.

50. Nakajima N, Nagahiro S, Sano T, Satomi J, Satoh K. Phenotypic modulation of smooth muscle cells in human cerebral aneurysmal walls. Acta Neuropathol 2000;100:475-80.

51. Aoki T, Kataoka H, Morimoto M, Nozaki K, Hashimoto N. Macrophage-derived matrix metalloproteinase-2 and -9 promote the progression of cerebral aneurysms in rats. Stroke 2007;38:162-9.

52. Aoki T, Kataoka H, Moriwaki T, Nozaki K, Hashimoto N. Role of TIMP-1 and TIMP-2 in the progression of cerebral aneurysms. Stroke 2007;38:2337-45.

53. Kolega J, Gao L, Mandelbaum M, Mocco J, Siddiqui AH, Natarajan SK, Meng H. Cellular and molecular responses of the basilar terminus to hemodynamics during intracranial aneurysm initiation in a rabbit model. J Vasc Res 2011;48:429-42.

54. Owens GK, Kumar MS, Wamhoff BR. Molecular regulation of vascular smooth muscle cell differentiation in development and disease. Physiol Rev 2004;84:767-801.

55. Galis ZS, Muszynski M, Sukhova GK, Simon-Morrissey E, Unemori EN, Lark MW, Amento E, Libby P. Cytokine-stimulated human vascular smooth muscle cells synthesize a complement of enzymes required for extracellular matrix digestion. Circ Res 1994;75:181-9.

56. Bourcier T, Sukhova G, Libby P. The nuclear factor kappa-B signaling pathway participates in dysregulation of vascular smooth muscle cells in vitro and in human atherosclerosis. J Biol Chem 1997;272:15817-24.

57. Meng H, Metaxa E, Gao L, Liaw N, Natarajan SK, Swartz DD, Siddiqui AH, Kolega J, Mocco J. Progressive aneurysm development following hemodynamic insult. J Neurosurg 2011;114:1095-103.

58. Hazama F, Hashimoto N. An animal model of cerebral aneurysms. Neuropathol Appl Neurobiol 1987;13:77-90.

59. Kataoka K, Taneda M, Asai T, Kinoshita A, Ito M, Kuroda R. Structural fragility and inflammatory response of ruptured cerebral aneurysms. A comparative study between ruptured and unruptured cerebral aneurysms. Stroke 1999;30:1396-401.

60. Frösen J, Piippo A, Paetau A, Kangasniemi M, Niemelä M, Hernesniemi J, Jääskeläinen J. Remodeling of saccular cerebral artery aneurysm wall is associated with rupture: histological analysis of 24 unruptured and 42 ruptured cases. Stroke 2004;35:2287-93.

61. Bhaskar V, Yin J, Mirza AM, Phan D, Vanegas S, Issafras H, Michelson K, Hunter JJ, Kantak SS. Monoclonal antibodies targeting IL-1 beta reduce biomarkers of atherosclerosis in vitro and inhibit atherosclerotic plaque formation in Apolipoprotein E-deficient mice. Atherosclerosis 2011;216:313-20.

62. Wang X, Feuerstein GZ, Gu JL, Lysko PG, Yue TL. Interleukin-1 beta induces expression of adhesion molecules in human vascular smooth muscle cells and enhances adhesion of leukocytes to smooth muscle cells. Atherosclerosis 1995;115:89-98.

63. Alexander MR, Murgai M, Moehle CW, Owens GK. Interleukin-1ß modulates smooth muscle cell phenotype to a distinct inflammatory state relative to PDGF-DD via NF-κB-dependent mechanisms. Physiol Genomics 2012;44:417-29.

64. Moriwaki T, Takagi Y, Sadamasa N, Aoki T, Nozaki K, Hashimoto N. Impaired progression of cerebral aneurysms in interleukin-1beta-deficient mice. Stroke 2006;37:900-5.

65. Aoki T, Kataoka H, Nishimura M, Ishibashi R, Morishita R, Miyamoto S. Ets-1 promotes the progression of cerebral aneurysm by inducing the expression of MCP-1 in vascular smooth muscle cells. Gene Ther 2010;17:1117-23.

66. Hashiya N, Jo N, Aoki M, Matsumoto K, Nakamura T, Sato Y, Ogata N, Ogihara T, Kaneda Y, Morishita R. In vivo evidence of angiogenesis induced by transcription factor Ets-1: ets-1 is located upstream of angiogenesis cascade. Circulation 2004;109:3035-41.

67. Kavurma MM, Bobryshev Y, Khachigian LM. Ets-1 positively regulates Fas ligand transcription via cooperative interactions with Sp1. J Biol Chem 2002;277:36244-52.

68. Hultgårdh-Nilsson A, Cercek B, Wang JW, Naito S, Lövdahl C, Sharifi B, Forrester JS, Fagin JA. Regulated expression of the ets-1 transcription factor in vascular smooth muscle cells in vivo and in vitro. Circ Res 1996;78:589-95.

69. Zhan Y, Brown C, Maynard E, Anshelevich A, Ni W, Ho IC, Oettgen P. Ets-1 is a critical regulator of Ang II-mediated vascular inflammation and remodeling. J Clin Invest 2005;115:2508-16.

70. Charo IF, Ransohoff RM. The many roles of chemokines and chemokine receptors in inflammation. N Engl J Med 2006;354:610-21.

71. Mandelbaum M, Kolega J, Dolan JM, Siddiqui AH, Meng H. A critical role for proinflammatory behavior of smooth muscle cells in hemodynamic initiation of intracranial aneurysm. PLoS One 2013;8:e74357.

72. Aoki T, Kataoka H, Ishibashi R, Nozaki K, Egashira K, Hashimoto N. Impact of monocyte chemoattractant protein-1 deficiency on cerebral aneurysm formation. Stroke 2009;40:942-51.

73. Chalouhi N, Points L, Pierce GL, Ballas Z, Jabbour P, Hasan D. Localized increase of chemokines in the lumen of human cerebral aneurysms. Stroke 2013;44:2594-7.

74. Newby AC. Dual role of matrix metalloproteinases (matrixins) in intimal thickening and atherosclerotic plaque rupture. Physiol Rev 2005;85:1-31.

75. Galis ZS, Khatri JJ. Matrix metalloproteinases in vascular remodeling and atherogenesis: the good, the bad, and the ugly. Circ Res 2002;90:251-62.

76. Allaire E, Forough R, Clowes M, Starcher B, Clowes AW. Local overexpression of TIMP-1 prevents aortic aneurysm degeneration and rupture in a rat model. J Clin Invest 1998;102:1413-20.

77. Eskandari MK, Vijungco JD, Flores A, Borensztajn J, Shively V, Pearce WH. Enhanced abdominal aortic aneurysm in TIMP-1-deficient mice. J Surg Res 2005;123:289-93.

78. Ikonomidis JS, Gibson WC, Butler JE, McClister DM, Sweterlitsch SE, Thompson RP, Mukherjee R, Spinale FG. Effects of deletion of the tissue inhibitor of matrix metalloproteinases-1 gene on the progression of murine thoracic aortic aneurysms. Circulation 2004;110:II268-73.

79. Koullias GJ, Ravichandran P, Korkolis DP, Rimm DL, Elefteriades JA. Increased tissue microarray matrix metalloproteinase expression favors proteolysis in thoracic aortic aneurysms and dissections. Ann Thorac Surg 2004;78:2106-10.

80. Ogata T, Shibamura H, Tromp G, Sinha M, Goddard KA, Sakalihasan N, Limet R, MacKean GL, Arthur C, Sueda T, Land S, Kuivaniemi H. Genetic analysis of polymorphisms in biologically relevant candidate genes in patients with abdominal aortic aneurysms. J Vasc Surg 2005;41:1036-42.

81. Bruno G, Todor R, Lewis I, Chyatte D. Vascular extracellular matrix remodeling in cerebral aneurysms. J Neurosurg 1998;89:431-40.

82. Kim SC, Singh M, Huang J, Prestigiacomo CJ, Winfree CJ, Solomon RA, Connolly ES, Jr. Matrix metalloproteinase-9 in cerebral aneurysms. Neurosurgery 1997;41:642-66.

83. Horstmann S, Su Y, Koziol J, Meyding-Lamadé U, Nagel S, Wagner S. MMP-2 and MMP-9 levels in peripheral blood after subarachnoid hemorrhage. J Neurol Sci 2006;251:82-6.

84. Hossain M, Sathe T, Fazio V, Mazzone P, Weksler B, Janigro D, Rapp E, Cucullo L. Tobacco smoke: a critical etiological factor for vascular impairment at the blood-brain barrier. Brain Res 2009;1287:192-205.

85. Kangavari S, Matetzky S, Shah PK, Yano J, Chyu KY, Fishbein MC, Cercek B. Smoking increases inflammation and metalloproteinase expression in human carotid atherosclerotic plaques. J Cardiovasc Pharmacol Ther 2004;9:291-8.

86. Ali A, Yu Z, Jabbour PM, Tjoumakaris SI, Gonzalez LF, Rosenwasser RH, Koch WJ, Dumont AS. Cigarette smoke directly produces a pro-inflammatory/matrix remodeling phenotype in cerebral vascular smooth muscle cells via KLF4: a potential mechanism for cerebral aneurysm pathogenesis. Stroke 2012;43:A112.

87. Edvinsson LI, Povlsen GK. Vascular plasticity in cerebrovascular disorders. J Cereb Blood Flow Metab 2011;31:1554-71.

88. Maddahi A, Kruse LS, Chen QW, Edvinsson L. The role of tumor necrosis factor-α and TNF-α receptors in cerebral arteries following cerebral ischemia in rat. J Neuroinflammation 2011;8:107.

89. Jayaraman T, Berenstein V, Li X, Mayer J, Silane M, Shin YS, Niimi Y, Kiliç T, Gunel M, Berenstein A. Tumor necrosis factor alpha is a key modulator of inflammation in cerebral aneurysms. Neurosurgery 2005;57:558-64.

90. Zhang HF, Zhao MG, Liang GB, Song ZQ, Li ZQ. Expression of pro-inflammatory cytokines and the risk of intracranial aneurysm. Inflammation 2013;36:1195-200.

91. Low SK, Zembutsu H, Takahashi A, Kamatani N, Cha PC, Hosono N, Kubo M, Matsuda K, Nakamura Y. Impact of LIMK1, MMP2 and TNF-α variations for intracranial aneurysm in Japanese population. J Hum Genet 2011;56:211-6. in Japanese

92. Fontanella M, Rainero I, Gallone S, Rubino E, Fenoglio P, Valfrè W, Garbossa D, Carlino C, Ducati A, Pinessi L. Tumor necrosis factor-alpha gene and cerebral aneurysms. Neurosurgery 2007;60:668-72.

93. Wilson AG, Symons JA, McDowell TL, McDevitt HO, Duff GW. Effects of a polymorphism in the human tumor necrosis factor alpha promoter on transcriptional activation. Proc Natl Acad Sci U S A 1997;94:3195-9.

94. Hajeer AH, Hutchinson IV. Influence of TNFalpha gene polymorphisms on TNFalpha production and disease. Hum Immunol 2001;62:1191-9.

95. Grell M, Zimmermann G, Hülser D, Pfizenmaier K, Scheurich P. TNF receptors TR60 and TR80 can mediate apoptosis via induction of distinct signal pathways. J Immunol 1994;153:1963-72.

96. Bazzoni F, Beutler B. The tumor necrosis factor ligand and receptor families. N Engl J Med 1996;334:1717-25.

97. Pimentel-Mui-os FX, Seed B. Regulated commitment of TNF receptor signaling: a molecular switch for death or activation. Immunity 1999;11:783-93.

98. Tang V, Dhirapong A, Yabes AP, Weiss RH. TNF-alpha-mediated apoptosis in vascular smooth muscle cells requires p73. Am J Physiol Cell Physiol 2005;289:C199-206.

99. Stehbens WE. Apoptosis and matrix vesicles in the genesis of arterial aneurysms of cerebral arteries. Stroke 1998;29:1478-80.

100. Starke RM, Ali MS, Jabbour PM, Tjoumakaris SI, Gonzalez F, Hasan DM, Rosenwasser RH, Owens GK, Koch WJ, Dumont AS. Cigarette smoke modulates vascular smooth muscle phenotype: implications for carotid and cerebrovascular disease. PLoS One 2013;8:e71954.

101. Chen QW, Edvinsson L, Xu CB. Cigarette smoke extract promotes human vascular smooth muscle cell proliferation and survival through ERK1/2- and NF-κB-dependent pathways. ScientificWorldJournal 2010;10:2139-56.

102. Ali MS, Starke RM, Jabbour PM, Tjoumakaris SI, Gonzalez LF, Rosenwasser RH, Owens GK, Koch WJ, Greig NH, Dumont AS. TNF-α induces phenotypic modulation in cerebral vascular smooth muscle cells: implications for cerebral aneurysm pathology. J Cereb Blood Flow Metab 2013;33:1564-73.

103. Chalouhi N, Ali MS, Jabbour PM, Tjoumakaris SI, Gonzalez LF, Rosenwasser RH, Koch WJ, Dumont AS. Biology of intracranial aneurysms: role of inflammation. J Cereb Blood Flow Metab 2012;32:1659-76.

104. Krischek B, Kasuya H, Tajima A, Akagawa H, Sasaki T, Yoneyama T, Ujiie H, Kubo O, Bonin M, Takakura K, Hori T, Inoue I. Network-based gene expression analysis of intracranial aneurysm tissue reveals role of antigen presenting cells. Neuroscience 2008;154:1398-407.

105. Ishibashi R, Aoki T, Nishimura M, Hashimoto N, Miyamoto S. Contribution of mast cells to cerebral aneurysm formation. Curr Neurovasc Res 2010;7:113-24.

106. Hasan D, Chalouhi N, Jabbour P, Hashimoto T. Macrophage imbalance (M1 vs. M2) and upregulation of mast cells in wall of ruptured human cerebral aneurysms: preliminary results. J Neuroinflammation 2012;9:222.

107. Hashimoto T, Meng H, Young WL. Intracranial aneurysms: links among inflammation, hemodynamics and vascular remodeling. Neurol Res 2006;28:372-80.

108. Kanematsu Y, Kanematsu M, Kurihara C, Tada Y, Tsou TL, van Rooijen N, Lawton MT, Young WL, Liang EI, Nuki Y, Hashimoto T. Critical roles of macrophages in the formation of intracranial aneurysm. Stroke 2011;42:173-8.

109. Ruzevick J, Jackson C, Pradilla G, Garzon-Muvdi T, Tamargo RJ. Aneurysm formation in proinflammatory, transgenic haptoglobin 2-2 mice. Neurosurgery 2013;72:70-6.

110. Chyatte D, Bruno G, Desai S, Todor DR. Inflammation and intracranial aneurysms. Neurosurgery 1999;45:1137-46.

111. Hasan DM, Chalouhi N, Jabbour P, Dumont AS, Kung DK, Magnotta VA, Young WL, Hashimoto T, Richard Winn H, Heistad D. Evidence that acetylsalicylic acid attenuates inflammation in the walls of human cerebral aneurysms: preliminary results. J Am Heart Assoc 2013;2:e000019.

112. Hasan DM, Chalouhi N, Jabbour P, Magnotta VA, Kung DK, Young WL. Imaging aspirin effect on macrophages in the wall of human cerebral aneurysms using ferumoxytol-enhanced MRI: preliminary results. J Neuroradiol 2013;40:187-91.

113. Hasan DM, Mahaney KB, Brown RD Jr, Meissner I, Piepgras DG, Huston J, Capuano AW, Torner JC; International Study of Unruptured Intracranial Aneurysms Investigators. Aspirin as a promising agent for decreasing incidence of cerebral aneurysm rupture. Stroke 2011;42:3156-62.

114. Hasan D, Chalouhi N, Jabbour P, Dumont AS, Kung DK, Magnotta VA, Young WL, Hashimoto T, Winn HR, Heistad D. Early change in ferumoxytol-enhanced magnetic resonance imaging signal suggests unstable human cerebral aneurysm: a pilot study. Stroke 2012;43:3258-65.

115. Nakaoka H, Tajima A, Yoneyama T, Hosomichi K, Kasuya H, Mizutani T, Inoue I. Gene expression profiling reveals distinct molecular signatures associated with the rupture of intracranial. aneurysm. Stroke 2014;45:2239-45.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Amenta PS, Valle E, Dumont AS, Medel R. Inflammation and intracranial aneurysms: mechanisms of initiation, growth, and rupture. Neurosciences 2015;2:68-76. http://dx.doi.org/10.4103/2347-8659.153975

AMA Style

Amenta PS, Valle E, Dumont AS, Medel R. Inflammation and intracranial aneurysms: mechanisms of initiation, growth, and rupture. Neuroimmunology and Neuroinflammation. 2015; 2: 68-76. http://dx.doi.org/10.4103/2347-8659.153975

Chicago/Turabian Style

Amenta, Peter S., Edison Valle, Aaron S. Dumont, Ricky Medel. 2015. "Inflammation and intracranial aneurysms: mechanisms of initiation, growth, and rupture" Neuroimmunology and Neuroinflammation. 2: 68-76. http://dx.doi.org/10.4103/2347-8659.153975

ACS Style

Amenta, PS.; Valle E.; Dumont AS.; Medel R. Inflammation and intracranial aneurysms: mechanisms of initiation, growth, and rupture. Neurosciences. 2015, 2, 68-76. http://dx.doi.org/10.4103/2347-8659.153975

About This Article

Special Issue

This article belongs to the Special Issue The Role of Inflammation in Cerebral Aneurysms
This is an open access article distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 License (http://creativecommons.org/licenses/by-nc-sa/3.0/), which allows others to remix, tweak and build upon the work non-commercially, as long as the author is credited and the new creations are licensed under the identical terms.

Data & Comments

Data

Views
6505
Downloads
1064
Citations
3
Comments
0
0

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 4 clicks
Like This Article 0 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Neuroimmunology and Neuroinflammation
ISSN 2349-6142 (Online) 2347-8659 (Print)

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/