Introduction

Intracerebral hemorrhage (ICH) is defined as non-traumatic extravasation of blood into the brain parenchyma. ICH is the second most common subtype of stroke, accounting for approximately 10–20% of all types of strokes (Feigin et al. 2009; Sacco et al. 2009). Notably, ICH is associated with very high morbidity and mortality and regarded as one of the deadliest stroke subtypes. Around 5.3 million cases of ICH were reported worldwide in 2010, and the ICH-related fatality rate ranges from 35% at 7 days to 59% at 1 year (Hemphill et al. 2015; Rincon and Mayer 2013; Sacco et al. 2009). In addition, survivors of ICH are often left with lifelong disabilities, further implicating the devastating nature of the disease.

The pathophysiological processes following ICH involve both primary and secondary brain injury. The primary brain injury generally occurs within minutes to hours of initial cerebral bleeding and is caused by the mass effect of hematoma whereas the secondary brain injury, resulting from inflammation, oxidative stress, excitotoxicity, and blood-induced cytotoxicity persists for a longer period of time and is mostly responsible for both acute as well as long-term neurological deficits.

Importantly, there is no efficient therapy for ICH, and the clinical management of ICH is largely restricted to supportive care. Lack of treatment options reflects on the complexity as well as the poorly defined pathophysiology of ICH warranting further investigation. To this end, herein, we employ a quantitative proteomics approach combined with bioinformatics in search for novel molecular targets and signaling pathways following ICH. Of note, very few studies, to date, have employed a proteomic approach in elucidating the pathobiology of ICH. More importantly, we isolated single-cell suspensions of brain tissue, which enables the characterization of changes that occur at the cellular level after the hemorrhagic brain injury, and that also limits the extracellular protein-mediated masking of signals derived from low-abundant proteins during the proteomic analysis.

Materials and Methods

Induction of ICH

All animal studies were performed according to protocols approved by the Institutional Animal Care and Use Committee in accordance with the NIH and USDA guidelines. As reported previously (Bonsack et al. 2016; Sukumari-Ramesh et al. 2012a, b), ICH was induced in adult male CD-1 mice (8–12 weeks, Charles River, USA). Briefly, mice were anesthetized with ketamine and xylazine and positioned prone on a stereotaxic frame (Stoelting, WI, USA.). Throughout the surgical procedure, the body temperature was maintained at 37 ± 0.5 °C employing a temperature controller (David Kopf Instruments, USA). Using a dental drill (Dremel, USA), a burr hole (0.5 mm) was made 2.2 mm lateral to bregma, not damaging the underlying dura. A Hamilton syringe (26-G) containing 0.04 U of bacterial type IV collagenase (Sigma, USA) in 0.5 μL phosphate-buffered saline (PBS, pH 7.4) was inserted stereotaxically into the left striatum to induce spontaneous ICH. After the needle was removed, the burr hole was covered with bone wax and the incision was sutured. Sham mice underwent the same surgical procedure, but only PBS (0.5 μL) was injected.

Magnetic Resonance Imaging

Mice were anesthetized with isoflurane (2.5% for induction, 0.7% to 1.5% for maintenance mixed with O2). The body temperature was maintained by a water-heated pad and both respiration rate and temperature were monitored and controlled throughout the procedure. Magnetic resonance imaging (MRI) was obtained with a 7 Tesla small animal imaging system from Bruker (Bruker Biospin USR 70/20) using a transmit-receive surface coil. Mice were imaged using a dedicated mouse table fitted with holders to minimize movement during the imaging procedure. After positioning using a triplanar FLASH sequence, MRI studies were performed using susceptibility-weighted sequences with parameters TR = 624.7 ms, TE = 15 ms, 35° flip angle, 24 × 24 FOV, 256 × 256 matrix, 20 slices with a slice thickness of 0.5 mm, and NEX = 2, and T2-weighted MR images were obtained with parameters, TR = 2500 ms, TE = 35 ms, 24 × 24 FOV, 256 × 256 matrix, and 20 slices with a slice thickness of 0.5 mm.

Neurological Evaluation

Mice were evaluated for neurobehavioral deficits, as per our laboratory and others (Clark et al. 1998; Rosenberg et al. 1990; Sukumari-Ramesh and Alleyne 2016). The neurological deficit was estimated on a 24-point scale and the neurobehavioral analysis was comprised of 6 different tests; circling, climbing, beam walking, compulsory circling, bilateral grasp, and whisker response. Each test was graded from 0 (performance without impairment) to 4 (severe impairment) and a composite neurological deficit score was calculated as the sum of the scores on all the six tests, with a maximum neurological deficit score of 24.

Sample Preparation for LC-MS/MS (Liquid Chromatography-Mass Spectrometry/Mass Spectrometry) Analysis

Mice were deeply anesthetized and perfused transcardially with ice-cold PBS. The ipsilateral hemisphere containing both hematoma and perihematomal brain tissue was collected from the ICH animals. The respective brain area from sham animals served as the control. To study the proteomic changes that occur at the cellular level, single-cell suspensions were prepared by dissociating the brain tissue by passing through a 100-μm cell strainer using the plunger end of a 1-ml syringe and centrifuged at 1200 rpm for 5 min at 4 °C. The resultant pellet was sonicated and centrifuged at 14,000 rpm for 5 min at 4 °C and the supernatant was collected for LC-MS/MS analysis and using BCA protein assay kit (Pierce, Rockford, IL, USA) the protein concentration was estimated. One hundred micrograms of protein was reduced with dithiothreitol, alkylated using iodoacetamide, and digested overnight using sequencing-grade trypsin (catalogue number: PI90305, Thermo Scientific, USA). Digested peptides were cleaned up using C18 spin column (catalogue number: 74-4607, Harvard Apparatus, USA) and then lyophilized.

LC-MS/MS Analysis

Peptide digests were analyzed on an Orbitrap Fusion tribrid mass spectrometer (Thermo Scientific, USA) coupled with an Ultimate 3000 nano-UPLC (ultra-performance liquid chromatography) system (Thermo Scientific, USA). Two microliters of reconstituted peptide was first trapped and washed on a Pepmap100 C18 trap (5 μm, 0.3 × 5mm) at 20 μl/min using 2% acetonitrile in water (with 0.1% formic acid) for 10 min and then separated on a Pepman 100 rapid-separation liquid chromatography (RSLC) C18 column (2.0 μm, 75-μm × 150-mm) using a gradient of 2 to 40% acetonitrile with 0.1% formic acid over 100 min at a flow rate of 300 nl/min and a column temperature of 40 °C.

Samples were analyzed by data-dependent acquisition in positive mode using Orbitrap MS analyzer for precursor scan at 120,000 FWHM (full width at half maximum) from 300 to 1500 m/z and ion-trap MS analyzer for MS/MS scans at top speed mode (3-s cycle time). Collision-induced dissociation (CID) was used as a fragmentation method. Raw MS data were processed using Proteome Discoverer (v1.4, Thermo Scientific, USA) and submitted for the Sequest HT database searching algorithm against the Uniprot mouse database. The percolator PSM (peptide spectrum match) validator algorithm was used for peptide spectrum matching validation. SequestHT search parameters included 10 ppm precursor and 0.6 Da product ion mass tolerance, with static Carbamidomethylation of cysteine and dynamic oxidation of methionine.

Tissue Preparation and Western Blotting

For western blotting, the brain tissue samples in RIPA buffer containing protease and phosphatase inhibitors were centrifuged at 14,000 rpm for 5 min and the supernatant was collected. The protein concentrations were estimated by using BCA protein assay kit (Pierce, Rockford, IL, USA). Fifty micrograms of protein were separated on a 4–20% sodium dodecyl sulfate-polyacrylamide gel and transferred on to a polyvinylidene fluoride (PVDF) membrane. By using 5% non-fat dry milk, the membrane was blocked. Then, the blots were incubated with appropriate primary antibody overnight at 4 °C and followed by 2 h incubation by fluorescent-tagged secondary antibody at room temperature. Using a Li-Cor Odyssey near infrared imaging system, as previously reported by our group (Sukumari-Ramesh et al. 2016; Sukumari-Ramesh et al. 2010, 2011; Wakade et al. 2010), the blots were visualized. The densitometry analysis was performed using Quantity One software (Bio-Rad, USA), as reported earlier (Sukumari-Ramesh et al. 2016, 2010, 2011; Wakade et al. 2010). Briefly, the pixel identity of the protein band and the background was estimated, and background correction was done by background subtraction. The data was normalized with β-Actin, which served as the loading control and data was analyzed using GraphPad Prism 8 software (GraphPad Software Inc., USA).

Statistical Analysis

For quantitative LC-MS/MS analysis, the peptide spectrum match (PSM) count for each identified protein was used as a semi-quantitative measure for protein level. The PSM count for each protein in a specific sample was normalized using the total PSM counts for that sample to compensate for possible sample loading variations. After normalization across all samples, the mean PSM count for the three replicates is calculated for each protein and each sample group. The average PSM counts for each sample group were used for statistical analysis with Microsoft Excel (Microsoft Corporation, USA). The differences between the two groups were assessed by Student’s t test. A p < 0.05 was considered to be statistically significant.

Results

Mass Spectrometry-Based Identification of Differentially Expressed Proteins After ICH

ICH was induced in the striatum of male CD1 mice using collagenase injection method. Induction of ICH resulted in a hematoma as well as hemorrhagic lesion, as demonstrated by susceptibility-weighted as well as T2-weighted magnetic resonance (MR) image, respectively (Fig. 1a, b). For proteomic analysis employing mass spectrometry, we collected both hematomal and perihematomal ipsilateral brain tissue from ICH animals on day 3 post-injury, an acute time point which exhibited prominent induction of both microglial and astrocyte activation (Bonsack et al. 2016; Sukumari-Ramesh et al. 2012a, b), critical brain responses that contribute to secondary brain damage and brain recovery after ICH. Of note, single-cell suspensions of brain tissue were isolated, as described in methods and subjected to LC-MS/MS analysis. The striatal brain area from sham animals served as the experimental control. Soon before the collection of brain samples for proteomic studies, the animals were subjected to neurobehavioral analysis to confirm the induction of ICH. Notably, ICH animals exhibited very profound neurobehavioral deficits in comparison to sham (p < 0.001) (Fig. 1c).

Fig. 1
figure 1

ICH was induced in male CD1 mice, as detailed in the “Methods” section, and the susceptibility-weighted (a), as well as T2-weighted (b), magnetic resonance (MR) image depicts ICH-mediated induction of hematoma as well as the hemorrhagic lesion, respectively, in the mouse brain striatum on day 3 post-surgery (scale bar = 1 mm). The induction of ICH resulted in significant neurobehavioral deficits, which was estimated using a 24-point scale as detailed in methods, on day 3 post-surgery in comparison to sham that served as the experimental control (c). n = 8–9/animals per group. ***p < 0.001 vs. sham

Initial proteomic analysis revealed the expression of 4833 proteins between sham and ICH and the principal component analysis differentiated the proteomic profile between sham and ICH groups (Fig. 2a). Of note, 207 proteins exhibited a significant difference in their expression profile between sham and ICH, on the basis of quantitative proteomic analysis. Precisely, 46 proteins were found to be significantly upregulated, and 161 were significantly downregulated after ICH in comparison to sham (Fig. 2b, Supplementary data Table 1). To validate the data, three proteins were randomly selected and their expression changes were analyzed using western blotting. Consistent with the mass spectrometry analysis, nestin and heme oxygenase 1 were found to be upregulated after ICH, whereas clathrin light chain B was found to be downregulated after ICH in comparison to sham (Fig. 3) ascertaining the efficacy of the proteomic approach.

Fig. 2
figure 2

The principal component analysis was performed with normalized PSM counts of proteins identified by mass spectrometry using the statistical computing software R and it illustrates that the replicates in each experimental group (sham/ICH) are closely clustered (a). Heatmap representation of differentially expressed proteins between sham and ICH (b). The heatmap was generated with normalized PSM counts of differentially expressed proteins using a web server, heatmapper available at http://www.heatmapper.ca using the distance function (pearson) and clustering (average linkage). Forty-six proteins were significantly upregulated and 161 proteins were significantly downregulated after ICH in comparison to sham. Green: high expression, and red: low expression. n = 3–4 animals/group. The complete list of differentially expressed proteins is provided as supplementary data Table 1 

Fig. 3
figure 3

Validation of proteomic analysis. To validate the proteomic analysis, western blotting was performed and analyzed the expression of proteins, nestin, heme oxygenase-1 (HO-1), and clathrin light chain B after sham or ICH, as detailed in methods (a). Consistent with the proteomic studies, the densitometry analysis of the western blotting data demonstrated significant upregulation of nestin (b) as well as heme oxygenase-1 (HO-1) (c) and significant downregulation of clathrin light chain B (d) after ICH in comparison to sham. β-actin was used as a loading control (** p < 0.01, * p < 0.05 vs. sham). n = 4 animals/group

Functional Characterization of Differentially Expressed Proteins by Bioinformatics Approach

To gain insight into the functional roles of differentially expressed proteins as well as the associated signaling pathways, we analyzed the data derived from the proteomic analysis using a bioinformatics application, PANTHER (Mi et al., 2019), against the mice protein ontology database using UniPort accession number. The bioinformatics analysis resulted in the categorization of proteins into various groups based on molecular function, biological process, cellular components, protein class, and signaling pathways. As per the analysis, the prominent molecular functions that the proteins carried were of catalytic activity (38.2%) and binding (32.9%), (Fig. 4a), and the major biological processes that the proteins involved were of cellular process (46.9%), metabolic process (26.1%), localization (15.5%), and biological regulation (12.1%) (Fig. 4b). Also, most of the differentially expressed proteins were derived from cell (38.6%) and organelle (20.3%) (Fig. 4c) and based on protein class, the majority of the proteins belonged to oxidoreductase (10.1%), enzyme modulators (9.7%), cytoskeletal proteins (8.7%), hydrolases (8.2%), and transferase (6.3%) (Fig. 4d, Supplementary data Table 2).

Fig. 4
figure 4figure 4

Gene ontology analysis of the differentially expressed proteins between sham and ICH employing bioinformatics application, PANTHER. The analysis resulted in the categorization of proteins into various groups based on molecular function (a), biological process (b), cellular components (c), and protein class (d). The classification of differentially expressed proteins based on protein class is provided as supplementary data Table 2

Identification of Molecular Pathways and Associated Proteins in Relation to ICH

Further, the PANTHER-based analysis of proteomic data identified 67 different molecular pathways, and of those, the key pathways which were modulated between ICH and sham included blood coagulation, plasminogen activating cascade, p53 pathway, EGF receptor signaling pathway, FGF signaling pathway, cytoskeletal regulation by Rho GTPase, angiogenesis, integrin signaling pathway, Parkinson’s disease, Huntington disease, dopamine receptor mediated signaling pathway, T cell activation, and Wnt signaling (Fig. 5 and Supplementary data Table 3). Upon further evaluation employing PANTHER, the key proteins that were associated with the signaling cascades were alpha-2-macroglobulin, kininogen-1, plasminogen, high mobility group protein B1 (HMGB1), cyclin-dependent-like kinase 5, E3 ubiquitin-protein ligase, protein phosphatase 2A-alpha, protein phosphatase 2A-beta, serine/threonine-protein kinase PAK1, alpha-actinin-4, calpain-8, axin-1, NCK1 (non-catalytic region of tyrosine kinase adaptor protein 1), and septin-4 (Table 1 and Fig. 6). Notably, proteins such as alpha-2-macroglobulin, kininogen-1, plasminogen, and HMGB1were found to be upregulated after ICH in comparison to sham (p < 0.05) whereas cyclin-dependent-like kinase 5, E3 ubiquitin-protein ligase, protein phosphatase 2A-alpha, protein phosphatase 2A-beta, serine/threonine-protein kinase PAK1, alpha-actinin-4, calpain-8, axin-1, NCK1 (non-catalytic region of tyrosine kinase adaptor protein 1), and septin-4 were found to be downregulated after ICH in comparison to sham (p < 0.05) (Table 1).

Fig. 5
figure 5

Bioinformatics analysis of differentially expressed proteins based on signaling pathways. The prominent signaling pathways are indicated and the complete list of pathways associated with the differentially expressed proteins is provided as supplementary data Table 3

Table 1 Major signaling pathways associated with the differentially expressed proteins between sham and ICH. The description of key proteins associated with the pathways and ICH-induced protein modulation in comparison to sham is provided
Fig. 6
figure 6

The volcano plot of the entire proteome (sham and ICH) identified by mass spectrometry demonstrates the protein distribution on which proteins with statistically significant differential expression (p < 0.05) are located above the horizontal line (p value =0 .05) and the vertical lines denote a fold change of approximately ± 1.5. The key proteins associated with the identified pathways are labeled

Discussion

Intracerebral hemorrhage is a fatal stroke subtype and major public health problem. Importantly, the mortality of ICH has not changed over the past 20 years (Flaherty et al. 2006). Also, approximately 80% of survivors of ICH are often left with neurobehavioral disabilities (Agnihotri et al. 2011). Given the fact that advanced age is a risk factor of ICH (van Asch et al. 2010), the incidence of ICH is expected to increase over the next decades due to possible changes in demographics. Notably, the pathophysiology of ICH remains largely undefined.

Herein, we employed a comprehensive proteomics approach combined with bioinformatics in characterizing the molecular changes in the ipsilateral brain region at the cellular level following ICH. Using this method, we identified that the differentially expressed proteins between sham and ICH were associated with 13 prominent signaling pathways, which were blood coagulation, plasminogen activating cascade, p53 pathway, EGF receptor signaling pathway, FGF signaling pathway, cytoskeletal regulation by Rho GTPase, angiogenesis, integrin signaling pathway, Parkinson’s disease, Huntington disease, dopamine receptor mediated signaling pathway, T cell activation, and Wnt signaling.

Blood coagulation involves a sequential conversion of a series of proenzymes to functionally active proteases (Lee et al. 1996). As a consequence of ICH, blood accumulates in the brain tissue surrounding the blood vessel rupture leading to the formation of a hematoma. Cerebral hemorrhage activates blood coagulation cascade, which in turn results in the release and subsequent brain accumulation of large amounts of thrombin, a proteolytic enzyme that plays a critical role in ICH-induced perihematomal edema, a contributing factor in delayed neurologic deterioration. To this end, the mass spectrometry analysis revealed augmented expression of alpha-2-macroglobulin, a thrombin/protease inhibitor, in ICH samples compared to sham. Elevated levels of serum protein, alpha-2-macroglobulin after ICH could be due to ICH-induced damage of blood-brain barrier and subsequent accumulation of alpha-2-macroglobulin in the brain. However, the role of alpha-2-macroglobulin in modulating the proteolytic activity of thrombin and thereby ICH-induced brain injury is yet to be defined. Also, a serum protein that plays a key role in fibrinolysis, plasminogen, was found to be higher after ICH in comparison to sham and bioinformatics analysis also identified plasminogen activating cascade as a key signaling pathway after ICH. In addition, another blood coagulation pathway-related protein that was identified by the proteomic approach was kininogen-1. High-molecular-weight kininogen is an important constituent of the plasma contact-kinin system, which represents a network of serially connected serine proteases. Though kininogen appears to be a critical regulator of thrombus formation and inflammation after brain ischemia (Langhauser et al., 2012), its functional role remains largely unexplored after ICH.

ICH results in the brain accumulation of various blood components such as hemoglobin and its degradation products such as hemin, which play critical roles in secondary brain damage and neurological deficits (Madangarli et al., 2019). Consistently, hemoglobin (Hb) subunit alpha, Hb subunit beta, Hb subunit epsilon and ferritin light chain (a subunit of iron storage protein ferritin), and a heme catabolizing enzyme, heme oxygenase-1, were found to be elevated in the brain of ICH animals, in comparison to sham by mass spectrometry analysis. In addition, the complement components which enter the brain tissue following ICH (Cao et al.

Abbreviations

ICH:

intracerebral hemorrhage

PBS:

phosphate-buffered saline

LC-MS/MS:

liquid chromatography-mass spectrometry/mass spectrometry

UPLC:

ultra-performance liquid chromatography

RSLC:

rapid-separation liquid chromatography

FWHM:

full width at half maximum

CID:

collision-induced dissociation

PSM:

peptide spectrum match

RIPA:

radioimmunoprecipitation assay

BCA:

bicinchoninic acid assay

PVDF:

polyvinylidene fluoride

MRI:

magnetic resonance imaging

HMGB1:

high mobility group protein B1

NCK1:

NCK adaptor protein 1

EGF:

epidermal growth factor

EGFR:

epidermal growth factor receptor

Hb:

hemoglobin

HO-1:

heme oxygenase-1

MAC:

membrane attack complex

BDNF:

brain-derived neurotrophic factor

BBB:

blood-brain barrier

MAPK:

mitogen-activated protein kinase

FGF:

fibroblast growth factor

Pak1:

p21-activated serine/threonine kinase 1

References

  • Agnihotri S, Czap A, Staff, I, Fortunato G, McCullough LD (2011) Peripheral leukocyte counts and outcomes after intracerebral hemorrhage. J Neuroinflammation 8:160

    PubMed  PubMed Central  Google Scholar 

  • Bonsack Ft, Alleyne CH Jr, Sukumari-Ramesh S (2016) Augmented expression of TSPO after intracerebral hemorrhage: a role in inflammation? J Neuroinflammation 13:151

    PubMed  PubMed Central  Google Scholar 

  • Brooks PC (1996) Role of integrins in angiogenesis. Eur J Cancer 32A:2423–2429

    CAS  PubMed  Google Scholar 

  • Cao S, Zheng M, Hua Y, Chen G, Keep RF, ** G (2016) Hematoma changes during clot resolution after experimental intracerebral hemorrhage. Stroke 47:1626–1631

    PubMed  PubMed Central  Google Scholar 

  • Chen L, Liu W, Wang P, Xue Y, Su Q, Zeng C, Shang X (2015) Endophilin-1 regulates blood-brain barrier permeability via EGFR-JNK signaling pathway. Brain Res 1606:44–53

    CAS  PubMed  Google Scholar 

  • Clark W, Gunion-Rinker L, Lessov N, Hazel K (1998) Citicoline treatment for experimental intracerebral hemorrhage in mice. Stroke 29:2136–2140

    CAS  PubMed  Google Scholar 

  • Ducruet AF, Zacharia BE, Hickman ZL, Grobelny BT, Yeh ML, Sosunov SA, Connolly ES Jr (2009) The complement cascade as a therapeutic target in intracerebral hemorrhage. Exp Neurol 219:398–403

    CAS  PubMed  PubMed Central  Google Scholar 

  • Feigin VL, Lawes CM, Bennett DA, Barker-Collo SL, Parag V (2009) Worldwide stroke incidence and early case fatality reported in 56 population-based studies: a systematic review. Lancet Neurol 8:355–369

    PubMed  Google Scholar 

  • Flaherty ML, Haverbusch M, Sekar P, Kissela B, Kleindorfer D, Moomaw CJ, Sauerbeck L, Schneider A, Broderick JP, Woo D (2006) Long-term mortality after intracerebral hemorrhage. Neurology 66:1182–1186

    CAS  PubMed  Google Scholar 

  • Gonzalez-Forero D, Montero F, Garcia-Morales V, Dominguez G, Gomez-Perez L, Garcia-Verdugo JM, Moreno-Lopez B (2012) Endogenous rho-kinase signaling maintains synaptic strength by stabilizing the size of the readily releasable pool of synaptic vesicles. J Neurosci 32:68–84

    CAS  PubMed  PubMed Central  Google Scholar 

  • Guo FQ, Li XJ, Chen LY, Yang H, Dai HY, Wei YS, Huang YL, Yang YS, Sun HB, Xu YC, Yang ZL (2006) Study of relationship between inflammatory response and apoptosis in perihematoma region in patients with intracerebral hemorrhage. Zhongguo Wei Zhong Bing Ji Jiu Yi Xue 18:290–293

    CAS  PubMed  Google Scholar 

  • Hemphill JC 3rd, Greenberg SM, Anderson CS, Becker K, Bendok BR, Cushman M, Fung GL, Goldstein JN, Macdonald RL, Mitchell PH et al (2015) Guidelines for the management of spontaneous intracerebral hemorrhage: a guideline for healthcare professionals from the American Heart Association/American Stroke Association. Stroke 46:2032–2060

    PubMed  Google Scholar 

  • Hua Y, ** G, Keep RF, Hoff JT (2000) Complement activation in the brain after experimental intracerebral hemorrhage. J Neurosurg 92:1016–1022

    CAS  PubMed  Google Scholar 

  • Huang B, Krafft PR, Ma Q, Rolland WB, Caner B, Lekic T, Manaenko A, Le M, Tang J, Zhang JH (2012) Fibroblast growth factors preserve blood-brain barrier integrity through RhoA inhibition after intracerebral hemorrhage in mice. Neurobiol Dis 46:204–214

    CAS  PubMed  PubMed Central  Google Scholar 

  • Huangfu WC, Fuchs SY (2010) Ubiquitination-dependent regulation of signaling receptors in cancer. Genes Cancer 1:725–734

    CAS  PubMed  PubMed Central  Google Scholar 

  • Ihara M, Yamasaki N, Hagiwara A, Tanigaki A, Kitano A, Hikawa R, Tomimoto H, Noda M, Takanashi M, Mori H, Hattori N, Miyakawa T, Kinoshita M (2007) Sept4, a component of presynaptic scaffold and Lewy bodies, is required for the suppression of alpha-synuclein neurotoxicity. Neuron 53:519–533

    CAS  PubMed  Google Scholar 

  • Jayaraman L, Moorthy NC, Murthy KG, Manley JL, Bustin M, Prives C (1998) High mobility group protein-1 (HMG-1) is a unique activator of p53. Genes Dev 12:462–472

    CAS  PubMed  PubMed Central  Google Scholar 

  • Jiang J, Tang S, **a J, Wen J, Chen S, Shu X, Huen MSY, Deng Y (2018) C9orf140, a novel Axin1-interacting protein, mediates the negative feedback loop of Wnt/beta-catenin signaling. Oncogene 37:2992–3005

    CAS  PubMed  PubMed Central  Google Scholar 

  • Kang W, Balordi F, Su N, Chen L, Fishell G, Hebert JM (2014) Astrocyte activation is suppressed in both normal and injured brain by FGF signaling. Proc Natl Acad Sci U S A 111:E2987–E2995

    CAS  PubMed  PubMed Central  Google Scholar 

  • Keep RF, **ang J, Ennis SR, Andjelkovic A, Hua Y, ** G, Hoff JT (2008) Blood-brain barrier function in intracerebral hemorrhage. Acta Neurochir Suppl 105:73–77

    CAS  PubMed  Google Scholar 

  • Langhauser F, Gob E, Kraft P, Geis C, Schmitt J, Brede M, Gobel K, Helluy X, Pham M, Bendszus M et al (2012) Kininogen deficiency protects from ischemic neurodegeneration in mice by reducing thrombosis, blood-brain barrier damage, and inflammation. Blood 120:4082–4092

    CAS  PubMed  PubMed Central  Google Scholar 

  • Lee KR, Colon GP, Betz AL, Keep RF, Kim S, Hoff JT (1996) Edema from intracerebral hemorrhage: the role of thrombin. J Neurosurg 84:91–96

    CAS  PubMed  Google Scholar 

  • Li Z, Chen X, Zhang X, Ren X, Chen X, Cao J, Zang W, Liu X, Guo F (2017) Small interfering RNA targeting Dickkopf-1 contributes to neuroprotection after intracerebral hemorrhage in rats. J Mol Neurosci 61:279–288

    CAS  PubMed  Google Scholar 

  • Lim TC, Mandeville E, Weng D, Wang LS, Kurisawa M, Leite-Morris K, Selim MH, Lo EH, Spector M (2019) Hydrogel-based therapy for brain repair after intracerebral hemorrhage. Transl Stroke Res

  • Liu W, Wang P, Shang C, Chen L, Cai H, Ma J, Yao Y, Shang X, Xue Y (2014) Endophilin-1 regulates blood-brain barrier permeability by controlling ZO-1 and occludin expression via the EGFR-ERK1/2 pathway. Brain Res 1573:17–26

    CAS  PubMed  Google Scholar 

  • Loftspring MC, McDole J, Lu A, Clark JF, Johnson AJ (2009) Intracerebral hemorrhage leads to infiltration of several leukocyte populations with concomitant pathophysiological changes. J Cereb Blood Flow Metab 29:137–143

    CAS  PubMed  Google Scholar 

  • Ma QL, Yang F, Frautschy SA, Cole GM (2012) PAK in Alzheimer disease, Huntington disease and X-linked mental retardation. Cell Logist 2:117–125

    PubMed  PubMed Central  Google Scholar 

  • Madangarli N, Bonsack F, Dasari R, Sukumari-Ramesh S (2019) Intracerebral hemorrhage: blood components and neurotoxicity. Brain Sci 9

  • Mi H, Muruganujan A, Huang X, Ebert D, Mills C, Guo X, Thomas PD (2019) Protocol update for large-scale genome and gene function analysis with the PANTHER classification system (v.14.0). Nat Protoc 14:703–721

    CAS  PubMed  PubMed Central  Google Scholar 

  • Mracsko E, Javidi E, Na SY, Kahn A, Liesz A, Veltkamp R (2014) Leukocyte invasion of the brain after experimental intracerebral hemorrhage in mice. Stroke 45:2107–2114

    PubMed  Google Scholar 

  • Murakami M, Nguyen LT, Zhuang ZW, Moodie KL, Carmeliet P, Stan RV, Simons M (2008) The FGF system has a key role in regulating vascular integrity. J Clin Invest 118:3355–3366

    CAS  PubMed  PubMed Central  Google Scholar 

  • Pillai DR, Shanbhag NC, Dittmar MS, Bogdahn U, Schlachetzki F (2013) Neurovascular protection by targeting early blood-brain barrier disruption with neurotrophic factors after ischemia-reperfusion in rats. J Cereb Blood Flow Metab 33:557–566

    CAS  PubMed  PubMed Central  Google Scholar 

  • Rincon F, Mayer SA (2013) The epidemiology of intracerebral hemorrhage in the United States from 1979 to 2008. Neurocrit Care 19:95–102

    PubMed  Google Scholar 

  • Rosenberg GA, Mun-Bryce S, Wesley M, Kornfeld M (1990) Collagenase-induced intracerebral hemorrhage in rats. Stroke 21:801–807

    CAS  PubMed  Google Scholar 

  • Rowell JP, Simpson KL, Stott K, Watson M, Thomas JO (2012) HMGB1-facilitated p53 DNA binding occurs via HMG-Box/p53 transactivation domain interaction, regulated by the acidic tail. Structure 20:2014–2024

    CAS  PubMed  Google Scholar 

  • Sacco S, Marini C, Toni D, Olivieri L, Carolei A (2009) Incidence and 10-year survival of intracerebral hemorrhage in a population-based registry. Stroke 40:394–399

    PubMed  Google Scholar 

  • Schittenhelm L, Hilkens CM, Morrison VL (2017) beta2 integrins as regulators of dendritic cell, monocyte, and macrophage function. Front Immunol 8:1866

    PubMed  PubMed Central  Google Scholar 

  • Silva R, D'Amico G, Hodivala-Dilke KM, Reynolds LE (2008) Integrins: the keys to unlocking angiogenesis. Arterioscler Thromb Vasc Biol 28:1703–1713

    CAS  PubMed  Google Scholar 

  • Stromblad S, Cheresh DA (1996) Integrins, angiogenesis and vascular cell survival. Chem Biol 3:881–885

    CAS  PubMed  Google Scholar 

  • Stupack DG, Cheresh DA (2004) Integrins and angiogenesis. Curr Top Dev Biol 64:207–238

    CAS  PubMed  Google Scholar 

  • Sukumari-Ramesh S, Alleyne CH Jr (2016) Post-injury administration of tert-butylhydroquinone attenuates acute neurological injury after intracerebral hemorrhage in mice. J Mol Neurosci 58:525–531

    CAS  PubMed  Google Scholar 

  • Sukumari-Ramesh S, Laird MD, Singh N, Vender JR, Alleyne CH Jr, Dhandapani KM (2010) Astrocyte-derived glutathione attenuates hemin-induced apoptosis in cerebral microvascular cells. Glia 58:1858–1870

    PubMed  PubMed Central  Google Scholar 

  • Sukumari-Ramesh S, Singh N, Jensen MA, Dhandapani KM, Vender JR (2011) Anacardic acid induces caspase-independent apoptosis and radiosensitizes pituitary adenoma cells. J Neurosurg 114:1681–1690

    CAS  PubMed  Google Scholar 

  • Sukumari-Ramesh S, Alleyne CH Jr, Dhandapani KM (2012a) Astrocyte-specific expression of survivin after intracerebral hemorrhage in mice: a possible role in reactive gliosis? J Neurotrauma 29:2798–2804

    PubMed  PubMed Central  Google Scholar 

  • Sukumari-Ramesh S, Alleyne CH Jr, Dhandapani KM (2012b) Astrogliosis: a target for intervention in intracerebral hemorrhage? Transl Stroke Res 3:80–87

    CAS  PubMed  Google Scholar 

  • Sukumari-Ramesh S, Alleyne CH Jr, Dhandapani KM (2016) The histone deacetylase inhibitor suberoylanilide hydroxamic acid (SAHA) confers acute neuroprotection after intracerebral hemorrhage in mice. Transl Stroke Res 7:141–148

    CAS  PubMed  Google Scholar 

  • van Asch CJ, Luitse MJ, Rinkel GJ, van der Tweel I, Algra A, Klijn CJ (2010) Incidence, case fatality, and functional outcome of intracerebral haemorrhage over time, according to age, sex, and ethnic origin: a systematic review and meta-analysis. Lancet Neurol 9:167–176

    PubMed  Google Scholar 

  • Wakade C, Sukumari-Ramesh S, Laird MD, Dhandapani KM, Vender JR (2010) Delayed reduction in hippocampal postsynaptic density protein-95 expression temporally correlates with cognitive dysfunction following controlled cortical impact in mice. J Neurosurg 113:1195–1201

    CAS  PubMed  PubMed Central  Google Scholar 

  • Wang J, Chen R, Liu X, Shen J, Yan Y, Gao Y, Tao T, Shi J (2017) Hck promotes neuronal apoptosis following intracerebral hemorrhage. Cell Mol Neurobiol 37:251–261

    CAS  PubMed  Google Scholar 

  • Wang G, Li Z, Li S, Ren J, Suresh V, Xu D, Zang W, Liu X, Li W, Wang H, Guo F (2019) Minocycline preserves the integrity and permeability of BBB by altering the activity of DKK1-Wnt signaling in ICH model. Neuroscience 415:135–146

    CAS  PubMed  Google Scholar 

  • Westlin WF (2001) Integrins as targets of angiogenesis inhibition. Cancer J 7(Suppl 3):S139–S143

    PubMed  Google Scholar 

  • Xue M, Del Bigio MR (2000) Intracerebral injection of autologous whole blood in rats: time course of inflammation and cell death. Neurosci Lett 283:230–232

    CAS  PubMed  Google Scholar 

  • Xue M, Del Bigio MR (2003) Comparison of brain cell death and inflammatory reaction in three models of intracerebral hemorrhage in adult rats. J Stroke Cerebrovasc Dis 12:152–159

    PubMed  Google Scholar 

  • Yamada M, Ikeuchi T, Hatanaka H (1997) The neurotrophic action and signalling of epidermal growth factor. Prog Neurobiol 51:19–37

    CAS  PubMed  Google Scholar 

  • Yang Z, Yu A, Liu Y, Shen H, Lin C, Lin L, Wang S, Yuan B (2014) Regulatory T cells inhibit microglia activation and protect against inflammatory injury in intracerebral hemorrhage. Int Immunopharmacol 22:522–525

    CAS  PubMed  Google Scholar 

  • Zhao Y, Wei ZZ, Zhang JY, Zhang Y, Won S, Sun J, Yu SP, Li J, Wei L (2017) GSK-3beta inhibition induced neuroprotection, regeneration, and functional recovery after intracerebral hemorrhagic stroke. Cell Transplant 26:395–407

    PubMed  PubMed Central  Google Scholar 

  • Zuo S, Pan P, Li Q, Chen Y, Feng H (2017) White matter injury and recovery after hypertensive intracerebral hemorrhage. Biomed Res Int 2017:6138424

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

We would like to thank Dr. Daniel F. Linder, Biostatistician, Department of Population Health Sciences, Augusta University, for assisting us with the principal component analysis.

Funding

This work was supported by grants from the National Institutes of Health (R01NS107853) and American Heart Association (14SDG18730034) to SSR.

Author information

Authors and Affiliations

Authors

Contributions

Original draft preparation, RD and SSR; Data collection and analysis, RD, WZ, FB, and SSR; Conceptualization, writing, editing and funding acquisition, SSR.

Corresponding author

Correspondence to Sangeetha Sukumari-Ramesh.

Ethics declarations

All animal studies were performed according to protocols approved by the Institutional Animal Care and Use Committee in accordance with the NIH and USDA guidelines.

Conflict of Interest

The authors declare that they have no conflict of interest.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic Supplementary Material

ESM 1

(PDF 261 kb)

ESM 2

(DOCX 23 kb)

ESM 3

(DOCX 22 kb)

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dasari, R., Zhi, W., Bonsack, F. et al. A Combined Proteomics and Bioinformatics Approach Reveals Novel Signaling Pathways and Molecular Targets After Intracerebral Hemorrhage. J Mol Neurosci 70, 1186–1197 (2020). https://doi.org/10.1007/s12031-020-01526-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12031-020-01526-7

Keywords