AJA Asian Journal of Anesthesiology

Advancing, Capability, Improving lives

Review Article
Volume 53, Issue 1, Pages 23-28
Chung-HsiHsing 1.2.3 , Jhi-JoungWang 1.2
2981 Views


Abstract

Cytokines are key modulators of inflammatory responses, and play an important role in the defense and repair mechanisms following trauma. After traumatic injury, an immuno-inflammatory response is initiated immediately, and cytokines rapidly appear and function as a regulator of immunity. In pathologic conditions, imbalanced cytokines may provide systemic inflammatory responses or immunosuppression. Expression of perioperative cytokines vary by different intensities of surgical trauma and types of anesthesia and anesthetic agents. Inflammatory cytokines play important roles in postoperative organ dysfunction including central nervous system, cardiovascular, lung, liver, and kidney injury. Inhibition of cytokines could protect against traumatic injury in some circumstances, therefore cytokine inhibitors or antagonists might have the potential for reducing postoperative tissue/organ dysfunction. Cytokines are also involved in wound healing and post-traumatic pain. Application of cytokines for the improvement of surgical wound healing has been reported. Anesthesia-related immune response adjustment might reduce perioperative morbidity because it reduces proinflammatory cytokine expression; however, the overall effects of anesthetics on postoperative immune-inflammatory responses needs to be further investigated.

Keywords

anesthesia; cytokines; immunosuppression; surgery; systemic inflammation;


1. Introduction

Inflammation after surgical injury is characterized by increased blood flow and vascular permeability, accumulation of leukocytes, and upregulation of inflammatory mediators.1 Cytokines are key modulators of inflammation and play both inflammatory and anti-inflammatory roles.12 Over recent decades, cytokines have gained more attention in the understanding of physiological changes after trauma or surgery. Cytokines participate in acute and chronic inflammation in a complex network of interactions. Under physiologic conditions, pro- and anti-inflammatory cytokines serve as immunomodulatory elements that limit potential injury or excess inflammatory reactions. Under pathologic conditions, imbalanced cytokines may provide systemic inflammatory responses or immunosuppression.23 A dynamic and balanced shift exists between pro- and anti-inflammatory cytokines which affects organ dysfunction, immunity and infection, as well as wound healing and pain after surgery.456 In this review, we discuss the functions and changes of cytokines and the potential clinical implication of cytokine/anticytokine therapy in the perioperative period.

2. Immuno-inflammatory responses following surgical injury

Patients with surgical injury induce endogenous mediators that alter hemodynamic, metabolic, and immune responses. This immuno-inflammatory response is initiated immediately following traumatic injury.7After surgical injury, polymorphonuclear leukocytes (PMNs), endothelial cells, macrophages, and lymphocytes all become activated by the secretion of various mediators including cytokines and other molecules such as reactive oxygen species, nitric oxide, platelet activating factor, growth factors, and eicosanoids.7 Furthermore, several physiological events occur to sustain the injury: the release of adrenaline suppresses insulin secretion but stimulates secretion of growth hormone and rennin, proteolysis and glycogenolysis which enhances hepatic mediated gluconeogenesis. Glucagon is released by pancreatic islet cells which increases hepatic glucose production from a substrate that arises from tissue catabolism. The liver synthesizes a group of acute phase reactants such as C-reactive protein (CRP), protease inhibitors, and fibrinogen. Complement is also activated, resulting in limiting hemorrhage and enhanced immunity.7 Cytokines are the key mediators in the immuno-inflammatory responses. The inflammatory response to surgical injury involves a complex crosstalk between several hormones such as catecholamines, adrenocorticotropic hormone (ACTH), cortisol, glucagons, eicosanoids, and cytokines. Exposure to anesthesia and major surgery affects many of the functions of the immune-inflammatory system, and most likely damages the immune response.8 Surgery is a major traumatic element in postoperative immunodepression in normal people.9 Damage of the immune response could increase perioperative morbidity and mortality rates from infection in exposed patients.10 Both humoral and cellular immunity are dampened by surgery injury. A higher degree of surgical trauma determines greater immunodepression.11

3. Types and functions of cytokines

Cytokines are a broad and loose category of heterogeneous low molecular polypeptides or glycoproteins (8–25 kDa) including chemokines, interferons, interleukins, lymphokines, and tumor necrosis factor. They act on specific cell-surface receptors that activate intracellular JAK-STAT signals.12 Cytokines are secreted proteins whose function is communication between cells predominantly in autocrine and paracrine mechanisms.12 The functions of cytokines include cell differentiation, proliferation, survival, or even apoptosis/cell death, and inducing cytokine production and regulating immune responses.12 Cytokines are produced by immune cells (macrophages, lymphocytes, and mast cells) and nonimmune cells (endothelial cells, fibroblasts, and various stromal cells).1213 One cytokine may be produced by more than one type of cell. Cytokines play an important role in the defense and repair mechanisms following trauma, but this highly controlled system may become over exuberant after severe injuries to the host.14 Application of recombinant cytokines such as TNF-α in animal models can evoke systemic inflammatory response syndrome (SIRS), and blocking it can have beneficial effects on diseases.6 TNF-α, IL-1β, IL-6, IL-8, IL-12, and IFN-γ are probably the most important and well-studied proinflammatory cytokines after trauma.14 Another category of cytokines called alarmins that are present in systemic inflammation without evidence of a bacterial focus, suggests the presence of endogenous triggers in immune activation after trauma. Alarmins are characterized as groups of pathogen-associated molecular pattern (PAMPs) and damage-associated molecular pattern (DAMPs), which are released either after nonprogrammed cell death, excluding apoptosis, or produced and released by cells of the immune system.15 Alarmins include high mobility group box 1 (HMGB1), heat shock proteins (HSPs), defensins, cathelicidin, eosinophil-derived neurotoxin (EDN) as well as others. These structurally diverse proteins serve as endogenous mediators of innate immunity as chemoattractants and activators of antigen presenting cells (APCs).16Defensins, cathelicidin, and EDN are rapidly released from storage compartments triggered by either PAMP/DAMP recognition or proinflammatory cytokines, and then trigger immune responses. HMGB1 is a nuclear protein released by injured cells, which not only influences nuclear transactions, but also plays an important role in signaling after tissue damage.17 The receptor dedicated to the different effects of HMBG1 is the receptor for advanced glycation end product (RAGE). It is released by necrotic but not apoptotic cells, as well as secreted by activated immune cells, macrophages, mature myeloid dendritic cells (DCs), and activated NK cells without using the Golgi apparatus pathway.1819 The active secretion of HMGB1 after lipopolysaccharide stimulation seems to be partially dependent on the TLR4-CD14 complex and TGF-beta, and is triggered by cytokines as TNF- α, IL-1 β, and interferon-r.19

4. Cytokines function as a regulator of immunity after injury

Cytokines rapid appearance after injury reflects active gene transcription and translation. They bind to specific cellular receptors resulting in activation of intracellular signaling pathways that regulate gene expressions.20 Cytokines can regulate the production and activity of other cytokines, and then either augment (proinflammatory) or attenuate (anti-inflammatory) the immuno-inflammatory response. There are significant overlaps in bioactivity among different cytokines. The capacity of cytokines to activate diverse cell types and responses, highlights the pleiotropism of these inflammatory mediators. Cytokines direct the inflammatory response to sites of injury and infection, and are essential for proper wound healing processes.21 However, dysregulation of cytokine expression such as excess production of proinflammatory cytokines can induce hemodynamic instability, metabolic derangements, or even muscle wasting. In severe injuries, persistently exaggerated proinflammatory cytokine responses may contribute to systemic inflammatory response syndromes (SIRS) or multiple organ failure (MOF) and late death.21 There is now a general agreement that SIRS are accompanied by the inability to regulate the inflammatory response.22 The overproduction of inflammatory cytokines generates a systemic activation that can lead to tissue necrosis and eventually to MOF and death.23 Proinflammatory cytokines incite the production of reactive oxygen species (ROS) from various cells. Excess production of ROS causes cellular damage in vital organs seen in septic shock.242526 Severe sepsis and SIRS also induce apoptosis, which contributes to multiple organ dysfunction.27282930 Notably, the production of anti-inflammatory cytokines in these periods may attenuate the exaggerated responses. However, excessive anti-inflammatory cytokine production compromises immunity and can lead to overwhelming infectious morbidity.21

5. The effects of cytokines on tissue injury

Inflammatory cytokines play important roles in postoperative organ dysfunction. In major surgery such as cardiac surgery with cardiopulmonary bypass (CPB) that induces the release of proinflammatory cytokines, such as TNF-α,3132 IL-1β,32 IL-6, IL-8,3334 and IL-193334 which has been involved in the inflammatory cascade. Post-CPB induced acute systemic inflammation is a typical SIRS in surgical patients. This inflammatory cascade contributes to the development of postoperative complications, including respiratory failure, renal dysfunction, bleeding disorders, neurologic dysfunction, altered liver function, and ultimately, multiple organ failure.3637 It has been shown that an anti-inflammatory response may also be initiated during and after CPB. IL-10, an anti-inflammatory cytokine is likely to be induced after CPB and may play an important role in limiting post-CPB complications.3839

5.1. Cytokines and central nervous system injury

After traumatic brain injury, there is rapid activation of glial cells and additional recruitment of granulocytes, T-cells and monocytes/macrophages from the blood stream triggered by the upregulation of cell adhesion molecules, chemokines, and cytokines.40 A cascade of inflammatory mediators is produced, and contributes to the pathological consequences of central nervous system (CNS) injury.41 Cytokines and inflammatory cells are mediators in the common pathways associated with perinatal brain injury induced by a variety of insults, such as hypoxic–ischemic injury, reperfusion injury, toxin-mediated injury, and infection.42 Neuroinflammation can cause neuronal damage, but also confers neuroprotection.43 After focal cerebral ischemia, neurotoxic mediators released by microglia such as the cytokines IL-1β and TNF-α are upregulated, which contributes to secondary infarct growth. Cytokine induction from ischemic lesions involves NMDA-mediated signaling pathways and confers neuroprotection.40 There is increasing evidence that neuroinflammation represents a double-edged sword. The opposing neurotoxic and neuroprotective properties of neuroinflammation during CNS injury provide currently unexplored research problems.40

5.2. Cytokines and cardiovascular injury

Current evidence indicates that cytokines are contributors in myocardial dysfunction and cardiomyocyte necrosis in ischemia-reperfusion (I/R) injury.44 Increase in the production TNF-α, IL-1β, and IL-6 contribute to the pathology of myocardial infarction and cardiopulmonary bypass surgery.44Besides, recent studies have showed TNF-α and IL-1β link TLR4 signaling in post-ischemic cardiac dysfunction.45 In acute and chronic heart failure, elevated proinflammatory cytokines released by immune cells play a pathogenetic role in myocardial dysfunction.46 Cytokines also have a regenerative capacity of the myocardium and its blood vessels.47 The cytokines granulocyte colony-stimulating factor (GSF) and erythropoietin may stimulate cell regeneration under normal physiologic conditions and in patients with myocardial injury. In experimental cardiac injury models, the addition of cytokines has been shown to improve myocardial function.47 In vascular injury, increased expression of adhesion molecules by endothelial cells and recruitment of inflammatory cells, growth factors, and cytokines have consequent effects on vessel injury.48 Circulating TNF-α, IL-1β, and IL-8 interact with specific receptors on endothelial cells that activate JAK-STAT, nuclear factor (NF)-κB, and Smad signaling pathways leading to induced cell adhesion, apoptosis, and permeability.48 Cytokines also interact with integrins and matrix metalloproteinases (MMPs) and modify extracellular matrix composition. Persistent increase of cytokines is associated with vascular dysfunctions such as atherosclerosis, aneurysm, and hypertension.

5.3. Cytokines and acute lung injury

Cytokine networks on cells of the alveolar–capillary membrane are necessary for cellular communication during pulmonary inflammation, and the subsequent events of these interactions are pivotal to the immuno-inflammatory responses leading to acute lung injury (ALI).49 Proinflammatory cytokines are known to play roles in ischemia–reperfusion injury of the lung.50 The dysregulated expression of cytokines and growth factors in response to infectious or harmful insults can provoke deleterious lung inflammation. Early response cytokines, adhesion molecules, and the chemokine IL-8 could promote the recruitment of neutrophils into lung tissue. TNF-α or IL-1β activate microvascular endothelium leading to the expression of endothelial cell-derived E- and P-selectins and ICAM-1, and then recruits leukocytes. These activated leukocytes can release reactive oxygen metabolites, proteolytic enzymes, and additional cytokines and then induce ALI. Recent studies further demonstrated that TNF-α plays a central role in the development of pulmonary edema in ALI through activation of TNF receptor p55-mediated caspase 8 death signaling.51

5.4. Cytokines and hepatic injury

Proinflammatory cytokines causes activation and priming of neutrophils for reactive oxygen formation and recruits them into the vascular beds of the liver and induce hepatic injury.52 The phenomenon is similar to cytokine-related lung injury. Cytokines as chemotactic signals from the parenchyma will trigger extravasation of leukocytes and attack hepatocytes. Leukocyte adhesion induces degranulation with release of proteases and formation of reactive oxygen species, which diffuse into hepatocytes and induce an intracellular oxidant stress and mitochondrial dysfunction and then cell death.52 In addition, necrotic cells release mediators such as high-mobility group box-1 (HMGB-1), which further promotes neutrophilic hepatitis and tissue damage.

5.5. Cytokines and acute kidney injury

Cytokines have been implicated in the pathobiology of acute kidney injury (AKI). Intravenous administration of TNF-α decreased glomerular filtration ratio (GFR) and led to damage of the glomerular endothelial surface, which is an important determinant of acute kidney injury in sepsis.53 In addition, IL-1β-mediated neutrophil recruitment is likely to be a key process in AKI.54Recently, the effects of IL-6 on AKI were confirmed.55 IL-6-deficient mice were resistant to HgCl2-induced AKI and neutrophil infiltration. Renal IL-6 expression and STAT3 activation in renal tubular epithelial cells were significantly increased during the development of kidney injury and correlate with the onset and severity of AKI.55 It is now believed that the IL-6/IL-6R axis plays a critical role in acute kidney injury.55 The recently discovered cytokine, IL-19 also mediates tissue damage in murine ischemic AKI.56

6. Cytokines in wound healing

After surgical injury, wound healing is proceeded by hemostasis, inflammation, proliferation, and tissue remodeling.57 Cutaneous wound healing is an instant response to a wound, which repairs damaged lesions and restores dermal structure and functions.58 The extracellular matrix, growth factors, and inflammatory mediators, and cytokines are critical for cutaneous wound healing.5960 Keratinocyte growth factor (KGF), a fibroblast growth factor (FGF) family of mitogens, is strongly upregulated in dermal fibroblasts after a skin injury6162, and is essential for wound re-epithelialization.6163Cytokines can induce KGF expression in fibroblasts.64 IL-1β, IL-6, and TNF-α were identified as strong stimulators of KGF expression in fibroblasts.6566Moreover, the expression of these cytokines after an injury correlated with the time course of KGF expression.67 A recent report showed that wounds in IL-6-deficient mice showed delays in macrophage infiltration, fibrin clearance, and wound contraction.68 IL-6 modulates immune responses and is essential for the wound healing process.68 Recently it was identified that IL-19 directly regulates KGF expression during wound healing.69 IL-19 induced IL-1β, IL-6, TGF-β, MMP2, MMP9, and CXCR4 expression,70 which contribute to cutaneous wound healing. Furthermore, applied IL-19 protein on surgical wounds in mice can promote a cutaneous wound healing process.69 We thus should consider that in some circumstances, inflammatory cytokines may be applied as a therapeutic agent for the improvement of surgical wounds.

7. Cytokines in post-traumatic pain

Evidence has showed that TNF-α plays an important role in T-cell-mediated tissue injury, and targeting anti-inflammatory treatment can ameliorate injury-induced neuropathic pain.71 A recent study showed the positive effects of TNF-α antagonist etanercept on functional recovery and reducing hypersensitivity after peripheral nerve crush injury.71 It was suggested that etanercept optimizes the involvement of macrophages and the secretion of inflammatory mediators in pain. Dahl and Cohen showed that perineural injection of etanercept can treat postamputation pain.72 They used perineural etanercept in six traumatic amputees with postamputation pain. Three months after injections, five of the six patients showed significant improvements in residual limb pain and functional capacity.72 Etanercept also reduced acute sciatica secondary to lumbar disc herniation in a triple-blind randomized controlled trial.73 IL-1β also affects post-traumatic pain. Schafer et al74 showed that IL-1β attenuated pain perception after surgery by promoting the release of β-endorphins from the pituitary gland and increasing the number of central opioid like receptors. Prostacyclin is an important mediator of peripheral pain sensation. Recently, Schuh et al75presented that early synthesis of prostacyclin at the site of injury causes accumulation of IL1β-expressing macrophages as a key step in neuropathic pain after traumatic injury. One may consider that inhibiting or antagonizing inflammatory cytokines could be applied to the treatment of post-traumatic pain in the future.

8. The impact of anesthesia and surgery on cytokine expression

Surgeries harm homeostasis and generate various hemodynamic, metabolic, and immunologic reactions. Surgical injury is correlated with impaired postoperative immune responses, which could be associated with dysregulated proinflammatory cytokines, or with the inhibition of cellular responses.976 In surgical patients, the major altered serum levels of cytokines are interleukin (IL)-1α, IL-1β, IL-2, IL-4, IL-6, IL-10, tumor necrosis factor-α (TNF-α) and interferon-γ (IFN-γ).76 Recently, newly discovered cytokines IL-19 and IL-22 were found to increase after cardiac surgery.35Anesthesia is also able to suppress immune function, thus influencing the postoperative immune-inflammatory responses. The combination between general anesthesia and surgical stress may have an impact on the inflammatory responses, which are vital for preserving the postoperative homeostatic state. Multiple anesthetics have been suggested to reduce the immune functions by altering the performance of immune-competent cells and adjusting stress responses.976 Anesthetic techniques can reduce stress response and cytokine activation after surgery.976 The different anesthetic techniques could affect various immune-inflammation responses after surgery. Although both general and regional anesthesia suppress host immunity, recent studies demonstrated that serum from patients with breast cancer surgery under general anesthesia showed lower levels of IL-1β and IL-10 and reduced NK cell activation ability as compared with regional block plus propofol sedation.77 Another effect of anesthetics on cytokine production is through influence on adrenergic activity. There is a positive correlation between the activity of adrenergic receptors and the production of cytokines. Elevated epinephrine concentrations can trigger TNF-α and IL-6 release by activating α2 and β2 adrenergic receptors in macrophages, respectively.7879The impact of volatile anesthetic on cytokine secretion could be related to intracellular calcium concentrations because Ca2+ is a vital step on cytokine regulation. 

 The intracellular Ca2+ pools can be altered by volatile anesthetics. Previous studies showed that halothane restricted ATP-stimulated Ca2+transients in endothelial cells.80 Thus, inhalational anesthetics reduce adrenergic activity, intracellular calcium concentration, inhibits inflammatory cytokine production and affects postoperative immunomodulatory responses. The intravenous anesthetic propofol is widely used for surgical anesthesia and patient sedation for intensive care. Propofol has anti-inflammatory effects on inhibition of stimuli-induced TNF-α, IL-1β, IL-6, IL-8, and IL-10 production.81828384 Propofol also suppresses inducible NO synthase/NO biosynthesis in endotoxin lipopolysaccharide (LPS)-activated macrophages.85The molecular mechanisms for anti-inflammatory propofol include inhibiting NF-κB activation f86 and decreasing mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase (ERK) activation.8788Recent studies demonstrated that propofol can protect multiple organ injuries in sepsis by reducing inflammatory cytokines production.8990 One may consider that anesthesia-related immune response adjustment could reduce perioperative morbidity by reducing proinflammatory cytokine production; however, the overall effects of anesthetics on postoperative pathophysiological responses especially in patients with pre-existing immunological disorders or malignancy needs to be further investigated.

9. Inhibit cytokines could protect traumatic injury

The implication of cytokine inhibitors in traumatic injury have been investigated, although the effects of cytokine inhibition on surgical injury needs to be further investigated. The TNF-α antagonist, etanercept, can be used to reduce traumatic brain injury (TBI) in rats.91 Cheong et al91 showed that neurological and motor deficits, cerebral contusion, and increased brain TNF-α contents caused by TBI can be attenuated by etanercept therapy. They found that etanercept may penetrate directly into the contused brain tissues and improve outcomes of TBI by reducing brain TNF-α and stimulating neurogenesis.91 The TNF-α antagonist also reduces apoptosis of neurons and oligodendroglia in rat spinal cord injury (SCI).92 This positive effect of etanercept on spinal cord injury is probably attributable to the suppression of TNF-α, TNFR1, TNFR2, and activated caspase-3 and caspase-8 overexpressions. Anti-IL-6-receptor antibody also promotes repair of spinal cord injury in mice.93 Mukaino et al93 used an anti-mouse IL-6 receptor antibody to improve motor function after SCI. This change was accompanied by reducing chemokines CCL2 and CCL5 recruitment and decreasing CXCL10 expression.93 In post-traumatic ischemic/reperfused hearts, etanercept markedly inhibited oxidative/nitrative stress and myocardial injury by reducing TNF-α.94 Inhibiting TNF-α could also improve bone formation following trauma. Previous studies also demonstrated that TNF-α negatively regulates bone formation at the injured growth plate in rats.95 TNF-α mediates p38 activation, influences osteoblast recruitment, proliferation and differentiation at the injured growth plate.95 HMGB1, released by injured cells, plays an important role in signaling after tissue damage.17 The HMGB1 inhibitor, glycyrrhizin can relieve the severity of traumatic pancreatitis in rats.96 Taken together, cytokine inhibitors or antagonists might reduce traumatic injury and has the potential for reducing postoperative tissue/organ dysfunction.

10. Conclusion

Surgical trauma may induce acute systemic inflammation which originally plays a role in immune defense from bacterial infection and in the wound healing process. Cytokines are major modulators of inflammatory responses; however, cytokine dysregulation may provide systemic inflammatory responses or immunosuppression leading to multiple organ dysfunction or infectious disorders. Inhibit cytokines could protect organ injury in some circumstances, therefore, cytokine inhibitors or antagonists might have the potential for reducing postoperative tissue/organ dysfunction. Anesthesia-related immune response adjustment might reduce proinflammatory cytokine production. The overall effects of anesthetics on perioperative cytokine production and pathophysiological responses needs to be further investigated.

Acknowledgments

This work was supported by grant NSC101-2314-B-384-003-MY3 from the Taiwan National Science Council.


References

1
A. Lenz, G.A. Franklin, W.G. Cheadle
Systemic inflammation after trauma
Injury, 38 (2007), pp. 1336-1345
2
C. Munoz, J. Carlet, C. Fitting, B. Misset, J.P. Bleriot, J.M. Cavaillon
Dysregulation of in vitro cytokine production by monocytes during sepsis
J Clin Invest, 88 (1991), pp. 1747-1754
3
T. Kasai, K. Inada, T. Takakuwa, Y. Yamada, Y. Inoue, T. Shimamura, et al.
Anti-inflammatory cytokine levels in patients with septic shock
Res Commun Mol Pathol Pharmacol, 98 (1997), pp. 34-42
4
S.M. Opal, A.S. Cross, J.W. Jhung, L.D. Young, J.E. Palardy, N.A. Parejo, et al.
Potential hazards of combination immunotherapy in the treatment of experimental septic shock
J Infect Dis, 173 (1996), pp. 1415-1421
5
E. Lin, S.E. Calvano, S.F. Lowry
Inflammatory cytokines and cell response in surgery
Surgery, 127 (2000), pp. 117-126
6
Y.M. Yao, H. Redl, S. Bahrami, G. Schlag
The inflammatory basis of trauma/shock-associated multiple organ failure
Inflamm Res, 47 (1998), pp. 201-210
7
M.D. Cipolle, M.D. Pasquale, F.B. Cerra
Secondary organ dysfunction. From clinical perspectives to molecular mediators
Crit Care Clin, 9 (1993), pp. 261-298
8
G.W. Stevenson, S.C. Hall, S. Rudnick, F.L. Seleny, H.C. Stevenson
The effect of anesthetic agents on the human immune response
Anesthesiology, 72 (1990), pp. 542-552
9
M. Salo
Effects of anaesthesia and surgery on the immune response
Acta Anaesthesiol Scand, 36 (1992), pp. 201-220
10
D.L. Bruce, D.W. Wingard
Anesthesia and the immune response
Anesthesiology, 34 (1971), pp. 271-282
11
B. Beilin, Y. Shavit, J. Hart, B. Mordashov, S. Cohn, I. Notti, et al.
Effects of anesthesia based on large versus small doses of fentanyl on natural killer cell cytotoxicity in the perioperative period
Anesth Analg, 82 (1996), pp. 492-497
12
J.B. Spangler, I. Moraga, J.L. Mendoza, K.C. Garcia
Insights into cytokine-receptor interactions from cytokine engineering
Annu Rev Immunol (2014) PMID: 25493332
Article  
13
C.H. Hsing, H.H. Li, Y.H. Hsu, C.L. Ho, S.S. Chuang, K.M. Lan, et al.
The distribution of interleukin-19 in healthy and neoplastic tissue
Cytokine, 44 (2008), pp. 221-228
14
L. Hoover, G.V. Bochicchio, L.M. Napolitano, M. Joshi, K. Bochicchio, W. Meyer, et al.
Systemic inflammatory response syndrome and nosocomial infection in trauma
J Trauma, 61 (2006), pp. 310-316 discussion 6–7
15
M.E. Bianchi
DAMPs, PAMPs and alarmins: all we need to know about danger
J Leukoc Biol, 81 (2007), pp. 1-5
16
J.J. Oppenheim, D. Yang
Alarmins: chemotactic activators of immune responses
Curr Opin Immunol, 17 (2005), pp. 359-365
17
H.E. Harris, A. Raucci
Alarmin(g) news about danger: workshop on innate danger signals and HMGB1
EMBO Rep, 7 (2006), pp. 774-778
18
S. Gardella, C. Andrei, D. Ferrera, L.V. Lotti, M.R. Torrisi, M.E. Bianchi, et al.
The nuclear protein HMGB1 is secreted by monocytes via a non-classical, vesicle-mediated secretory pathway
EMBO Rep, 3 (2002), pp. 995-1001
19
P. Scaffidi, T. Misteli, M.E. Bianchi
Release of chromatin protein HMGB1 by necrotic cells triggers inflammation
Nature, 418 (2002), pp. 191-195
20
J. Spink, J. Cohen
Synergy and specificity in induction of gene activity by proinflammatory cytokines: potential therapeutic targets
Shock, 7 (1997), pp. 405-412
21
J.P. Desborough
The stress response to trauma and surgery
Br J Anaesth, 85 (2000), pp. 109-117
22
E. Kilger, F. Weis, J. Briegel, L. Frey, A.E. Goetz, D. Reuter, et al.
Stress doses of hydrocortisone reduce severe systemic inflammatory response syndrome and improve early outcome in a risk group of patients after cardiac surgery
Crit Care Med, 31 (2003), pp. 1068-1074
23
M.G. Netea, J.W. van der Meer, M. van Deuren, B.J. Kullberg
Proinflammatory cytokines and sepsis syndrome: not enough, or too much of a good thing?
Trends Immunol, 24 (2003), pp. 254-258
24
E.D. Crouser
Therapeutic benefits of antioxidants during sepsis: is protection against oxidant-mediated tissue damage only half the story?
Crit Care Med, 32 (2004), pp. 589-590
25
C. Ritter, M. Andrades, M.L. Frota Junior, F. Bonatto, R.A. Pinho, M. Polydoro, et al.
Oxidative parameters and mortality in sepsis induced by cecal ligation and perforation
Intensive Care Med, 29 (2003), pp. 1782-1789
26
D.E. Taylor, A.J. Ghio, C.A. Piantadosi
Reactive oxygen species produced by liver mitochondria of rats in sepsis
Arch Biochem Biophys, 316 (1995), pp. 70-76
27
C. Power, N. Fanning, H.P. Redmond
Cellular apoptosis and organ injury in sepsis: a review
Shock, 18 (2002), pp. 197-211
28
M. Perl, C.S. Chung, J. Lomas-Neira, T.M. Rachel, W.L. Biffl, W.G. Cioffi, et al.
Silencing of Fas, but not caspase-8, in lung epithelial cells ameliorates pulmonary apoptosis, inflammation, and neutrophil influx after hemorrhagic shock and sepsis
Am J Pathol, 167 (2005), pp. 1545-1559
29
C.M. Coopersmith, P.E. Stromberg, W.M. Dunne, C.G. Davis, D.M. Amiot 2nd, T.G. Buchman, et al.
Inhibition of intestinal epithelial apoptosis and survival in a murine model of pneumonia-induced sepsis
Jama, 287 (2002), pp. 1716-1721
30
H. Blumberg, D. Conklin, W.F. Xu, A. Grossmann, T. Brender, S. Carollo, et al.
Interleukin 20: discovery, receptor identification, and role in epidermal function
Cell, 104 (2001), pp. 9-19
31
N.J. Jansen, W. van Oeveren, L. van den Broek, H.M. Oudemans-van Straaten, C.P. Stoutenbeek, M.C. Joen, et al.
Inhibition by dexamethasone of the reperfusion phenomena in cardiopulmonary bypass
J Thorac Cardiovasc Surg, 102 (1991), pp. 515-525
32
P. Menasche, S. Haydar, J. Peynet, C. Du Buit, R. Merval, G. Bloch, et al.
A potential mechanism of vasodilation after warm heart surgery. The temperature-dependent release of cytokines
J Thorac Cardiovasc Surg, 107 (1994), pp. 293-299
33
A. Finn, S. Naik, N. Klein, R.J. Levinsky, S. Strobel, M. Elliott
Interleukin-8 release and neutrophil degranulation after pediatric cardiopulmonary bypass
J Thorac Cardiovasc Surg, 105 (1993), pp. 234-241
34
T. Kawamura, R. Wakusawa, K. Okada, S. Inada
Elevation of cytokines during open heart surgery with cardiopulmonary bypass: participation of interleukin 8 and 6 in reperfusion injury
Can J Anaesth, 40 (1993), pp. 1016-1021
35
C.H. Hsing, M.Y. Hsieh, W.Y. Chen, E. Cheung So, B.C. Cheng, M.S. Chang
Induction of interleukin-19 and interleukin-22 after cardiac surgery with cardiopulmonary bypass
Ann Thorac Surg, 81 (2006), pp. 2196-2201
36
J. Cremer, M. Martin, H. Redl, S. Bahrami, C. Abraham, T. Graeter, et al.
Systemic inflammatory response syndrome after cardiac operations
Ann Thorac Surg, 61 (1996), pp. 1714-1720
37
L.H. Edmunds Jr.
Inflammatory response to cardiopulmonary bypass
Ann Thorac Surg, 66 (1998), pp. S12-S16 discussion S25–8
38
S. Wan, J.L. LeClerc, J.L. Vincent
Cytokine responses to cardiopulmonary bypass: lessons learned from cardiac transplantation
Ann Thorac Surg, 63 (1997), pp. 269-276
39
P. Giomarelli, S. Scolletta, E. Borrelli, B. Biagioli
Myocardial and lung injury after cardiopulmonary bypass: role of interleukin (IL)-10
Ann Thorac Surg, 76 (2003), pp. 117-123
40
G. Stoll, S. Jander, M. Schroeter
Detrimental and beneficial effects of injury-induced inflammation and cytokine expression in the nervous system
Adv Exp Med Biol, 513 (2002), pp. 87-113
41
A. Helmy, M.G. De Simoni, M.R. Guilfoyle, K.L. Carpenter, P.J. Hutchinson
Cytokines and innate inflammation in the pathogenesis of human traumatic brain injury
Prog Neurobiol, 95 (2011), pp. 352-372
42
R.M. McAdams, S.E. Juul
The role of cytokines and inflammatory cells in perinatal brain injury
Neurol Res Int, 2012 (2012), p. 561494
Article  
43
H.J. Kadhim, J. Duchateau, G. Sebire
Cytokines and brain injury: invited review
J Intensive Care Med, 23 (2008), pp. 236-249
44
M. Zhang, L. Chen
Status of cytokines in ischemia reperfusion induced heart injury
Cardiovasc Hematol Disord Drug Targets, 8 (2008), pp. 161-172
45
J. Cha, Z. Wang, L. Ao, N. Zou, C.A. Dinarello, A. Banerjee, et al.
Cytokines link Toll-like receptor 4 signaling to cardiac dysfunction after global myocardial ischemia
Ann Thorac Surg, 85 (2008), pp. 1678-1685
46
D. Chen, C. Assad-Kottner, C. Orrego, G. Torre-Amione
Cytokines and acute heart failure
Crit Care Med, 36 (2008), pp. S9-S16
47
G. Srinivas, P. Anversa, W.H. Frishman
Cytokines and myocardial regeneration: a novel treatment option for acute myocardial infarction
Cardiol Rev, 17 (2009), pp. 1-9
48
A.H. Sprague, R.A. Khalil
Inflammatory cytokines in vascular dysfunction and vascular disease
Biochem Pharmacol, 78 (2009), pp. 539-552
49
R.M. Strieter, S.L. Kunkel
Acute lung injury: the role of cytokines in the elicitation of neutrophils
J Investig Med, 42 (1994), pp. 640-651
50
B. Krishnadasan, B.V. Naidu, K. Byrne, C. Fraga, E.D. Verrier, M.S. Mulligan
The role of proinflammatory cytokines in lung ischemia-reperfusion injury
J Thorac Cardiovasc Surg, 125 (2003), pp. 261-272
51
B.V. Patel, M.R. Wilson, K.P. O'Dea, M. Takata
TNF-induced death signaling triggers alveolar epithelial dysfunction in acute lung injury
J Immunol, 190 (2013), pp. 4274-4282
52
H. Jaeschke
Mechanisms of Liver Injury. II. Mechanisms of neutrophil-induced liver cell injury during hepatic ischemia-reperfusion and other acute inflammatory conditions
Am J Physiol Gastrointest Liver Physiol, 290 (2006), pp. G1083-G1088
53
C. Xu, A. Chang, B.K. Hack, M.T. Eadon, S.L. Alper, P.N. Cunningham
TNF-mediated damage to glomerular endothelium is an important determinant of acute kidney injury in sepsis
Kidney Int, 85 (2014), pp. 72-81
54
M. Berry, M.R. Clatworthy
Immunotherapy for acute kidney injury
Immunotherapy, 4 (2012), pp. 323-334
55
Y. Nechemia-Arbely, D. Barkan, G. Pizov, A. Shriki, S. Rose-John, E. Galun, et al.
IL-6/IL-6R axis plays a critical role in acute kidney injury
J Am Soc Nephrol, 19 (2008), pp. 1106-1115
56
Y.H. Hsu, H.H. Li, J.M. Sung, W.T. Chen, Y.C. Hou, M.S. Chang
Interleukin-19 mediates tissue damage in murine ischemic acute kidney injury
PLoS One, 8 (2013), p. e56028
CrossRef  
57
P. Martin
Wound healing–aiming for perfect skin regeneration
Science, 276 (1997), pp. 75-81
58
D.T. Woodley, E.J. O'Keefe, M. Prunieras
Cutaneous wound healing: a model for cell-matrix interactions
J Am Acad Dermatol, 12 (1985), pp. 420-433
59
K.S. Midwood, L.V. Williams, J.E. Schwarzbauer
Tissue repair and the dynamics of the extracellular matrix
Int J Biochem Cell Biol, 36 (2004), pp. 1031-1037
60
P. Martin, J. Hopkinson-Woolley, J. McCluskey
Growth factors and cutaneous wound repair
Prog Growth Factor Res, 4 (1992), pp. 25-44
61
S. Werner, K.G. Peters, M.T. Longaker, F. Fuller-Pace, M.J. Banda, L.T. Williams
Large induction of keratinocyte growth factor expression in the dermis during wound healing
Proc Natl Acad Sci U S A, 89 (1992), pp. 6896-6900
62
H.D. Beer, M.G. Gassmann, B. Munz, H. Steiling, F. Engelhardt, K. Bleuel, et al.
Expression and function of keratinocyte growth factor and activin in skin morphogenesis and cutaneous wound repair
J Investig Dermatol Symp Proc, 5 (2000), pp. 34-39
63
S. Werner, H. Smola, X. Liao, M.T. Longaker, T. Krieg, P.H. Hofschneider, et al.
The function of KGF in morphogenesis of epithelium and reepithelialization of wounds
Science, 266 (1994), pp. 819-822
64
M. Brauchle, K. Angermeyer, G. Hubner, S. Werner
Large induction of keratinocyte growth factor expression by serum growth factors and pro-inflammatory cytokines in cultured fibroblasts
Oncogene, 9 (1994), pp. 3199-3204
65
M. Chedid, J.S. Rubin, K.G. Csaky, S.A. Aaronson
Regulation of keratinocyte growth factor gene expression by interleukin 1
J Biol Chem, 269 (1994), pp. 10753-10757
66
J.A. Quayle, S. Adams, R.C. Bucknall, S.W. Edwards
Cytokine expression by inflammatory neutrophils
FEMS Immunol Med Microbiol, 8 (1994), pp. 233-239
67
G. Hubner, M. Brauchle, H. Smola, M. Madlener, R. Fassler, S. Werner
Differential regulation of pro-inflammatory cytokines during wound healing in normal and glucocorticoid-treated mice
Cytokine, 8 (1996), pp. 548-556
68
M.M. McFarland-Mancini, H.M. Funk, A.M. Paluch, M. Zhou, P.V. Giridhar, C.A. Mercer, et al.
Differences in wound healing in mice with deficiency of IL-6 versus IL-6 receptor
J Immunol, 184 (2010), pp. 7219-7228
69
D.P. Sun, C.H. Yeh, E. So, L.Y. Wang, T.S. Wei, M.S. Chang, et al.
Interleukin (IL)-19 promoted skin wound healing by increasing fibroblast keratinocyte growth factor expression
Cytokine, 62 (2013), pp. 360-368
70
C.H. Hsing, H.C. Cheng, Y.H. Hsu, C.H. Chan, C.H. Yeh, C.F. Li, et al.
Upregulated IL-19 in breast cancer promotes tumor progression and affects clinical outcome
Clin Cancer Res, 18 (2012), pp. 713-725
71
K. Iwatsuki, T. Arai, H. Ota, S. Kato, T. Natsume, S. Kurimoto, et al.
Targeting anti-inflammatory treatment can ameliorate injury-induced neuropathic pain
PLoS One, 8 (2013), p. e57721
CrossRef  
72
E. Dahl, S.P. Cohen
Perineural injection of etanercept as a treatment for postamputation pain
Clin J Pain, 24 (2008), pp. 172-175
73
T. Okoro, S.I. Tafazal, S. Longworth, P.J. Sell
Tumor necrosis alpha-blocking agent (etanercept): a triple blind randomized controlled trial of its use in treatment of sciatica
J Spinal Disord Tech, 23 (2010), pp. 74-77
74
M. Schafer, L. Carter, C. Stein
Interleukin 1 beta and corticotropin-releasing factor inhibit pain by releasing opioids from immune cells in inflamed tissue
Proc Natl Acad Sci U S A, 91 (1994), pp. 4219-4223
75
C.D. Schuh, S. Pierre, A. Weigert, B. Weichand, K. Altenrath, Y. Schreiber, et al.
Prostacyclin mediates neuropathic pain through interleukin 1beta-expressing resident macrophages
Pain, 155 (2014), pp. 545-555
76
H.E. Hogevold, T. Lyberg, H. Kahler, E. Haug, O. Reikeras
Changes in plasma IL-1beta, TNF-alpha and IL-6 after total hip replacement surgery in general or regional anaesthesia
Cytokine, 12 (2000), pp. 1156-1159
77
A. Buckley, S. McQuaid, P. Johnson, D.J. Buggy
Effect of anaesthetic technique on the natural killer cell anti-tumour activity of serum from women undergoing breast cancer surgery: a pilot study
Br J Anaesth, 113 (Suppl. 1) (2014), pp. i56-62
78
K.A. Boost, M. Flondor, C. Hofstetter, I. Platacis, K. Stegewerth, S. Hoegl, et al.
The beta-adrenoceptor antagonist propranolol counteracts anti-inflammatory effects of isoflurane in rat endotoxemia
Acta Anaesthesiol Scand, 51 (2007), pp. 900-908
79
I.J. Elenkov, G. Hasko, K.J. Kovacs, E.S. Vizi
Modulation of lipopolysaccharide-induced tumor necrosis factor-alpha production by selective alpha- and beta-adrenergic drugs in mice
J Neuroimmunol, 61 (1995), pp. 123-131
80
V.K. Singh, V.S. Yadav
Role of cytokines and growth factors in radioprotection
Exp Mol Pathol, 78 (2005), pp. 156-169
81
P.E. Marik
Propofol: an immunomodulating agent
Pharmacotherapy, 25 (2005), pp. 28S-33S
CrossRef  
82
T. Taniguchi, H. Kanakura, Y. Takemoto, Y. Kidani, K. Yamamoto
Effects of ketamine and propofol on the ratio of interleukin-6 to interleukin-10 during endotoxemia in rats
Tohoku J Exp Med, 200 (2003), pp. 85-92
83
Y. Takemoto
Dose effects of propofol on hemodynamic and cytokine responses to endotoxemia in rats
J Anesth, 19 (2005), pp. 40-44
84
B.G. Hsu, F.L. Yang, R.P. Lee, T.C. Peng, H.I. Chen
Effects of post-treatment with low-dose propofol on inflammatory responses to lipopolysaccharide-induced shock in conscious rats
Clin Exp Pharmacol Physiol, 32 (2005), pp. 24-29
85
R.M. Chen, T.G. Chen, T.L. Chen, L.L. Lin, C.C. Chang, H.C. Chang, et al.
Anti-inflammatory and antioxidative effects of propofol on lipopolysaccharide-activated macrophages
Ann N Y Acad Sci, 1042 (2005), pp. 262-271
86
X.M. Song, Y.L. Wang, J.G. Li, C.Y. Wang, Q. Zhou, Z.Z. Zhang, et al.
Effects of propofol on pro-inflammatory cytokines and nuclear factor kappaB during polymicrobial sepsis in rats
Mol Biol Rep, 36 (2009), pp. 2345-2351
87
B. Jawan, Y.H. Kao, S. Goto, M.C. Pan, Y.C. Lin, L.W. Hsu, et al.
Propofol pretreatment attenuates LPS-induced granulocyte-macrophage colony-stimulating factor production in cultured hepatocytes by suppressing MAPK/ERK activity and NF-kappaB translocation
Toxicol Appl Pharmacol, 229 (2008), pp. 362-373
88
W.T. Chiu, Y.L. Lin, C.W. Chou, R.M. Chen
Propofol inhibits lipoteichoic acid-induced iNOS gene expression in macrophages possibly through downregulation of toll-like receptor 2-mediated activation of Raf-MEK1/2-ERK1/2-IKK-NFkappaB
Chem Biol Interact, 181 (2009), pp. 430-439
89
C.H. Yeh, W. Cho, E.C. So, C.C. Chu, M.C. Lin, J.J. Wang, et al.
Propofol inhibits lipopolysaccharide-induced lung epithelial cell injury by reducing hypoxia-inducible factor-1alpha expression
Br J Anaesth, 106 (2011), pp. 590-599
90
C.H. Hsing, W. Chou, J.J. Wang, H.W. Chen, C.H. Yeh
Propofol increases bone morphogenetic protein-7 and decreases oxidative stress in sepsis-induced acute kidney injury
Nephrol Dial Transplant, 26 (2011), pp. 1162-1172
91
C.U. Cheong, C.P. Chang, C.M. Chao, B.C. Cheng, C.Z. Yang, C.C. Chio
Etanercept attenuates traumatic brain injury in rats by reducing brain TNF- alpha contents and by stimulating newly formed neurogenesis
Mediators Inflamm, 2013 (2013), p. 620837
Article  
92
K.B. Chen, K. Uchida, H. Nakajima, T. Yayama, T. Hirai, S. Watanabe, et al.
Tumor necrosis factor-alpha antagonist reduces apoptosis of neurons and oligodendroglia in rat spinal cord injury
Spine (Phila Pa 1976), 36 (2011), pp. 1350-1358
93
M. Mukaino, M. Nakamura, O. Yamada, S. Okada, S. Morikawa, F. Renault-Mihara, et al.
Anti-IL-6-receptor antibody promotes repair of spinal cord injury by inducing microglia-dominant inflammation
Exp Neurol, 224 (2010), pp. 403-414
94
S. Liu, T. Yin, X. Wei, W. Yi, Y. Qu, Y. Liu, et al.
Downregulation of adiponectin induced by tumor necrosis factor alpha is involved in the aggravation of posttraumatic myocardial ischemia/reperfusion injury
Crit Care Med, 39 (2011), pp. 1935-1943
95
F.H. Zhou, B.K. Foster, X.F. Zhou, A.J. Cowin, C.J. Xian
TNF-alpha mediates p38 MAP kinase activation and negatively regulates bone formation at the injured growth plate in rats
J Bone Miner Res, 21 (2006), pp. 1075-1088
96
K. Xiang, L. Cheng, Z. Luo, J. Ren, F. Tian, L. Tang, et al.
Glycyrrhizin Suppresses the Expressions of HMGB1 and Relieves the Severity of Traumatic Pancreatitis in Rats
PLoS One, 9 (2014), p. e115982
CrossRef  

References

Close