Open Access

Types of acute phase reactants and their importance in vaccination (Review)

  • Authors:
    • Rafaat H. Khalil
    • Nabil Al-Humadi
  • View Affiliations

  • Published online on: February 11, 2020     https://doi.org/10.3892/br.2020.1276
  • Pages: 143-152
  • Copyright: © Khalil et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Vaccines are considered to be one of the most cost‑effective life‑saving interventions in human history. The body's inflammatory response to vaccines has both desired effects (immune response), undesired effects [(acute phase reactions (APRs)] and trade‑offs. Trade‑offs are more potent immune responses which may be potentially difficult to separate from potent acute phase reactions. Thus, studying acute phase proteins (APPs) during vaccination may aid our understanding of APRs and homeostatic changes which can result from inflammatory responses. Depending on the severity of the response in humans, these reactions can be classified as major, moderate or minor. In this review, types of APPs and their importance in vaccination will be discussed.

1. Introduction

The association between the strength of the acute phase response and vaccination efficacy is of key importance to human and veterinary medicine. Proteins which are expressed in the acute phase are potential biomarkers for the diagnosis of inflammatory disease, for example, acute phase proteins (APPs) are indicators of successful organ transplantation and can be used to predict the ameliorative effect of cancer therapy (1,2). APPs are primarily synthesized in hepatocytes. The acute phase response is a spontaneous reaction triggered by disrupted homeostasis resulting from environmental disturbances (3). Acute phase reactions (APRs) usually stabilize quickly, after recovering from a disruption to homeostasis within a few days to weeks; however, APPs expression levels often remain elevated in long lasting infection and chronic disease states, such as cancer (4-6).

Classification of acute phase reactions is dependent on the change in APP concentration: A 10-100-fold elevation is considered major; a 2-10-fold elevation is considered moderate; and a less than 2-fold elevation is considered minor (7). The APPs elevated in a major APR include C-reactive protein (CRP) and serum amyloid (SA); the APPs elevated in a moderate APR include α1-acid glycoprotein (AGP); and the APPs elevated in a minor APR include fibrinogen, haptoglobin (Hp) and ceruloplasmin (Cp) (8).

In response to infection, the liver synthesizes a large quantity of APPs (8). There are 8 proteins which are overexpressed in APRs denoted as ‘positive’ APPs, including Hp, SA, fibrinogen, Cp, AGP, α-1 antitrypsin (AAT), lactoferrin (Lf) and CRP. Similarly, there are a number of ‘negative’ APPs the expression levels of which are reduced, including albumin, transferrin and transthyretin (8).

The APP is elicited by cytokines, including those functioning as positive and negative growth factors and cytokines with proinflammatory or anti-inflammatory activity. Positive or negative growth factor cytokines involved include: Interleukin (IL)-2; IL-3; IL-4; IL-7; IL-10; IL-11; IL-12; and granulocyte-macrophage colony stimulating factor (9). Proinflammatory cytokines involved include tumour necrosis factor (TNF)-α/β; IL-1α/β; IL-6; IFN-α/γ; IL-8; and macrophage inhibitory protein-1(6). Cytokines involved in the anti-inflammatory response include: IL-1 receptor antagonists; soluble IL-1 receptors; IL-1 binding protein; and TNF-α binding protein. Table I shows acute phase reactants associated, inflammatory cytokines and references.

Table I

Acute phase reactants, associated inflammatory cytokines and references.

Table I

Acute phase reactants, associated inflammatory cytokines and references.

A, Positive acute phase reactants
Author, yearReactantAssociated cytokines(Refs.)
Sharpe-Timms et al, 2010HaptoglobinIL1β, IL-6, IL-8, TNF-α,(9)
He et al, 2006Serum amyloidIL1β, IL-6, IL-8, IL-12, IL-23(143)
Lu et al, 2015FibrinogenIL-6, IL-6, TNF-α, IL-1β, IL-8(144)
Lazzaro et al, 2014CeruloplasminIL-1β, TNF-α, IFN-l(145)
Martinez Cordero et al, 2008α-1 acid glycoproteinTNF-α, IL-1, IL-6(146)
de Serres and Blanco, 2014α-1 antitrypsinTNF-α, IL-6, IL-1β, IL-8, IL-32(147)
Haversen et al, 2002LactoferrinTNF-α, IL-6, IL-1β, IL-8(148)
Du Clos, 2000C-reactive proteinIL-6, IL-1a, IL-1β, TNF-α(149)
B, Negative acute phase reactants
Author, yearReactantAssociated cytokines(Refs.)
Spadaro et al, 2014AlbuminIL-6, TNF-α(150)
Feelders et al, 1998TransferrinIL-6, TNF-α, IL1(151)
Bartalena et al, 1992TransthyretinIL-6, IL-1, TNF-α(152)

[i] IL, interleukin; TNF, tumor necrosis factor.

2. Inflammation and APPs

The acute inflammatory reaction is an essential immune response required for vaccinations to initiate temporary simulated immunity. The innate response is the first branch of the immune system stimulated by invading pathogens, should they cross the body's anatomical and chemical barriers. The innate immune system is also activated by APP synthesis as these molecules mediate inflammation.

After vaccination, pro-inflammatory cells are activated and produce cytokines which diffuse into the extracellular fluid and circulate in the blood. In response, the liver upregulates the synthesis of APPs, preceding the specific immune reaction within a few hours. Monitoring APP expression levels may be an indicator of efficacy of a vaccine in stimulating the innate immune system and may be a useful biomarker in the future development of vaccines.

3. Types of APPs

C-reactive protein (CRP)

CRP is a member of the short evolutionarily conserved pentraxin group of plasma proteins, consisting of 5 identical non-glycosylated peptide subunits which link to form a cyclic pentamer structure (10,11). CRP is produced as a result of pro-inflammatory cytokine signaling primarily mediated by neutrophils and monocytes. CRP concentration is elevated during infection or inflammation as part of the innate immune response and alteration of CRP plasma concentration is dependent on the rate of CRP synthesis and the severity of infection (10). The half-life of CRP in plasma is ~19 h and is cleared by the urinary system (10,12) and CRP stimulates immune cells by binding to Fcγ receptors (FcγR) on leukocytes (monocytes, neutrophils and cells of a myeloid lineage) and increases production of IgG, linking the innate and adaptive immune systems (13). CRP-FcγR binding also facilitates the anti-inflammatory responses (14).

It has been proposed that CRP may have value as a diagnostic marker for active inflammation and infection. IL-6, IL-1 and TNF-α trigger hepatocyte mediated synthesis of CRP in response to active infection or inflammation. Subsequently, CRP binding to bacterial polysaccharides in the presence of calcium activates downstream compliment pathways and ultimately results in phagocytosis (15,16); however, the primary function of CRP is to recognize foreign pathogens and components of damaged cells through binding to phosphocholine (PC), a terminal head group of the lipoteichoic acid which is a component of the cell walls of certain Gram positive bacteria and a component of the mammalian cell membrane. In normal healthy cells, phosphocholine is not exposed however, when cells are damaged or are dying, CRP is able to bind to the PC present on the cell membrane (17). In addition, CRP promotes complement fixation, binding to phagocytized cells and triggers the elimination of cells targeted by the inflammatory response pathway (17-19). There is a direct association between the elevated levels of CRP and the risk of coronary artery and cerebrovascular thrombosis interfering with endothelial nitric oxide (NO) bioavailability, by decreasing endothelial NO synthase expression and increasing the production of reactive oxygen species (20,21).

In the USA, CRP levels in patients tend to be higher in females compared with males in healthy humans (2.7 vs. 1.6 mg/l, respectively) and these levels are exacerbated with age, for example one study reported CRP levels of 1.4 mg/l between the ages of 20-29 vs. 2.7 mg/l in those over 80(22). Meanwhile, ethnicity has little effect on CRP levels (22).

When CRP was studied in non-human species, rabbits served as the primary experimental model due to their susceptibility to infection and the similarity of pathogenesis to what is observed in humans, rabbit CRP behaves more similar to human CRP compared with rat CRP in terms of its dynamic changes during acute phase response (13,23).

Haptoglobin (Hp)

Hp is a glycoprotein synthesized in the liver and present in serum at concentrations of 3-30 µmol/l. Serum levels of Hp increase 3-8-fold in response to inflammation and injury and Hp is eliminated from the plasma in 3-5 days (24,25). Hp binds free hemoglobin (Hb) for detoxification and Hb is highly toxic due to its heme group which mediates the generation of hydroxyl radicals (24). The elimination of Hb and its iron constituent occurs by the formation of a noncovalent Hb-Hp complex which is released into the blood by intravascular hemolysis and subsequently removed by reticuloendothelial receptor-mediated endocytosis and tissue macrophages via interacting with CD163 cell-surface receptor (26). Free hemoglobin is very toxic to the human body due to its ability to bind NO which is a key modulator of vascular tone (27). Hp mediated NO binding prevents NO activity in vascular smooth muscle cells resulting in changes to vasomotor constriction, potentially causing endothelial damage which may contribute to cardiovascular disease, and pulmonary and systemic hypertension (28).

Hp has several functions in the cellular and humoral aspects of the innate and adaptive immune systems (24), including inhibiting prostaglandin production, enhancing the antibody production, leukocyte recruitment and migration, modulation of cytokine release, which are the regulatory mediators secreted by T cells and other immunoactive cells following an injury, infection and tissue repair (29-31). Conditions associated with a reduction or absence of Hp, such as hypohaptoglobinemia or ahaptoglobinemia, result in severe allergies involving the skin, lungs and anaphylaxis (29-31). Hp also suppresses T cell proliferation, including Th2 and Th4 cytokines synthesis (32), and cyclooxygenase and lipoxygenase activity, thus contributing to the regulation of the immune response to potentially damaging inflammation or infection (33).

Serum amyloid A (SAA)

SAA is an APP and a apolipoprotein, which binds to high-density lipoproteins (HDL) in the plasma, During an APR, the fraction of apoA1 in HDL falls while that of SAA rises, becoming the predominant apolipoprotein (apo SAA), exceeding apo A-1 (the major apolipoprotein of native HDL) and reduces the effectiveness of HDL recycling cholesterol to the liver during inflammation (34). SAA displaces apolipoprotein A-1 from HDL, and becomes the predominant circulating HDL3 apolipoprotein mediating reverse cholesterol transport and inhibiting the LDL oxidation. LDL oxidation promotes foam cell formation, and thus, this reduces the risk of atherosclerosis (27). SAA is found at concentrations of 40 ug/ml in healthy males in the UK (34). During acute infection, SAA plasma levels are elevated by up to 1,000-fold (34). SAA has been shown to be an efficient carrier of retinol during an infection, while retinol is important for promotion and maturation of innate immune cells, expression of retinol transporter is reduced during infection, providing insight into the underlying mechanisms involved in the redirection of retinol in response to infection compensating to the markedly reduced levels of serum retinol binding protein transporter following a microbial infection (35). During inflammation, macrophages and monocytes present at the inflammatory site release cytokines, such as IL-6, initiating the induction of SAA synthesis and release of SAA by the liver into the plasma (16).

SAA is a conserved protein which circulates in the plasma and is able to bind to surface ligands of microbial pathogens in a calcium-dependent manner (36). SAA binds to microbial polysaccharides, which are part of the cell wall of gram-negative bacteria, and matrix components via carbohydrate determinant links, including heparin, 6-phosphorylated mannose, 3-sulfated saccharides and the 4,6-cyclic pyruvate acetal of galactose (37,38). SAA binds to apoptotic and necrotic cells facilitating its clearance in vivo (39).

Most species possess two main acute phase apo SAA isoforms of hepatic origin in their serum, SAA1 and SAA2 are APPs with the ability to form amyloid proteins in vivo and SAA1 and SAA2 represent multiple allelic forms which are alternatively expressed by three different genes in humans (40). SAA4 is constitutively expressed across a number of tissues and has been shown to form amyloid when mutated (41). SAA1a is frequently present in amyloid fibrils and is possibly the most amyloidogenic form of SAA1. Although the majority of SAA1 and SAA2 are found bound to HDL, they are only a minor protein component in a healthy state. This classification helps differentiate between regulated acute phase reactants of hepatic origin or constitutive proteins (42).

Fibrinogen

Fibrinogen is an important protein involved in blood clotting, homeostasis, inflammation and tissue repair. Fibrinogen is a 340-kDa soluble glycoprotein found in the blood, and a major component of fibrin which is synthesized in the liver. In healthy adults, fibrinogen plasma levels are ~150-400 mg/dl, and during infection, expression levels of fibrinogen can increase by ≤20-fold (43). At a site of injury, fibrinogen facilitates aggregation of activated platelets through binding to glycoprotein IIb/IIIa cell surface receptor (43), triggering platelet adhesion, and subsequently, thrombin cleaves fibrinogen into fibrin monomers which polymerize to form a clot (44,45) and are stabilized by activated factor XIII (46). The strength of the fibrin clot is influenced by the concentration of fibrinogen (44). A structural scaffold is formed by the fibrin clot onto which leukocyte platelets and fibroblasts adhere and infiltrate the injury site. Extravascular plasma generates thrombin which ultimately leads to deposition of fibrinogen (47), therefore injury, infection and auto-immunity are associated with extravascular fibrinogen (48,49).

Ceruloplasmin (Cp)

Cp is a major copper transport protein present in the plasma and is produced by the hepatic parenchymal cells (50). Human Cp (hCp) is a 132 kDa α2-globulin which can bind up to six copper ions, and serum concentration levels in healthy individuals are ~0.2-0.6 mg/ml, which increases >2-fold during inflammation (51). Overall, ~95% of serum copper is bound to Cp (52). hCp has ferroxidase activity and functions in the mobilization of iron for transport by oxidizing Fe2+ to the less reactive Fe3+ and incorporating Fe3+ into apotransferrin (53). This oxidation prevents the formation of reactive oxygen species and toxic products of iron (54,55). Therefore, Cp has an essential role in iron metabolism and the elimination of free iron (56-58). Cp is an APR and Cp expression levels increase during infection, stress and inflammation (59). Cp also possesses antioxidant properties and functions in the removal of free radicals such as H2O2 during wound healing, collagen formation and the maturation phase which brings about extracellular matrix remodeling and resolution of the granulation of tissue (60,61). However, studies have shown that Cp can also act as a pro-oxidant by promoting the oxidation of low density lipoprotein (62,63).

α1-acid glycoprotein (AGP)

AGP is an APR which stabilizes the biological activity of plasminogen activator inhibitor-1, preventing platelet aggregation (64), and is present in the plasma of healthy humans at concentrations of 0.6-1.2 mg/ml (65). However, these expression levels increase 2-7-fold during an APR (53,66). AGP expression in the liver is induced by activation of IL-1β, IL-6 and TNF-α, and is inhibited by growth hormone (67,68). AGP is considered a natural anti-inflammatory agent with respect to its anti-neutrophilic activity. For example, AGP modulates neutrophil chemotactic migration and superoxide generation in a concentration-dependent manner assisting in the re-establishment of systemic homeostasis following an infection (59,69,70). AGP also inhibits monocyte chemotaxis and cellular leakage caused by histamine and bradykinin levels which are reduced by AGP, and additionally, AGP reduced the synthesis of soluble TNFα receptor leading to an inhibition of the inflammatory process (70). Meanwhile, AGP induces IL-1 receptor antagonism expressed on peripheral blood monocytes (71-73).

α-1 antitrypsin (AAT)

AAT is the most abundant serine protease inhibitor in human blood (65). AAT is present in bodily fluids, including the saliva, tears, urine, bile and circulating blood. AAT consists of a single polypeptide chain made of 394 amino acid residues containing one free cysteine residue and three asparagine-linked carbohydrate side-chains. AAT aids in the elimination of acute inflammation, tissue proteolytic damage by neutrophil elastase in the lungs and inhibits lipopolysaccharides and the release of inflammatory mediators such as TNF-α and IL-1β (65,74,75). AAT is synthesized in the liver but is also produced by other blood cells such as monocytes, macrophages, pulmonary alveolar cells and by intestinal and corneal epithelium (65,74,75). Synthesis of AAT occurs at a rate of 34 mg/kg and the protein clearance rate (half-life) is 3-5 days. As a result, high plasma concentrations of AAT usually range from 0.9-2 mg/ml in healthy individuals (76-80). During an inflammatory response, local AAT synthesis results in the invasion of inflammatory cells followed by an acute rise in AAT expression levels by as much as 11-fold (81).

Lactoferrin (Lf)

Lf is a multifunctional 80 kDa glycoprotein which binds to Fe3+ and an innate immunity factor present in a range of secretory fluids, including mammalian exocrine breast milk, saliva, tears and mucosal secretions (82). Lf is also present in mucosal surfaces and specific leukocyte granules and it can be found in feces following release from fecal leukocytes (83). Abundant antimicrobial peptides and APPs are present in airway surface liquid. Lf is a bacteriostatic protein which chelates iron from Fe3+ (82). Iron is required for bacterial cell division and for the development of biofilms, which are distinct, matrix-encased communities of bacteria, and the biofilm is protects against host defense mechanisms and antibiotics (84). LF chelation of iron occurs at a higher affinity in an acidic medium (85), therefore Lf binding to iron results in the prevention of growth and proliferation of iron dependent bacteria, donating this iron to beneficial bacteria, such as lactic acid bacteria (Lactobacillales) serves as a barrier against colonization of pathogenic bacteria on the intestinal surface thus preventing infection (84). Lf also has several other physiological roles, including stimulation of cell growth and proliferation, differentiation, development of immune competence, antifungal, antibacterial and antiviral activities, antioxidant and anti-inflammatory activity and anti-tumor activity (82,85). Recombinant Lf (TLf) produced in the filamentous fungus, Aspergillus awamori, possesses the same biological activities as human lactoferrin and has been reported to lower mortality in adults with severe blood poisoning (86) and to have anticancer activities (87,88). TLf and human lactoferrin (hLf) have been reported to display changes in immunogenicity and allergenicity in mice (89). Alteration of Lf glycosylation in human milk collected at different time points during lactation resulted in changes in bacterial binding to epithelial cells (89). Thus, hLf from milk and TLf may display different bioactivities. hLF is resistant to gastrointestinal tract digestion and may therefore play an important role in intestinal development during the prenatal period and infancy (90,91). More importantly, Lf promotes maturation of dendritic cells and therefore may function as a natural defense in neonates against bacterial invasion (90) as neonates primarily rely on innate immunity (92).

Albumin

Human serum albumin (HSA) is a major plasma protein synthesized by the liver functioning in the transport of several endogenous ligands, including fatty acids, ions, thyroxine, bilirubin; and exogenous ligands as well as drugs, such as warfarin, diazepam, phenytoin, non-estradiol anti-inflammatory drugs and digoxin (93). Albumin is a member of the family of α-fetoprotein, afamin (also called α-albumin) and vitamin D binding protein (94,95) and these proteins tend to be homologous, that is, all members of the albuminoid superfamily of proteins are suitably capable of ligand binding and transport, as they possess highly conserved intron/exon organization (95). HSA is the key regulator of fluid distribution between somatic regions of the body and body compartment (96). HSA functions to maintain plasma osmotic pressure and its synthesis is regulated by changes in blood osmotic pressure (94,97,98).

HSA is an important biomarker of inflammation in a number of diseases, including cancer, diabetes, rheumatoid arthritis, ischemia and obesity (99-101). Low expression levels of HSA may indicate malnutrition and decrease in hemoglobin levels (99). HSA serves as a valuable cell culture medium and was an additive in the production of pharmaceutical vaccines (102). HSA has also been used for decades in the management of a range of medical and surgical problems, such as for the treatment of acute hypovolemia (surgical blood loss, trauma, or hemorrhage) due to its effect on osmotic pressure providing an oncotic gradient favoring the entry of water from the interstitial space and thus reversing the movement of leaked fluids back into blood vessels (103,104).

Transferrin (Tf)

Tf is a member of the iron-binding glycoprotein family, including lactoferrin (present in intracellular compartments and secretions, such as milk), melanotransferrin (present on melanoma cells) and ovotransferrin (present in egg white of multiple species), and the metal and coordinating anion sites are well conserved in all vertebrates (93). Each of these iron-binding glycoproteins features a single iron-binding site and are monomeric proteins 76-81 kDa in weight. Tf consists of two structurally similar lobes, namely, the C- and the N-lobes. Plasma transferrin provides body tissues with iron, whereas the remaining transferrin are produced locally and transport iron to restricted regions, including the testes and brain, as these two sites are relatively inaccessible to proteins in the general circulation due to the presence of highly specialized barriers (104). Liver hepatocytes are the primary sites of Tf production; however, Tf is also synthesized in other organs to a lesser extent, including the choroid plexus (105). The primary function of Tf is to transport iron from absorption centers in the duodenum and in white blood cell macrophages to all tissues (105). Tf also functions as a constituent of the innate immune system where its levels decrease during inflammation. The Tf receptor is a receptor for the IgA1 class of antibodies (106). Tf serves an important role in the somatic regions where erythropoiesis and active cell division occur (106). Iron-bound Tf is internalized by cells expressing specific Tf receptors by receptor-mediated endocytosis thus contributing to an environment with low free iron capable of inhibiting bacterial growth and multiplication during infection (96).

Transthyretin (TTR)

TTR (previously referred to as prealbumin) is a negative APR synthesized and excreted by the kidneys and gastrointestinal tract with a half-life of 1.9 days and expression levels of TTR decrease significantly during inflammation which promotes an APR (96,107). TTR is also synthesized in the choroid plexus and forms a complex molecule with retinol binding protein allowing retinol and thyroxine transport (107). This process is mediated by pro-inflammatory cytokines, such as IL-6, IL-1a and TNFα (108). It has been reported that TTR functions as a biomarker for predicting poor short-term outcome and disease severity in patients with burn injuries (109,110) or respiratory failure (111), and was strongly correlated with the score of sequential organ failure assessment, and low levels of TTR were associated with an increase in mortality (112,113). Patients with low preoperative TTR levels (<0.2 g/l) are prone to increased risk of postoperative infections and need longer mechanical ventilation after heart surgery (113).

4. Vaccines and APP

Upon vaccination and introduction of antigens into the body, macrophages and dendritic cells are stimulated, producing inflammatory cytokines and triggering APP synthesis in hepatocytes, which function nonspecifically as part of the innate immune system detecting pathogens or vaccination components (114). Vaccines have been shown to cause inflammatory responses which has a direct impact on the maintenance of homeostasis, particularly in the kidney and the liver (115). Disruption of homeostasis during APR can negatively affect the host (3,7,16), for example, plasma levels of a number of microminerals, including iron, may change as they are taken up by hepatocytes and other cell types (7). As the APR increases the products of metabolism will also increase (4). The increase in the plasma proteins during APR affects the concentration of the free form of any drug administered at the time of stress, leading to drug dispersal, and variations in plasma protein concentrations during the acute phase response may result in the plasma levels of available drugs to decrease considerably whereas the total drug concentration (free and bound) will be only slightly affected, and thus the effective dose will be altered due to fluctuations in serum protein levels and decreased serum albumin expression levels (116,117). Individual variation in the expression levels of APRs in response to inflammatory stimuli, such as vaccines, has been previously reported in cattle (118), mice (119) and humans (120). This difference may be due to individual differences in inflammatory genes (120,121). Obesity can alter vaccination responses in mice, wherein the obesity was associated with decreased antibody production and ultimately reduced the efficacy of an influenza vaccine (122). Furthermore, decreased influenza-specific antibody levels and B-cell function in response to vaccination has also been linked to obesity in humans (123). Administration of vaccines containing foreign particles into the body triggers local inflammation due to an immediate change in cytokine production, which triggers the clinical symptoms associated with vaccination. Overall, a reduced response to vaccination is a likely indication of an unregulated inflammatory system which may lead to increased risk of disease as the majority of vaccines protect against infection through induction of a B-cell antibody response (124).

Pro-inflammatory cytokine IL-6 is an important regulator of inflammation and is one of the primary cytokines which stimulate APP synthesis in hepatocytes (125-127). IL-6 functions in several processes during inflammation, including chemokine production, development of B-cells and dendritic cells, secretion and maturation of antibodies, T-cell maturation, as well as linking the innate and adaptive immune responses (128,129). An acute spike of anti-IL-6 was observed 5 days post vaccination in a small study of yellow fever vaccine, indicating its potential role predict vaccine-induced protection (130). Kurtz et al (131) studied the protective mechanisms of IL-6 against a Francisella tularensis live vaccine administered intradermally or intranasally to IL-6 knockout (KO) and wild type mice. The group reported a decline in SSA and Hp in IL-6 KO mice compared with wild type controls, indicating the importance of IL-6 in protection against infection.

In humans, the efficacy of the influenza vaccination has been invaluable. In one study, CRP expression levels increased by 36%, further increasing to 40% above baseline at day 3, before returning to the normal levels by day 7 following vaccination (132). In a study of patients >65 years old, CRP plasma concentrations spiked at 2 days post vaccination with influenza and pneumococcal vaccines as well as a combination of the two (133). Another study investigating the influenza vaccine showed that CRP expression levels post-vaccination were dependent on the dose of vaccine (134). Table II show acute phase reactants associated with vaccines.

Table II

Acute phase reactants associated with vaccines.

Table II

Acute phase reactants associated with vaccines.

Author, yearVaccineAcute phase reactant(Refs.)
Louagie et al, 1993Hepatitis BHaptoglobin(153)
Borthwick et al, 2018Human immunodeficiency virus core DNASerum amyloid(154)
Creech et al, 2017S. aureus capsular polysaccharidesFibrinogen(155)
Blom et al, 1979Typhoid. AB.-choleraCeruloplasmin, α-1 glycoprotein(156)
Naylor et al, 2015Rotavirus and poliovirusα-1 antitrypsin(157)
Hwang et al, 2005Bacillus Calmette-GuerinLactoferrin(158)
Carty et al, 2006InfluenzaC-reactive protein, transthyretin(159)
Patel and Shah, 2015H1N1 influenzaAlbumin(160)
Bos et al, 2016Neisseria gonorrhoeaeTransferrin(161)

5. Effect of adjuvant systems on APP

Adjuvants are compounds joined with other constituents in vaccine formulations (135) where multiple immunostimulants are combined with adjuvant systems to enhance antibody production and adaptive immune responses to vaccination (136). Adjuvant components regulate and improve the specificity of the immune system, thus aiming to optimize the immune response to vaccination (136). To date, there have been at least six novel adjuvants approved for use as vaccine components in the last 20 years (135,136). Adjuvant systems can be useful when strong T cell responses are required to protect against complex pathogens formed of several different types of antigens, chronic infections, or in immunocompromised populations such as the elderly or neonates (137,138). A previous study on the effects of adjuvants in rabbits indicated that CRP and fibrinogen expression levels were increased following the administration of adjuvant systems AS01, AS03, AS15 and DTPw (139). The increase in CRP levels increased 9-26-fold after injection of AS01, AS03, or AS15. Comparable effects have also been shown in humans (140). However, the use of other adjuvants such as aluminum phosphate and aluminum hydroxide did not affect CRP expression levels in the elderly or in the young following a diphtheria-tetanus-poliomyelitis-typhoid vaccine (141), influenza vaccine (142) or yellow fever vaccine (120,130). Overall, these studies indicate the efficacy of these adjuvant systems in stimulating the adaptive immune response.

6. Conclusion

APRs a group of 11 key proteins, eight of which are positive and three are negative proteins. During infection, the liver hepatocytes respond by producing a large number of APRs. Upon vaccination, an acute-phase protein reaction may develop which is a part of the innate immune system. Vaccination has also been demonstrated to cause an inflammatory response, which results in homeostatic changes including changes in renal and liver functions. Changes in APR expression levels in response to an inflammatory stimulus may vary across individuals due to genetic, nutritional, age, or other health-associated factors. The health and nutritional status of the individual may affect the response to a given vaccine due to the inflammatory condition. For example, IL-6 (an inflammatory key mediator and regulator of the majority of APRs) functions in the acute-phase protein response in the innate and adaptive branches of the immune system. CRP, as a prototypical APP, is synthesized in response to pro-inflammatory cytokine signals and CRP expression levels increase in innate immune responses to infection or to injury. Further investigations into the relationships between vaccination and the effect on CRP and other APRs may help clarify the mechanisms by which beneficial and harmful inflammatory responses affect health and influence the immune response.

Acknowledgements

Not applicable.

Funding

No funding was received.

Availability of data and materials

Not applicable.

Authors' contributions

RHK and NAH wrote the manuscript, edited and critically reviewed the manuscript. Both authors read and approved the final version of the manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

This article reflects the views of the authors and should not be construed to represent the FDA's views or policies. Dr Humadi's contributions are an informal communication and represent his own best judgement. These comments do not bind or obligate FDA.

References

1 

Deans C and Wigmore SJ: Systemic inflammation, cachexia and prognosis in patients with cancer. Curr Opin Clin Nutr Metab Care. 8:265–269. 2005.PubMed/NCBI View Article : Google Scholar

2 

Ridker PM: Inflammatory biomarkers and risks of myocardial infarction, stroke, diabetes, and total mortality: Implications for longevity. Nutr Rev. 65 (Suppl):S253–S259. 2007.PubMed/NCBI View Article : Google Scholar

3 

Cray C, Zaias J and Altman NH: Acute phase response in animals: A review. Comp Med. 59:517–526. 2009.PubMed/NCBI

4 

Boosalis MG, Snowdon DA, Tully CL and Gross MD: Acute phase response and plasma carotenoid concentrations in older women: Findings from the nun study. Nutrition. 12:475–478. 1996.PubMed/NCBI View Article : Google Scholar

5 

Mackiewicz A: Acute phase proteins and transformed cells. Int Rev Cytol. 170:225–300. 1997.PubMed/NCBI View Article : Google Scholar

6 

Koj A: Termination of acute-phase response: Role of some cytokines and anti-inflammatory drugs. Gen Pharmacol. 31:9–18. 1998.PubMed/NCBI View Article : Google Scholar

7 

Ceron JJ, Eckersall PD and Martynez-Subiela S: Acute phase proteins in dogs and cats: Current knowledge and future perspectives. Vet Clin Pathol. 34:85–99. 2005.PubMed/NCBI View Article : Google Scholar

8 

Gómez-Laguna J, Salguero FJ, Pallarés FJ, Rodríguez-Gómez IM, Barranco I and Carrasco L: Acute phase proteins as biomarkers in animal health and welfare. In: Acute Phase Proteins As Early Non-Specific Biomarkers of Human and Veterinary Diseases. Veas F (ed). InTech, pp259-298, 2011.

9 

Sharpe-Timms KL, Nabli H, Zimmer RL, Birt JA and Davis JW: Inflammatory cytokines differentially up-regulate human endometrial haptoglobin production in women with endometriosis. Hum Reprod. 25:1241–1250. 2010.PubMed/NCBI View Article : Google Scholar

10 

Suk HJ, Ridker PM, Cook NR and Zee RY: Relation of polymorphism within the C-reactive protein gene and plasma CRP levels. Atherosclerosis. 178:139–145. 2005.PubMed/NCBI View Article : Google Scholar

11 

Macleod CM and Avery OT: The occurrence during acute infections of a protein not normally present in the blood: II. Isolation and properties of the reactive protein. J Exp Med. 73:183–190. 1941.PubMed/NCBI View Article : Google Scholar

12 

Ballou SP and Kushner I: C-reactive protein and the acute phase response. Adv Intern Med. 37:313–336. 1992.PubMed/NCBI

13 

Anderson HC and McCarty M: The occurrence in the rabbit of an acute phase protein analogous to human C reactive protein. J Exp Med. 93:25–36. 1951.PubMed/NCBI View Article : Google Scholar

14 

Mold C, Rodriguez W, Rodic-Polic B and Du Clos TW: C-reactive protein mediates protection from lipopolysaccharide through interactions with Fc gamma R. J Immunol. 169:7019–7025. 2002.PubMed/NCBI View Article : Google Scholar

15 

Volanakis JE and Kaplan MH: Specificity of C-reactive protein for choline phosphate residues of pneumococcal C-polysaccharide. Proc Soc Exp Biol Med. 136:612–614. 1971.PubMed/NCBI View Article : Google Scholar

16 

Gabay C and Kushner I: Acute-phase proteins and other systemic responses to inflammation. N Engl J Med. 340:448–454. 1999.PubMed/NCBI View Article : Google Scholar

17 

Volanakis JE: Acute phase proteins in rheumatic disease. In: Arthritis and allied conditions: A textbook of rheumatology. Koopman WJ (ed). 13 edition. Williams & Wilkins, Baltimore, MD, pp505-514, 1997.

18 

Rao AK, Chouhan V, Chen X, Sun L and Boden G: Activation of the tissue factor pathway of blood coagulation during prolonged hyperglycemia in young healthy men. Diabetes. 48:1156–1161. 1999.PubMed/NCBI View Article : Google Scholar

19 

Stegenga ME, van der Crabben SN, Levi M, de Vos AF, Tanck MW, Sauerwein HP and van der Poll T: Hyperglycemia stimulates coagulation, whereas hyperinsulinemia impairs fibrinolysis in healthy humans. Diabetes. 55:1807–1812. 2006.PubMed/NCBI View Article : Google Scholar

20 

Pepys MB and Hirschfield GM: C-reactive protein: A critical update. J Clin Invest. 111:1805–1812. 2003.PubMed/NCBI View Article : Google Scholar

21 

Falsey AR, Walsh EE, Francis CW, Looney RJ, Kolassa JE, Hall WJ and Abraham GN: Response of C-reactive protein and serum amyloid A to influenza A infection in older adults. J Infect Dis. 183:995–999. 2001.PubMed/NCBI View Article : Google Scholar

22 

Woloshin S and Schwartz LM: Distribution of C-reactive protein values in the United States. N Engl J Med. 352:1611–1613. 2005.PubMed/NCBI View Article : Google Scholar

23 

Yen-Watson B and Kushner I: Rabbit CRP response to endotoxin administration: Dose-response relationship and kinetics. Proc Soc Exp Biol Med. 146:1132–1136. 1974.PubMed/NCBI View Article : Google Scholar

24 

Langlois MR and Delanghe JR: Biological and clinical significance of haptoglobin polymorphism in humans. Clin Chem. 42:1589–1600. 1996.PubMed/NCBI

25 

Bowman BH and Kurosky A: Haptoglobin: The evolutionary product of duplication, unequal crossing over, and point mutation. Adv Hum Genet. 12:189–261, 453-184. 1982.PubMed/NCBI View Article : Google Scholar

26 

Alayash AI: Redox biology of blood. Antioxid Redox Signal. 6:941–943. 2004.PubMed/NCBI View Article : Google Scholar

27 

Eklund KK, Niemi K and Kovanen PT: Immune functions of serum amyloid A. Crit Rev Immunol. 32:335–348. 2012.PubMed/NCBI View Article : Google Scholar

28 

Minneci PC, Deans KJ, Zhi H, Yuen PS, Star RA, Banks SM, Schechter AN, Natanson C, Gladwin MT and Solomon SB: Hemolysis-associated endothelial dysfunction mediated by accelerated NO inactivation by decompartmentalized oxyhemoglobin. J Clin Invest. 115:3409–3417. 2005.PubMed/NCBI View Article : Google Scholar

29 

Gilstad CW: Anaphylactic transfusion reactions. Curr Opin Hematol. 10:419–423. 2003.PubMed/NCBI View Article : Google Scholar

30 

Larsen K, Macleod D, Nihlberg K, Gurcan E, Bjermer L, Marko-Varga G and Westergren-Thorsson G: Specific haptoglobin expression in bronchoalveolar lavage during differentiation of circulating fibroblast progenitor cells in mild asthma. J Proteome Res. 5:1479–1483. 2006.PubMed/NCBI View Article : Google Scholar

31 

Shimada E, Tadokoro K, Watanabe Y, Ikeda K, Niihara H, Maeda I, Isa K, Moriya S, Ashida T, Mitsunaga S, et al: Anaphylactic transfusion reactions in haptoglobin-deficient patients with IgE and IgG haptoglobin antibodies. Transfusion. 42:766–773. 2002.PubMed/NCBI View Article : Google Scholar

32 

Arredouani M, Matthijs P, Van Hoeyveld E, Kasran A, Baumann H, Ceuppens JL and Stevens E: Haptoglobin directly affects T cells and suppresses T helper cell type 2 cytokine release. Immunology. 108:144–151. 2003.PubMed/NCBI View Article : Google Scholar

33 

Saeed SA, Ahmad N and Ahmed S: Dual inhibition of cyclooxygenase and lipoxygenase by human haptoglobin: Its polymorphism and relation to hemoglobin binding. Biochem Biophys Res Commun. 353:915–920. 2007.PubMed/NCBI View Article : Google Scholar

34 

Pepys MB, Dash AC, Markham RE, Thomas HC, Williams BD and Petrie A: Comparative clinical study of protein SAP (amyloid P component) and C-reactive protein in serum. Clin Exp Immunol. 32:119–124. 1978.PubMed/NCBI

35 

Derebe MG, Zlatkov CM, Gattu S, Ruhn KA, Vaishnava S, Diehl GE, MacMillan JB, Williams NS and Hooper LV: Serum amyloid A is a retinol binding protein that transports retinol during bacterial infection. Elife. 3(e03206)2014.PubMed/NCBI View Article : Google Scholar

36 

Bottazzi B, Doni A, Garlanda C and Mantovani A: An integrated view of humoral innate immunity: Pentraxins as a paradigm. Annu Rev Immunol. 28:157–183. 2010.PubMed/NCBI View Article : Google Scholar

37 

Hind CR, Collins PM, Renn D, Cook RB, Caspi D, Baltz ML and Pepys MB: Binding specificity of serum amyloid P component for the pyruvate acetal of galactose. J Exp Med. 159:1058–1069. 1984.PubMed/NCBI View Article : Google Scholar

38 

Loveless RW, Floyd-O'Sullivan G, Raynes JG, Yuen CT and Feizi T: Human serum amyloid P is a multispecific adhesive protein whose ligands include 6-phosphorylated mannose and the 3-sulphated saccharides galactose, N-acetylgalactosamine and glucuronic acid. EMBO J. 11:813–819. 1992.PubMed/NCBI View Article : Google Scholar

39 

Familian A, Zwart B, Huisman HG, Rensink I, Roem D, Hordijk PL, Aarden LA and Hack CE: Chromatin-independent binding of serum amyloid P component to apoptotic cells. J Immunol. 167:647–654. 2001.PubMed/NCBI View Article : Google Scholar

40 

Husby G, Marhaug G, Dowton B, Sletten K and Sipe JD: Serum amyloid A (SAA): Biochemistry, genetics and the pathogenesis of AA amyloidosis. Int J Exu Clin Invest. 1:119–137. 1994. View Article : Google Scholar

41 

Wang S, Murphy CL, Kestler D, Macy SD, Williams TK, Weiss DT and Solomon A: SAA4-related AA amyloidosis. In: XIth International Symposium on Amyloidosis. CRC Press, Boca Raton, FL, 2008.

42 

Eriksen N and Benditt EP: Isolation and characterization of the amyloid-related apoprotein (SAA) from human high density lipoprotein. Proc Natl Acad Sci USA. 77:6860–6864. 1980.PubMed/NCBI View Article : Google Scholar

43 

Levy JH, Szlam F, Tanaka KA and Sniecienski RM: Fibrinogen and hemostasis: A primary hemostatic target for the management of acquired bleeding. Anesth Analg. 114:261–274. 2012.PubMed/NCBI View Article : Google Scholar

44 

Jennewein C, Tran N, Paulus P, Ellinghaus P, Eble JA and Zacharowski K: Novel aspects of fibrin(ogen) fragments during inflammation. Mol Med. 17:568–573. 2011.PubMed/NCBI View Article : Google Scholar

45 

Mosesson MW: Fibrinogen and fibrin structure and functions. J Thromb Haemost. 3:1894–1904. 2005.PubMed/NCBI View Article : Google Scholar

46 

Meyer MA, Ostrowski SR, Windelov NA and Johansson PI: Fibrinogen concentrates for bleeding trauma patients: What is the evidence? Vox Sang. 101:185–190. 2011.PubMed/NCBI View Article : Google Scholar

47 

Furie B and Furie BC: The molecular basis of blood coagulation. Cell. 53:505–518. 1988.PubMed/NCBI View Article : Google Scholar

48 

Colvin RB, Johnson RA, Mihm MC Jr and Dvorak HF: Role of the clotting system in cell-mediated hypersensitivity. I. Fibrin deposition in delayed skin reactions in man. J Exp Med. 138:686–698. 1973.PubMed/NCBI View Article : Google Scholar

49 

Labarrere CA, Nelson DR and Faulk WP: Myocardial fibrin deposits in the first month after transplantation predict subsequent coronary artery disease and graft failure in cardiac allograft recipients. Am J Med. 105:207–213. 1998.PubMed/NCBI View Article : Google Scholar

50 

Ryden L and Eaker D: Identification of the thiol groups in human ceruloplasmin. Eur J Biochem. 132:241–247. 1983.PubMed/NCBI View Article : Google Scholar

51 

Jamieson GA: Studies on Glycoproteins. I. The carbohydrate portion of human ceruloplasmin. J Biol Chem. 240:2019–2027. 1965.PubMed/NCBI

52 

Harris ZL, Klomp LW and Gitlin JD: Aceruloplasminemia: An inherited neurodegenerative disease with impairment of iron homeostasis. Am J Clin Nutr. 67 (5 Suppl):972S–977S. 1998.PubMed/NCBI View Article : Google Scholar

53 

Colombo S, Buclin T, Decosterd LA, Telenti A, Furrer H, Lee BL, Biollaz J and Eap CB: Swiss HIV Cohort Study: Orosomucoid (alpha1-acid glycoprotein) plasma concentration and genetic variants: Effects on human immunodeficiency virus protease inhibitor clearance and cellular accumulation. Clin Pharmacol Ther. 80:307–318. 2006.PubMed/NCBI View Article : Google Scholar

54 

Osaki S, Johnson DA and Frieden E: The possible significance of the ferrous oxidase activity of ceruloplasmin in normal human serum. J Biol Chem. 241:2746–2751. 1966.PubMed/NCBI

55 

Osaki S, Johnson DA and Frieden E: The mobilization of iron from the perfused mammalian liver by a serum copper enzyme, ferroxidase I. J Biol Chem. 246:3018–3023. 1971.PubMed/NCBI

56 

Attieh ZK, Mukhopadhyay CK, Seshadri V, Tripoulas NA and Fox PL: Ceruloplasmin ferroxidase activity stimulates cellular iron uptake by a trivalent cation-specific transport mechanism. J Biol Chem. 274:1116–1123. 1999.PubMed/NCBI View Article : Google Scholar

57 

Harris ZL, Takahashi Y, Miyajima H, Serizawa M, MacGillivray RT and Gitlin JD: Aceruloplasminemia: Molecular characterization of this disorder of iron metabolism. Proc Natl Acad Sci USA. 92:2539–2543. 1995.PubMed/NCBI View Article : Google Scholar

58 

Lahey ME, Gubler CJ, Chase MS, Cartwright GE and Wintrobe MM: Studies on copper metabolism. II. Hematologic manifestations of copper deficiency in swine. Blood. 7:1053–1074. 1952.PubMed/NCBI

59 

Hochepied T, Berger FG, Baumann H and Libert C: Alpha(1)-acid glycoprotein: An acute phase protein with inflammatory and immunomodulating properties. Cytokine Growth Factor Rev. 14:25–34. 2003.PubMed/NCBI View Article : Google Scholar

60 

Giurgea N, Constantinescu MI, Stanciu R, Suciu S and Muresan A: Ceruloplasmin-acute-phase reactant or endogenous antioxidant? The case of cardiovascular disease. Med Sci Monit. 11:RA48–RA51. 2005.PubMed/NCBI

61 

Gutteridge JM: Caeruloplasmin: A plasma protein, enzyme, and antioxidant. Ann Clin Biochem. 15:293–296. 1978.PubMed/NCBI View Article : Google Scholar

62 

Mukhopadhyay CK, Ehrenwald E and Fox PL: Ceruloplasmin enhances smooth muscle cell- and endothelial cell-mediated low density lipoprotein oxidation by a superoxide-dependent mechanism. J Biol Chem. 271:14773–14778. 1996.PubMed/NCBI View Article : Google Scholar

63 

Mukhopadhyay CK, Mazumder B, Lindley PF and Fox PL: Identification of the prooxidant site of human ceruloplasmin: A model for oxidative damage by copper bound to protein surfaces. Proc Natl Acad Sci USA. 94:11546–11551. 1997.PubMed/NCBI View Article : Google Scholar

64 

Smolarczyk K, Gils A, Boncela J, Declerck PJ and Cierniewski CS: Function-stabilizing mechanism of plasminogen activator inhibitor type 1 induced upon binding to alpha1-acid glycoprotein. Biochemistry. 44:12384–12390. 2005.PubMed/NCBI View Article : Google Scholar

65 

Geboes K, Ray MB, Rutgeerts P, Callea F, Desmet VJ and Vantrappen G: Morphological identification of alpha-I-antitrypsin in the human small intestine. Histopathology. 6:55–60. 1982.PubMed/NCBI View Article : Google Scholar

66 

Kremer JM, Wilting J and Janssen LH: Drug binding to human alpha-1-acid glycoprotein in health and disease. Pharmacol Rev. 40:1–47. 1988.PubMed/NCBI

67 

Barraud B, Balavoine S, Feldmann G and Lardeux B: Effects of insulin, dexamethasone and cytokines on alpha 1-acid glycoprotein gene expression in primary cultures of normal rat hepatocytes. Inflammation. 20:191–202. 1996.PubMed/NCBI View Article : Google Scholar

68 

Mejdoubi N, Henriques C, Bui E, Durand G, Lardeux B and Porquet D: Growth hormone inhibits rat liver alpha-1-acid glycoprotein gene expression in vivo and in vitro. Hepatology. 29:186–194. 1999.PubMed/NCBI View Article : Google Scholar

69 

Costello MJ, Gewurz H and Siegel JN: Inhibition of neutrophil activation by alpha1-acid glycoprotein. Clin Exp Immunol. 55:465–472. 1984.PubMed/NCBI

70 

Laine E, Couderc R, Roch-Arveiller M, Vasson MP, Giroud JP and Raichvarg D: Modulation of human polymorphonuclear neutrophil functions by alpha 1-acid glycoprotein. Inflammation. 14:1–9. 1990.PubMed/NCBI View Article : Google Scholar

71 

Tilg H, Vannier E, Vachino G, Dinarello CA and Mier JW: Antiinflammatory properties of hepatic acute phase proteins: Preferential induction of interleukin 1 (IL-1) receptor antagonist over IL-1 beta synthesis by human peripheral blood mononuclear cells. J Exp Med. 178:1629–1636. 1993.PubMed/NCBI View Article : Google Scholar

72 

Samak R, Edelstein R and Israel L: Immunosuppressive effect of acute-phase reactant proteins in vitro and its relevance to cancer. Cancer Immunol Immunother. 13:38–43. 1982.PubMed/NCBI View Article : Google Scholar

73 

Muchitsch EM, Teschner W, Linnau Y and Pichler L: In vivo effect of alpha 1-acid glycoprotein on experimentally enhanced capillary permeability in guinea-pig skin. Arch Int Pharmacodyn Ther. 331:313–321. 1996.PubMed/NCBI

74 

Perlmutter DH, Kay RM, Cole FS, Rossing TH, Van Thiel D and Colten HR: The cellular defect in alpha 1-proteinase inhibitor (alpha 1-PI) deficiency is expressed in human monocytes and in Xenopus oocytes injected with human liver mRNA. Proc Natl Acad Sci USA. 82:6918–6921. 1985.PubMed/NCBI View Article : Google Scholar

75 

Ray MB, Geboes K, Callea F and Desmet VJ: Alpha-1-antitrypsin immunoreactivity in gastric carcinoid. Histopathology. 6:289–297. 1982.PubMed/NCBI View Article : Google Scholar

76 

Berman MB, Barber JC, Talamo RC and Langley CE: Corneal ulceration and the serum antiproteases. I. Alpha 1-antitrypsin. Invest Ophthalmol. 12:759–770. 1973.PubMed/NCBI

77 

Chowanadisai W and Lonnerdal B: Alpha(1)-antitrypsin and antichymotrypsin in human milk: Origin, concentrations, and stability. Am J Clin Nutr. 76:828–833. 2002.PubMed/NCBI View Article : Google Scholar

78 

Huang CM: Comparative proteomic analysis of human whole saliva. Arch Oral Biol. 49:951–962. 2004.PubMed/NCBI View Article : Google Scholar

79 

Janciauskiene S, Toth E, Sahlin S and Eriksson S: Immunochemical and functional properties of biliary alpha-1-antitrypsin. Scand J Clin Lab Invest. 56:597–608. 1996.PubMed/NCBI View Article : Google Scholar

80 

Poortmans J and Jeanloz RW: Quantitative immunological determination of 12 plasma proteins excreted in human urine collected before and after exercise. J Clin Invest. 47:386–393. 1968.PubMed/NCBI View Article : Google Scholar

81 

Boskovic G and Twining SS: Retinol and retinaldehyde specifically increase alpha1-proteinase inhibitor in the human cornea. Biochem J. 322:751–756. 1997.PubMed/NCBI View Article : Google Scholar

82 

Lonnerdal B and Iyer S: Lactoferrin: Molecular structure and biological function. Annu Rev Nutr. 15:93–110. 1995.PubMed/NCBI View Article : Google Scholar

83 

Garcia-Montoya IA, Cendon TS, Arevalo-Gallegos S and Rascon-Cruz Q: Lactoferrin a multiple bioactive protein: An overview. Biochim Biophys Acta. 1820:226–236. 2012.PubMed/NCBI View Article : Google Scholar

84 

Singh PK, Parsek MR, Greenberg EP and Welsh MJ: A component of innate immunity prevents bacterial biofilm development. Nature. 417:552–555. 2002.PubMed/NCBI View Article : Google Scholar

85 

Vogel HJ: Lactoferrin, a bird's eye view. Biochem Cell Biol. 90:233–244. 2012.PubMed/NCBI View Article : Google Scholar

86 

Guntupalli K, Dean N, Morris PE, Bandi V, Margolis B, Rivers E, Levy M, Lodato RF, Ismail PM, Reese A, et al: A phase 2 randomized, double-blind, placebo-controlled study of the safety and efficacy of talactoferrin in patients with severe sepsis. Crit Care Med. 41:706–716. 2013.PubMed/NCBI View Article : Google Scholar

87 

Digumarti R, Wang Y, Raman G, Doval DC, Advani SH, Julka PK, Parikh PM, Patil S, Nag S, Madhavan J, et al: A randomized, double-blind, placebo-controlled, phase II study of oral talactoferrin in combination with carboplatin and paclitaxel in previously untreated locally advanced or metastatic non-small cell lung cancer. J Thorac Oncol. 6:1098–1103. 2011.PubMed/NCBI View Article : Google Scholar

88 

Parikh PM, Vaid A, Advani SH, Digumarti R, Madhavan J, Nag S, Bapna A, Sekhon JS, Patil S, Ismail PM, et al: Randomized, double-blind, placebo-controlled phase II study of single-agent oral talactoferrin in patients with locally advanced or metastatic non-small-cell lung cancer that progressed after chemotherapy. J Clin Oncol. 29:4129–4136. 2011.PubMed/NCBI View Article : Google Scholar

89 

Almond RJ, Flanagan BF, Antonopoulos A, Haslam SM, Dell A, Kimber I and Dearman RJ: Differential immunogenicity and allergenicity of native and recombinant human lactoferrins: Role of glycosylation. Eur J Immunol. 43:170–181. 2013.PubMed/NCBI View Article : Google Scholar

90 

Lonnerdal B, Jiang R and Du X: Bovine lactoferrin can be taken up by the human intestinal lactoferrin receptor and exert bioactivities. J Pediatr Gastroenterol Nutr. 53:606–614. 2011.PubMed/NCBI View Article : Google Scholar

91 

Davidson LA and Lonnerdal B: Persistence of human milk proteins in the breast-fed infant. Acta Paediatr Scand. 76:733–740. 1987.PubMed/NCBI View Article : Google Scholar

92 

Wynn JL and Levy O: Role of innate host defenses in susceptibility to early-onset neonatal sepsis. Clin Perinatol. 37:307–337. 2010.PubMed/NCBI View Article : Google Scholar

93 

Lambert LA: Molecular evolution of the transferrin family and associated receptors. Biochim Biophys Acta. 1820:244–255. 2012.PubMed/NCBI View Article : Google Scholar

94 

Peters T Jr: All about albumin. Biochemistry, Genetics and Medical Applications. Academic Press, Inc., San Diego, CA, 1996.

95 

Fasano M, Fanali G, Leboffe L and Ascenzi P: Heme binding to albuminoid proteins is the result of recent evolution. IUBMB Life. 59:436–440. 2007.PubMed/NCBI View Article : Google Scholar

96 

Aisen P: Transferrin receptor 1. Int J Biochem Cell Biol. 36:2137–2143. 2004.PubMed/NCBI View Article : Google Scholar

97 

Evans TW: Review article: Albumin as a drug-biological effects of albumin unrelated to oncotic pressure. Aliment Pharmacol Ther. 16 (Suppl 5):S6–S11. 2002.PubMed/NCBI View Article : Google Scholar

98 

Mendez CM, McClain CJ and Marsano LS: Albumin therapy in clinical practice. Nutr Clin Pract. 20:314–320. 2005.PubMed/NCBI View Article : Google Scholar

99 

Gupta D and Lis CG: Pretreatment serum albumin as a predictor of cancer survival: A systematic review of the epidemiological literature. Nutr J. 9(69)2010.PubMed/NCBI View Article : Google Scholar

100 

Koga M and Kasayama S: Clinical impact of glycated albumin as another glycemic control marker. Endocr J. 57:751–762. 2010.PubMed/NCBI View Article : Google Scholar

101 

Sbarouni E, Georgiadou P and Voudris V: Ischemia modified albumin changes-review and clinical implications. Clin Chem Lab Med. 49:177–184. 2011.PubMed/NCBI View Article : Google Scholar

102 

Marth E and Kleinhappl B: Albumin is a necessary stabilizer of TBE-vaccine to avoid fever in children after vaccination. Vaccine. 20:532–537. 2001.PubMed/NCBI View Article : Google Scholar

103 

Haynes GR, Navickis RJ and Wilkes MM: Albumin administration-what is the evidence of clinical benefit? A systematic review of randomized controlled trials. Eur J Anaesthesiol. 20:771–793. 2003.PubMed/NCBI View Article : Google Scholar

104 

Alderson P, Bunn F, Lefebvre C, Li WP, Li L, Roberts I and Schierhout G: Albumin Reviewers: Human albumin solution for resuscitation and volume expansion in critically ill patients. Cochrane Database Syst Rev. 2004(CD001208)2004.PubMed/NCBI View Article : Google Scholar

105 

Idzerda RL, Huebers H, Finch CA and McKnight GS: Rat transferrin gene expression: Tissue-specific regulation by iron deficiency. Proc Natl Acad Sci USA. 83:3723–3727. 1986.PubMed/NCBI View Article : Google Scholar

106 

Macedo MF and de Sousa M: Transferrin and the transferrin receptor: Of magic bullets and other concerns. Inflamm Allergy Drug Targets. 7:41–52. 2008.PubMed/NCBI View Article : Google Scholar

107 

Ingenbleek Y and Young V: Transthyretin (prealbumin) in health and disease: Nutritional implications. Annu Rev Nutr. 14:495–533. 1994.PubMed/NCBI View Article : Google Scholar

108 

Ballmer PE: Causes and mechanisms of hypoalbuminaemia. Clin Nutr. 20:271–273. 2001.PubMed/NCBI View Article : Google Scholar

109 

Cynober L, Prugnaud O, Lioret N, Duchemin C, Saizy R and Giboudeau J: Serum transthyretin levels in patients with burn injury. Surgery. 109:640–644. 1991.PubMed/NCBI

110 

Yang HT, Yim H, Cho YS, Kim D, Hur J, Kim JH, Lee JW, Lee YK, Lee J, Han SW and Chun W: Serum transthyretin level is associated with clinical severity rather than nutrition status in massively burned patients. JPEN J Parenter Enteral Nutr. 38:966–972. 2014.PubMed/NCBI View Article : Google Scholar

111 

Le Bricon T, Guidet B, Coudray-Lucas C, Staikowsky F, Gabillet JM, Offenstadt G, Giboudeau J and Cynober L: Biochemical assessment of nutritional status in patients with chronic obstructive pulmonary disease and acute respiratory failure on admission to an intensive care unit. Clin Nutr. 13:98–104. 1994.PubMed/NCBI View Article : Google Scholar

112 

Germano Borges de Oliveira Nascimento Freitas R, Negrao Nogueira RJ and Hessel G: Protein needs of Critically Ill patients receiving parenteral nutrition. Nutr Hosp. 32:250–255. 2015.PubMed/NCBI View Article : Google Scholar

113 

Yu PJ, Cassiere HA, Dellis SL, Manetta F, Kohn N and Hartman AR: Impact of preoperative prealbumin on outcomes after cardiac surgery. JPEN J Parenter Enteral Nutr. 39:870–874. 2015.PubMed/NCBI View Article : Google Scholar

114 

Alsemgeest SP, Kalsbeek HC, Wensing T, Koeman JP, van Ederen AM and Gruys E: Concentrations of serum amyloid-A (SAA) and haptoglobin (HP) as parameters of inflammatory diseases in cattle. Vet Q. 16:21–23. 1994.PubMed/NCBI View Article : Google Scholar

115 

Mills PC, Auer DE, Kramer H, Barry D and Ng JC: Effects of inflammation-associated acute-phase response on hepatic and renal indices in the horse. Aust Vet J. 76:187–194. 1998.PubMed/NCBI View Article : Google Scholar

116 

Mills PC, Ng JC and Auer DE: The effect of inflammation on the disposition of phenylbutazone in thoroughbred horses. J Vet Pharmacol Ther. 19:475–481. 1996.PubMed/NCBI View Article : Google Scholar

117 

Mills PC, Ng JC and Auer DE: The effect of the acute-phase response on in vitro drug metabolism and plasma protein binding in the horse. Vet Res Commun. 21:361–368. 1997.PubMed/NCBI View Article : Google Scholar

118 

Jacobsen S, Andersen PH, Toelboell T and Heegaard PM: Dose dependency and individual variability of the lipopolysaccharide-induced bovine acute phase protein response. J Dairy Sci. 87:3330–3339. 2004.PubMed/NCBI View Article : Google Scholar

119 

Araujo LM, Ribeiro OG, Siqueira M, De Franco M, Starobinas N, Massa S, Cabrera WH, Mouton D, Seman M and Ibañez OM: Innate resistance to infection by intracellular bacterial pathogens differs in mice selected for maximal or minimal acute inflammatory response. Eur J Immunol. 28:2913–2920. 1998.PubMed/NCBI View Article : Google Scholar

120 

Verschuur M, van der Beek MT, Tak HS, Visser LG and de Maat MP: Interindividual variation in the response by fibrinogen, C-reactive protein and interleukin-6 to yellow fever vaccination. Blood Coagul Fibrinolysis. 15:399–404. 2004.PubMed/NCBI View Article : Google Scholar

121 

Elsasser TH, Blum JW and Kahl S: Characterization of calves exhibiting a novel inheritable TNF-alpha hyperresponsiveness to endotoxin: Associations with increased pathophysiological complications. J Appl Physiol (1985). 98:2045–2055. 2005.PubMed/NCBI View Article : Google Scholar

122 

Park HL, Shim SH, Lee EY, Cho W, Park S, Jeon HJ, Ahn SY, Kim H and Nam JH: Obesity-induced chronic inflammation is associated with the reduced efficacy of influenza vaccine. Hum Vaccin Immunother. 10:1181–1186. 2014.PubMed/NCBI View Article : Google Scholar

123 

Frasca D, Ferracci F, Diaz A, Romero M, Lechner S and Blomberg BB: Obesity decreases B cell responses in young and elderly individuals. Obesity (Silver Spring). 24:615–625. 2016.PubMed/NCBI View Article : Google Scholar

124 

Lisciandro JG and van den Biggelaar AH: Neonatal immune function and inflammatory illnesses in later life: Lessons to be learnt from the developing world? Clin Exp Allergy. 40:1719–1731. 2010.PubMed/NCBI View Article : Google Scholar

125 

Eckersall PD, Lawson FP, Bence L, Waterston MM, Lang TL, Donachie W and Fontaine MC: Acute phase protein response in an experimental model of ovine caseous lymphadenitis. BMC Vet Res. 3(35)2007.PubMed/NCBI View Article : Google Scholar

126 

EL-Deeb WM and El-Bahr SM: Selected biochemical indicators of equine rhabdomyolysis in Arabian horses: Acute phase proteins and trace elements. Equine Vet Sci. 34:484–488. 2014. View Article : Google Scholar

127 

EL-Deeb WM, El-Moslemany AM and Salem MA: Cardiac troponin I and immune inflammatory response in horses with strangles. Equine Vet Sci. 51:18–23. 2017. View Article : Google Scholar

128 

Jones SA: Directing transition from innate to acquired immunity: Defining a role for IL-6. J Immunol. 175:3463–3468. 2005.PubMed/NCBI View Article : Google Scholar

129 

Chomarat P, Banchereau J, Davoust J and Palucka AK: IL-6 switches the differentiation of monocytes from dendritic cells to macrophages. Nat Immunol. 1:510–514. 2000.PubMed/NCBI View Article : Google Scholar

130 

van der Beek MT, Visser LG and de Maat MP: Yellow fever vaccination as a model to study the response to stimulation of the inflammation system. Vascul Pharmacol. 39:117–121. 2002.PubMed/NCBI View Article : Google Scholar

131 

Kurtz SL, Foreman O, Bosio CM, Anver MR and Elkins KL: Interleukin-6 is essential for primary resistance to Francisella tularensis live vaccine strain infection. Infect Immun. 81:585–597. 2013.PubMed/NCBI View Article : Google Scholar

132 

Tsai MY, Hanson NQ, Straka RJ, Hoke TR, Ordovas JM, Peacock JM, Arends VL and Arnett DK: Effect of influenza vaccine on markers of inflammation and lipid profile. J Lab Clin Med. 145:323–327. 2005.PubMed/NCBI View Article : Google Scholar

133 

Posthouwer D, Voorbij HA, Grobbee DE, Numans ME and van der Bom JG: Influenza and pneumococcal vaccination as a model to assess C-reactive protein response to mild inflammation. Vaccine. 23:362–365. 2004.PubMed/NCBI View Article : Google Scholar

134 

Treanor JJ, Taylor DN, Tussey L, Hay C, Nolan C, Fitzgerald T, Liu G, Kavita U, Song L, Dark I and Shaw A: Safety and immunogenicity of a recombinant hemagglutinin influenza-flagellin fusion vaccine (VAX125) in healthy young adults. Vaccine. 28:8268–8274. 2010.PubMed/NCBI View Article : Google Scholar

135 

Njoroge L, Khayata M, Zacharias M and ElAmm C: Acute myocarditis following influenza vaccine in a heart transplant patient. J Cardiac Failure. 24 (Suppl):S45–S46. 2018. View Article : Google Scholar

136 

Garcon N, Leroux-Roels G and Cheng W: Vaccine adjuvants. In: Understanding Modern Vaccines: Perspectives in Vaccinology: Elsevier, pp89-113, 2011.

137 

Mutwiri G, Gerdts V, van Drunen Littel-van den Hurk S, Auray G, Eng N, Garlapati S, Babiuk LA and Potter A: Combination adjuvants: The next generation of adjuvants? Expert Rev Vaccines. 10:95–107. 2011.PubMed/NCBI View Article : Google Scholar

138 

Singh M, Ugozzoli M, Kazzaz J, Chesko J, Soenawan E, Mannucci D, Titta F, Contorni M, Volpini G, Del Guidice G and O'Hagan DT: A preliminary evaluation of alternative adjuvants to alum using a range of established and new generation vaccine antigens. Vaccine. 24:1680–1686. 2006.PubMed/NCBI View Article : Google Scholar

139 

Leroux-Roels G: Unmet needs in modern vaccinology: Adjuvants to improve the immune response. Vaccine. 28 (Suppl 3):C25–C36. 2010.PubMed/NCBI View Article : Google Scholar

140 

Destexhe E, Prinsen MK, van Schöll I, Kuper CF, Garçon N, Veenstra S and Segal L: Evaluation of C-reactive protein as an inflammatory biomarker in rabbits for vaccine nonclinical safety studies. J Pharmacol Toxicol Methods. 68:367–373. 2013.PubMed/NCBI View Article : Google Scholar

141 

El Yousfi M, Mercier S, Breuillé D, Denis P, Papet I, Mirand PP and Obled C: The inflammatory response to vaccination is altered in the elderly. Mech Ageing Dev. 126:874–881. 2005.PubMed/NCBI View Article : Google Scholar

142 

Liuba P, Aburawi EH, Pesonen E, Andersson S, Truedsson L, Ylä-Herttuala S and Holmberg L: Residual adverse changes in arterial endothelial function and LDL oxidation after a mild systemic inflammation induced by influenza vaccination. Ann Med. 39:392–399. 2007.PubMed/NCBI View Article : Google Scholar

143 

He R, Shepard LW, Chen J, Pan ZK and Ye RD: Serum amyloid A is an endogenous ligand that differentially induces IL-12 and IL-23. J Immunol. 177:4072–4079. 2006.PubMed/NCBI View Article : Google Scholar

144 

Lu P, Liu J and Pang X: Pravastatin inhibits fibrinogen- and FDP-induced inflammatory response via reducing the production of IL-6, TNF-alpha and iNOS in vascular smooth muscle cells. Mol Med Rep. 12:6145–6151. 2015.PubMed/NCBI View Article : Google Scholar

145 

Lazzaro M, Bettegazzi B, Barbariga M, Codazzi F, Zacchetti D and Alessio M: Ceruloplasmin potentiates nitric oxide synthase activity and cytokine secretion in activated microglia. J Neuroinflammation. 11(164)2014.PubMed/NCBI View Article : Google Scholar

146 

Martinez Cordero E, Gonzalez MM, Aguilar LD, Orozco EH and Hernandez Pando R: Alpha-1-acid glycoprotein, its local production and immunopathological participation in experimental pulmonary tuberculosis. Tuberculosis (Edinb). 88:203–211. 2008.PubMed/NCBI View Article : Google Scholar

147 

de Serres F and Blanco I: Role of alpha-1 antitrypsin in human health and disease. J Intern Med. 276:311–335. 2014.PubMed/NCBI View Article : Google Scholar

148 

Haversen L, Ohlsson BG, Hahn-Zoric M, Hanson LA and Mattsby-Baltzer I: Lactoferrin down-regulates the LPS-induced cytokine production in monocytic cells via NF-kappa B. Cell Immunol. 220:83–95. 2002.PubMed/NCBI View Article : Google Scholar

149 

Du Clos TW: Function of C-reactive protein. Ann Med. 32:274–278. 2000.PubMed/NCBI View Article : Google Scholar

150 

Spadaro M, Montone M, Arigoni M, Cantarella D, Forni G, Pericle F, Pascolo S, Calogero RA and Cavallo F: Recombinant human lactoferrin induces human and mouse dendritic cell maturation via Toll-like receptors 2 and 4. FASEB J. 28:416–429. 2014.PubMed/NCBI View Article : Google Scholar

151 

Feelders RA, Vreugdenhil G, Eggermont AM, Kuiper-Kramer PA, van Eijk HG and Swaak AJ: Regulation of iron metabolism in the acute-phase response: Interferon gamma and tumour necrosis factor alpha induce hypoferraemia, ferritin production and a decrease in circulating transferrin receptors in cancer patients. Eur J Clin Invest. 28:520–527. 1998.PubMed/NCBI View Article : Google Scholar

152 

Bartalena L, Farsetti A, Flink IL and Robbins J: Effects of interleukin-6 on the expression of thyroid hormone-binding protein genes in cultured human hepatoblastoma-derived (Hep G2) cells. Mol Endocrinol. 6:935–942. 1992.PubMed/NCBI View Article : Google Scholar

153 

Louagie H, Delanghe J, Desombere I, De Buyzere M, Hauser P and Leroux-Roels G: Haptoglobin polymorphism and the immune response after hepatitis B vaccination. Vaccine. 11:1188–1190. 1993.PubMed/NCBI View Article : Google Scholar

154 

Borthwick NJ, Lane T, Moyo N, Crook A, Shim JM, Baines I, Wee EG, Hawkins PN, Gillmore JD, Hanke T and Pepys MB: Randomized phase I trial HIV-CORE 003: Depletion of serum amyloid P component and immunogenicity of DNA vaccination against HIV-1. PLoS One. 13(e0197299)2018.PubMed/NCBI View Article : Google Scholar

155 

Creech CB, Frenck RW Jr, Sheldon EA, Seiden DJ, Kankam MK, Zito ET, Girgenti D, Severs JM, Immermann FW, McNeil LK, et al: Safety, tolerability, and immunogenicity of a single dose 4-antigen or 3-antigen Staphylococcus aureus vaccine in healthy older adults: Results of a randomised trial. Vaccine. 35:385–394. 2017.PubMed/NCBI View Article : Google Scholar

156 

Blom M, Prag JB and Norredam K: Alpha 1-Acid glycoprotein, alpha 1-antitrypsin, and ceruloplasmin in human intestinal helminthiases. Am J Trop Med Hyg. 28:76–83. 1979.PubMed/NCBI View Article : Google Scholar

157 

Naylor C, Lu M, Haque R, Mondal D, Buonomo E, Nayak U, Mychaleckyj JC, Kirkpatrick B, Colgate R, Carmolli M, et al: Environmental enteropathy, oral vaccine failure and growth faltering in infants in bangladesh. EBioMedicine. 2:1759–1766. 2015.PubMed/NCBI View Article : Google Scholar

158 

Hwang SA, Kruzel ML and Actor JK: Lactoferrin augments BCG vaccine efficacy to generate T helper response and subsequent protection against challenge with virulent Mycobacterium tuberculosis. Int Immunopharmacol. 5:591–599. 2005.PubMed/NCBI View Article : Google Scholar

159 

Carty CL, Heagerty P, Nakayama K, McClung EC, Lewis J, Lum D, Boespflug E, McCloud-Gehring C, Soleimani BR, Ranchalis J, et al: Inflammatory response after influenza vaccination in men with and without carotid artery disease. Arterioscler Thromb Vasc Biol. 26:2738–2744. 2006.PubMed/NCBI View Article : Google Scholar

160 

Patel C and Shah HH: Membranous nephropathy and severe acute kidney injury following influenza vaccination. Saudi J Kidney Dis Transpl. 26:1289–1293. 2015.PubMed/NCBI View Article : Google Scholar

161 

Bos MP, Poolman J, Stork M, Tommassen JPM and Weynants V: Vaccine comprising protein NMB0964 from Neisseria meningitidis. US Patent 9,259,461. Filed March 6, 2009; issued August 4, 2011.

Related Articles

Journal Cover

April-2020
Volume 12 Issue 4

Print ISSN: 2049-9434
Online ISSN:2049-9442

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Khalil RH and Khalil RH: Types of acute phase reactants and their importance in vaccination (Review). Biomed Rep 12: 143-152, 2020
APA
Khalil, R.H., & Khalil, R.H. (2020). Types of acute phase reactants and their importance in vaccination (Review). Biomedical Reports, 12, 143-152. https://doi.org/10.3892/br.2020.1276
MLA
Khalil, R. H., Al-Humadi, N."Types of acute phase reactants and their importance in vaccination (Review)". Biomedical Reports 12.4 (2020): 143-152.
Chicago
Khalil, R. H., Al-Humadi, N."Types of acute phase reactants and their importance in vaccination (Review)". Biomedical Reports 12, no. 4 (2020): 143-152. https://doi.org/10.3892/br.2020.1276