Chapter 27

Cellular and Physiological Effects of Arginine in Seniors

Vance L. Albaugh, Melissa K. Stewart and Adrian Barbul,    Vanderbilt University School of Medicine, Nashville, TN, United States

Abstract

Arginine is a conditionally essential amino acid that plays a number of critical roles in human health and disease. Since discovering arginine to be the nitrogen source for nitric oxide, biomedical research has intensely focused on it and its vasodilatory effects. However, other recent work has highlighted not only these potential cardiovascular benefits of arginine and its metabolites but also the advances in the knowledge of arginine biochemistry, organ–organ trafficking, and insights from clinical studies examining its role of arginine as a therapeutic modality in a variety of disease states. These advances, along with a comprehensive review of arginine biochemistry and physiology, are reviewed within this chapter and aim to highlight this robust area of medical research.

Keywords

Arginine; citrulline; nutritional physiology; metabolism; arginase; nitric oxide; nitric oxide synthase

Introduction

Arginine (2-amino-5-guanidinovaleric acid) is an alpha amino acid that has garnered significant research interest for its involvement in a several physiological and pathophysiological processes in human health and disease. Arginine was first shown to be a constituent of animal proteins in 1895 by Hedin (Rogers and Visek, 1985), which spurred interest in the compound’s biologic actions. The discovery of arginine as the nitrogen source of nitric oxide (Palmer et al., 1988) has reinvigorated basic and clinical research on arginine and its potential uses as a nutraceutical or dietary supplement. Given its broad range of physiological actions, many of which are affected by normal physiological aging as well as pathological processes, the aim of this chapter is to review arginine’s broad reach in human biology and identify the current state of research in many of these areas.

Sources of Arginine

Dietary Arginine and Nutritional Importance

Arginine is considered a “conditionally essential” amino acid, which means that its nutritional status (i.e., essential versus nonessential) depends on the health or developmental status of the organism. The human body has the ability to synthesize arginine de novo, but if arginine demand increases, the endogenous supply can be insufficient. Thus, during developmental periods or time of rapid growth and cellular turnover, arginine requirements are significantly increased and necessitate exogenous arginine supply. This need for additional arginine has been confirmed by numerous studies demonstrating that endogenous production is insufficient to meet the metabolic needs of growing infants and other neonates (e.g., piglets, young rats), of adults under highly catabolic conditions (e.g., sepsis, burn injury, trauma), or following extensive bowel resection (Wakabayashi et al., 1994).

Arginine has three potential sources in humans: (1) dietary arginine as a constituent of consumed protein, (2) recycled arginine from endogenous protein turnover, and (3) arginine synthesized de novo. Dietary arginine is available in a variety of foods but is most notably found in significant quantifies in seeds, nuts, seafood, algae, meats, rice protein concentrate, and soy protein isolate (Ros, 2015). In general, these foods have the highest per weight amounts of arginine (Hu et al., 1998; King et al., 2008). In addition to naturally occurring dietary sources, supplements are commercially available to consumers for several putative indications.

Like other dietary amino acids that are digested and liberated by intestinal peptidases, arginine is taken up in the diet through apical enterocyte membranes. Under normal conditions, approximately 40% (30–44%) of dietary arginine is extracted by the splanchnic tissues (Castillo et al., 1993), which is likely secondary to the relatively high arginase activity in the intestinal mucosa, a tissue with a high rate of cellular turnover. Dietary arginine uptake by enterocytes uses several amino acid transporters. These transport systems continue to be identified and studied but include three major categories: system y+image(cationic amino acid-specific transporters), systems BO,+image and bO,+image (the sodium-dependent and sodium-independent broad-scope transporters, respectively), and the system y+Limage (the cation-modulated broad-scope transporter). These transporters have been extensively reviewed elsewhere and continue to be characterized (for reviews, see Devés and Boyd, 1998). Although a discussion of these transporters is beyond the scope of this chapter, it is important to recognize that they are expressed on a variety of cell types in addition to enterocytes and determine the flux of arginine and other amino acids into and across cells and tissues, depending on the density of expression and cell polarity. Virtually all cells have the capacity for arginine transport across the plasma membrane, which facilitates its wide range of biologic uses (Devés and Boyd, 1998).

Endogenous Arginine and Whole-Body Arginine Flux

Arginine homeostasis, like that of most amino acids in the plasma, is a dynamic process. Typical arginine plasma concentrations in adult humans range from 80 to 120 µM, relative to the 2.5-mM concentration of total plasma amino acids (Brosnan, 2003). The half-life of arginine in humans and many other mammalian species is approximately one hour (Wu et al., 2007). Human whole-body arginine flux in the postabsorptive state is approximately 70 µM/h to 90 µM/kg/h (Castillo, 1996; Castillo et al., 1995). Thus, extrapolating that flux to a 24 h rate corresponds to approximately 15–20 g of arginine per day. Given that average daily dietary arginine consumption is only 4–6 g per day (Visek, 1986), a significant amount of recycling has to occur to meet basal arginine requirements.

Aside from exogenous dietary sources, baseline metabolic requirements for arginine are fulfilled by endogenous production. This occurs via three separate pathways with recycling from protein turnover being responsible for approximately 85% of available circulating arginine in humans (Castillo, 1996; Dejong et al., 1998). In addition to arginine recovery from protein turnover, arginine can be synthesized from other amino acid metabolites. In the adult human, de novo arginine synthesis can account for approximately 10–15% of arginine production (Luiking et al., 2012). This endogenous synthesis occurs from the conversion of proline, glutamate or glutamine, and citrulline, a nonproteinogenic amino acid that can be converted to arginine through organ-to-organ flux and metabolism by the intestine and kidney. The de novo synthesis of arginine from citrulline accounts for the largest amount of endogenously synthesized arginine (approximately 60–80%) (Curis et al., 2007; 2005). In fact, oral citrulline supplementation can be used to increase plasma arginine quite effectively, as citrulline has an exceedingly high bioavailability and virtually no citrulline is lost in the urine (Rougé et al., 2007; Urschel et al., 2006).

Biochemistry of Arginine

Our understanding of the regulation of arginine metabolism has significantly increased over the last several decades. Arginine is well known for its role in maintaining long-term nitrogen balance as part of the urea cycle, and as the source of nitrogen in the synthesis of nitric oxide (NO). Our appreciation for arginine biochemistry and homeostasis continues to grow and includes a previously unrecognized organ-to-organ trafficking and subcellular localization of arginine-metabolizing enzymes. Each of these processes adds a number of layers of complexity to the role arginine plays in cellular metabolism. However, we have little knowledge of how the expression of catabolic or anabolic enzymes in these pathways changes with aging. We also do not know if arginine flux changes over an organism’s or human’s lifetime.

Arginine Catabolism

Several pathways degrade arginine for use in vital biologic processes. These enzymes have a wide tissue distribution, and the absence of any one of them can have detrimental effects. One of the most studied arginine-metabolizing enzymes is arginase, the final enzyme of the urea cycle that regenerates ornithine. Unlike the other urea cycle enzymes that exist as single isoforms, two human arginase isoforms are encoded by separate genes: arginase I (AI), which is cytosolic, and arginase II (AII), which localizes to the mitochondrion. (Note: Kmimage values for AI and AII isoforms are ~0.1 and ~5 mM, respectively (Moinard et al., 2015). Previously the AI and AII isoforms were referred to as hepatic arginase and extrahepatic arginase, respectively. Both arginase isoforms catalyze the hydrolysis of arginine to form ornithine and urea:

arginine+H2Oornithine+ureaimage

The AI gene was cloned in 1986 by Mori and colleagues (Haraguchi et al., 1987) and has a wide tissue distribution, including a high expression in the liver and, to a lesser extent, erythrocytes, gastrointestinal tract, thymus, skin, uterus, and sympathetic ganglia (Yu et al., 2001, 2003). The AII enzyme was discovered in a patient with known AI deficiency who was noted to have residual arginase activity. The AII gene includes a mitochondrial localization signal that targets it for the mitochondrial matrix. Unlike AI, the AII protein is a ubiquitously expressed enzyme (Morris, 2002) with its highest expression in the kidney, gastrointestinal tract, brain, and lactating mammary gland (Yu et al., 2003).

First discovered in the 1980s, nitric oxide synthase is another well studied enzyme that catalyzes the generation of NO and citrulline from arginine:

arginine+NADPH+O2+H+NO+citrulline+NADP+

image

There are three isotypes of NOS: endothelial NOS (eNOS), inducible NOS (iNOS), and neuronal NOS (nNOS). NOS1 (neuronal NOS) and NOS3 (eNOS) are constitutive enzymes that are controlled by intracellular calcium and calmodulin (Luiking et al., 2012). NOS2 is inducible at the level of gene transcription, calcium independent, and expressed by macrophages and other tissues in response to (pro)inflammatory mediators (Luiking et al., 2012). From a functional perspective, the Kmimage values for arginine of the nNOS, eNOS, and iNOS enzymes are approximately 1.8, 3.6, and 12.5 µM, respectively (Gad, 2010).

The relationship between arginase and NOS, and their competition for available arginine, has led to a significant amount of investigation. The biological basis for the so-called arginine paradox remains unknown—that is, the Kmimage of NOS is in the micromolar range while the Kmimage of arginase is in the millimolar range. These enzymes appear to compete for arginine substrate, although biochemically this cannot be explained by Kmimage alone. Thus, the subcellular localization of these enzymes or local concentrations of arginine substrate (or both) must underlie the regulation of these two enzymes. Regardless, several studies have examined the possibilities of arginase contributing to the effects of aging such as endothelial dysfunction and hypertension. Exogenous arginine administration induces the expression of both the AI and AII enzymes and leads to early cellular senescence (Xiong et al., 2014). Increased arginase activity negatively influences eNOS efficiency in producing NO. This negative effect has been observed in hypertensive humans in whom exogenous arginine is ineffective in inducing vasodilation (Holowatz and Kenney, 2007). It is thought that increased arginase expression and activity leads to decreased availability of arginine for NO production. Consistent with this hypothesis, the ornithine produced by excessive arginase activity may be shunted to polyamine biosynthesis for vascular smooth muscle hypertrophy, further exacerbating vascular dysfunction with aging (Durante, 2013). Arginase expression is increased with aging in vascular smooth muscle cells, and this increased activity may decrease local arginine availability for vascular endothelial cells, thus decreasing NO biosynthesis. Not all NOS isoforms behave in the same manner, however, with iNOS actually promoting arginase activation as part of a negative feedback mechanism (Santhanam et al., 2007; Shin et al., 2012). Regardless, these cell-to-cell interactions between arginase and NOS affect each other’s activity by decreasing arginine availability (Johnson et al., 2005; Kim et al., 2009).

Creatine, a molecule important in transient energy storage from high-energy phosphates of adenosine triphosphate (ATP), is a third molecule synthesized from arginine. The creatine synthetic pathway is important in tissues with fluctuating energy needs such as muscle but is important as a rapid source of ATP-derived energy in many other tissues, including the brain, retina, and spermatozoa (Wyss and Kaddurah-Daouk, 2000). Arginine is metabolized by the enzyme arginine:glycine amidinotransferase (Km9.2 mMimage) (Sipilä, 1980), which catalyzes the condensation of arginine and glycine in the following reaction to form a guanidino intermediate:

arginine+glycineguanidinoacetatecreatine

image

Production of this guanidinoacetate intermediate is the rate-limiting step in creatine production, an irreversible process. Creatine is also obtained from diet, so overall flux of arginine into this metabolic pathway is largely determined by the availability of dietary creatine. It is estimated that 20% of dietary intake of arginine is used for creatine phosphate synthesis (Brosnan and Brosnan, 2010). Significant amounts of dietary creatine can act as a feedback repressor of this enzyme to limit de novo creatine production from arginine (Stead et al., 2001; Wyss and Kaddurah-Daouk, 2000).

The fourth and last pathway that utilizes arginine as a direct precursor is the synthesis of agmatine, a compound involved in the synthesis of polyamines. The polyamines (i.e., putrescine, spermine, and spermidine) are vital to cellular proliferation; dysregulated polyamine synthesis is associated with cancer. Agmatine is an intermediate whose biologic functions have not been completely determined, but recent evidence suggests that it may play an inhibitory role on polyamine biosynthesis (Satriano, 2003). The enzyme that generates agmatine from arginine is arginine decarboxylase (Kmimage 0.75 mM) (Regunathan and Reis, 2000). As its name implies, it decarboxylates arginine to form agmatine and carbon dioxide:

arginineagmatine+CO2image

The agmatine generated in this reaction can then be acted on by another enzyme, agmatinase (Iyer et al., 2002), which generates putrescine (a polyamine) and urea:

agmatineurea+putrescineimage

The biologic actions of agmatine continue to be debated, as both the liver and kidney have high levels of arginine decarboxylase to produce agmatine, two tissues with high capacities for arginine biosynthesis (Lortie et al., 1996; Morrissey et al., 1995). There are conflicting reports on whether or not mammalian cells express significant arginine decarboxylase activity (Morris, 2004). The role of agmatine, agmatinase, and arginine decarboxylase continue to be an area of research and may potentially reveal further roles in mammals for this arginine-derived compound in the future.

Arginine Synthesis

Intestinal–Renal Axis

Arginine is synthesized endogenously through two related pathways. The first pathway is referred to as the intestinal–renal axis. First identified in the 1980s, it involves the coordinated actions of the intestine and kidneys in organ-to-organ metabolite shunting that culminates in de novo renal arginine synthesis and release (Fig. 27.1). Metabolic precursors of arginine are converted to citrulline, a common metabolic intermediate.

image
Figure 27.1 The intestinal–renal axis of arginine synthesis.
Endogenous arginine synthesis involves a coordinated organ-to-organ trafficking of arginine precursors made possible by the expression of a partial urea cycle in the intestines and kidney, as well as a lack of hepatic citrulline clearance. The complete urea cycle consists of five enzymes: (1) carbamoyl phosphate synthetase 1, (2) ornithine transcarbamoylase (OTC), (3) argininosuccinate synthetase (AS), (4) argininosuccinate lyase (ASL), and (5) arginase (Arg). Dietary nutrients (e.g., amino acids, carbohydrates) are absorbed by CPS-1 and OTC-expressing intestinal enterocytes, which convert arginine precursors to citrulline. This conversion arginine precursors to bypass hepatic portal clearance, which is essentially zero for citrulline. This effectively shunts arginine precursors in the form of citrulline to the kidney. Once in the kidney, citrulline can be converted to arginine by the expression of AS and ASL, releasing arginine into circulation for use by peripheral tissues. Unlike arginine precursors and citrulline, other excess amino acid nitrogen is cleared in the hepatic portal circulation and converted to urea or NH3, which is then excreted into the urine by the kidney.

Cytosolic arginine can be hydrolyzed by arginase in the intestine to produce ornithine and urea:

arginine+H2Ourea+ornithineimage

Of plasma arginine flux, 15% enters this extrahepatic arginase pathway that degrades arginine to ornithine and urea (Castillo, 1996). Similarly, proline, glutamine, and glutamate can also be metabolized to ornithine, which is subsequently converted to citrulline. Proline is first metabolized to pyrroline-5-carboxylate (P5C) via proline oxidase, and then ornithine aminotransferase coverts P5C to ornithine:

prolinepyrroline~5~carboxylateornithineimage

Glutamine can also be metabolized by glutaminase to form glutamate, which is then converted to P5C and ornithine via ornithine aminotransferase (similar to proline):

glutamineglutamatepyrroline~5~carboxylateornithineimage

Carbamoyl phosphate and ornithine are then finally condensed in a reaction catalyzed by ornithine transcarbamoylase:

ornithine+carbamoyl phosphatecitrulline+Piimage

which generates citrulline in the mitochondria and is then transported to the cytosol ready for export. All of these enzymes that lead to the production of citrulline are expressed within the intestinal enterocytes and represent the expression of a partial urea cycle that exists within the small intestine. This partial urea cycle converts ornithine to citrulline before exporting it from the cell. In addition, dietary ornithine or amino acids that are degraded to ornithine can be recycled to arginine using the classical urea cycle enzymes.

After citrulline is produced by the intestinal enterocytes, it is released into the portal circulation and then recirculates via the arterial circulation to the kidneys. The kidneys efficiently extract the circulating citrulline and convert it to arginine (Fig. 27.1), which becomes available for the rest of the cells and tissues of the body. Experimental data have demonstrated that this de novo arginine synthesis via the intestinal–renal axis is tightly linked to citrulline availability (Luiking et al., 2012). This pathway of converting arginine precursors to a common nonarginine compound (i.e., citrulline) appears advantageous because the liver has virtually zero clearance of citrulline from the portal circulation (Morris, 2002). In stark contrast, the kidney efficiently clears circulating citrulline (Wu and Morris, 2003).

Citrulline comes from several possible sources, including (1) diet; (2) arginine from the NOS pathway; (3) ornithine catabolism of proline, glutamate, or glutamine, and (4) asymmetric dimethyl arginine breakdown. In mice, and potentially humans, the predominant citrulline precursors in this situation are dietary arginine as well as citrulline (Marini, 2012).

Once in the kidney, two cytosolic enzymes that are classically expressed in the urea cycle allow for the conversion of ornithine to arginine. The first enzyme is argininosuccinate synthetase (AS), which condenses citrulline and aspartate to form argininosuccinate. Then the second enzyme, argininosuccinate lyase, generates arginine as well as fumarate as a by-product:

citrulline+aspartateargininosuccinatearginine+fumarateimage

Note that this is an energetically costly process to the cell, using two molar equivalents of ATP for the cycling process, although the benefit of shunting arginine around the liver appears to be desirable to provide sufficient arginine to the remaining tissues of the body.

Citrulline–Nitric Oxide Cycle

The citrulline–nitric oxide cycle is a theoretical pathway for which the magnitude of flux is inherently difficult to measure and not currently well appreciated. This pathway, however, seems to be active in cells with endogenous NOS activity, potentially as a way to ensure a continuous supply of arginine for NO production. Cells in which arginine would otherwise be limiting can recycle the citrulline produced as a by-product of NOS using Krebs cycle intermediates to continuously generate arginine (Wu and Morris, 2003). Similar to the intestinal–renal axis, the citrulline–NO cycle uses urea cycle enzymes. The citrulline that is produced through the activity of NOS is converted back to arginine via the action of the enzymes argininosuccinate synthase and argininosuccinate lyase. In the first reaction, argininosuccinate synthase condenses citrulline with aspartate:

aspartate+citrullineargininosuccinateimage

Subsequently, argininosuccinate is converted back to arginine by the enzyme argininosuccinate lyase (ASL):

argininosuccinatearginine+fumarateimage

Hypothetically, this citrulline–NO cycle could continuously generate arginine from citrulline as long as aspartate does not become a limiting factor. Thus, the other vital portion of this cycle is the continuous supply of aspartate. The beauty and simplicity of this cycle, however, is the pathway for the regeneration of aspartate and removal of fumarate that could otherwise build up significant concentrations. The cell appears to handle this via two routes. The first is by the actions of fumarase, which catalyzes the conversion of fumarate to malate, which then can be metabolized by malate dehydrogenase to form oxaloacetate:

fumaratemalateoxaloacetateimage

Oxaloacetate can be subsequently acted on by the enzyme aspartate transaminase to regenerate aspartate:

oxaloacetate+glutamateα~ketoglutarate+aspartateimage

Alternatively, the fumarate can be converted to malate; then, instead of malate dehydrogenase, malic enzyme can convert the malate to pyruvate. Subsequently, pyruvate can be acted on by pyruvate carboxylase to reform oxaloacetate, which can eventually generate aspartate using the enzymatic conversions above:

fumaratemalatepyruvateoxaloacetateimage

As previously mentioned above, nearly all known cell types express AS and ASL. Thus, further studies are needed to determine the importance and flux rates of these processes in vivo (Morris, 2002). Regardless, experimental data from a number of investigators have suggested that this pathway is induced along with iNOS activity (Mori and Gotoh, 2000), which makes sense teleologically as a means to circumvent a potential arginine deficiency in a time of stress or need (e.g., infection and iNOS activity).

Arginine Physiology and Biological Importance

As mentioned, arginine and its metabolites are important in numerous biologic processes. Arginine is most notable as a urea cycle intermediate and carrier of excretable nitrogen, as well as a precursor of NO that is important in vasodilation and the immune response. More recent studies continue to examine arginine and the arginases, which have critical importance in vascular pathology (Durante et al., 2007). Given our expanded understanding of the biologic actions of arginine, the physiologic effects of arginine supplementation have also been an area of intense investigation. How arginine metabolism changes over time and with aging of the body is less well understood and represent important areas for future research. Each of the aforementioned categories has a specific connection and interest to the aging population and will be broadly discussed in the following section.

Protein Synthesis and Degradation

Other important but less well recognized biological effects involving arginine are its role in targeting proteins for degradation as they come off the ribosome. This so-called N-end rule for protein degradation uses an amino terminal arginine as a marker for protein degradation, which may act as a quality control for cellular protein synthesis (Tasaki and Kwon, 2007; Varshavsky, 1996). Conversely, arginine is a potent stimulator as well as a substrate in protein synthesis. It is estimated that approximately 80% of arginine needed for protein translation comes from recycled arginine (Luiking et al., 2012). Moreover, arginine has been shown in several studies to stimulate the mammalian target of rapamycin (mTOR) pathway, a known regulator of protein translational initiation (Corl et al., 2008; Wang et al., 2012). This effect on protein translation via mTOR is independent of NO (Bauchart-Thevret et al., 2010). Specifically, arginine directly stimulates p70 S6 kinase and phosphorylation of eukaryotic initiation factor 4E binding protein-1 (Ban et al., 2004). In older individuals, evidence suggests that arginine may have protective effects on muscle wasting and could improve lean muscle mass. These studies, however, have examined arginine in combination therapies and not arginine supplementation alone (Clark et al., 2000; Flakoll et al., 2004; May et al., 2002). Overall, these studies have examined functional outcomes in groups with single or multiple amino acid supplements. The effects of these supplements may be either direct effects on skeletal muscle or indirect effects through other hormonal pathways such as growth hormone and insulin and insulin-like growth factor 1.

Acid–Base Homeostasis and the Urea Cycle

In addition to being an important anaplerotic precursor of the Krebs cycle through conversion to glutamate (Owen et al., 2002), arginine and its metabolism are vitally important in humans as an intermediate in the urea cycle. The final step in the pathway is catalysis of arginine by the enzyme arginase, which produces ornithine and urea, allowing the urea to be available for excretion and regenerating ornithine for the cycle to continue to function. It is important to note, as stressed by Morris, that many textbooks erroneously state that a significant amount of arginine comes from its synthesis from the urea cycle. This is far from the case, as virtually no net synthesis of arginine occurs in the liver (Morris, 2009).

Beyond its classic role in nitrogenous waste detoxification, arginine and the urea cycle also play an important role in acid–base homeostasis. This may be less well clinically appreciated, but the urea cycle is bicarbonate consuming (Häussinger, 1986) and critical in the maintenance of whole-body acid–base balance (Häussinger et al., 1984). Aside from the link between the urea cycle and acid–base homeostasis, recent evidence has demonstrated a role for sirtuins, a family of protein deacetylases, as a link between aging and the regulation of the urea cycle (Nakagawa and Guarente, 2009; Nakagawa et al., 2009).

Immunology

As the immune system ages, its capabilities decline and there is increased susceptibility to infection and cancer, with an increased incidence of autoimmune disorders (Agarwal and Busse, 2010). Given such, the use of arginine to prevent or overcome immunosenescence has received much attention.

The role of all amino acids in immunity, including arginine, has been well studied. Arginine is an essential determinant in immune cell metabolism. Specifically, arginine has been shown to be important in T-cell activation and proliferation, B-cell development, and cytokine production and activity (Bronte and Zanovello, 2005; Feldmeyer et al., 2012). Both enzymes of arginine metabolism, NOS and arginase, are felt to be responsible for its immunologic activity.

Absence of L-arginine in culture media results in significant reduction in proliferation of T lymphocytes in response to mitogens (Brittenden et al., 1994). The optimal L-arginine concentration in media for maximal T-lymphocyte proliferation is similar to those found in human plasma (Efron et al., 1991). Arginine is also required for the effective induction of cytotoxic T-cell function in vitro (Moriguchi et al., 1987). Jurkat T cells cultured in media lacking L-arginine manifest rapid reduction in the expression of the T-cell receptor (TCR), CD36ζ, caused by a short half-life of CD36ζ-mRNA(Rodriguez et al., 2002). It is important to note that CD36ζ phosphorylation is the rate-limiting step in the assembly and membrane expression of the TCR (Minami et al., 1987). Such reduction in mRNA decreases TCR internalization, thus leading to a decrease in total TCR cell membrane expression (Bronstein-Sitton et al., 1999; Valitutti et al., 1997).

Arginine has also been identified as being important in a specific stage of murine B-cell development (de Jonge et al., 2002a, b). The investigators engineered a transgenic mouse expressing arginase I under the control of the rat intestinal fatty-acid binding promoter and enhancer element. The mice demonstrated a 30–40% reduction of plasma arginine and showed a disruption in the B-cell development at the progenitor B to precursor B cell interface in the bone marrow. There were also decreased B cells in secondary lymphoid organs such as the spleen and Peyer’s patches in the initial 3 weeks of the neonatal period. The defects were reversed with arginine supplementation (Fafournoux et al., 2000; LeBien, 2002).

Arginine is metabolized in macrophages and lymphocytes by two independent enzymatic pathways, NOS and arginase (Ochoa et al., 2001a). Several clinical conditions—including trauma (Munder et al., 1998), sepsis (Carraway et al., 2015; Witte, 2003), and liver transplantation (Ikemoto et al., 1998; Längle et al., 1995)—have been associated with increased arginase activity, leading to excessive catabolism and subsequent depletion of arginine. This coincides with decreased T-cell proliferation. Supplementation with L-arginine increases CD4+ cells (Kirk et al., 1992), suggesting that arginine may play an important role in reversing the immunosuppression observed during periods of stress (Bernstein, 1996). Specifically, trauma is associated with a greater than 10-fold increase in arginase activity (Ochoa et al., 2001b; 1991). In addition, increased arginase activity in the liver has been implicated in the increased tolerance of the liver to organ rejection (Callery et al., 1991; Schrempf Decker et al., 1983). Moreover, growing evidence suggests that arginine is also an important contributor to intestinal immune function (Li et al., 2007).

Following trauma or surgery, arginine-rich diets were found to reverse the alteration of T-cell function that was initially seen as secondary to the pathologic process. In addition, arginine supplementation in patients undergoing major abdominal operations for gastrointestinal malignancies resulted in increased in vitro immune responses, which correlated to decreased wound infections and decreased length of hospital stay (Daly et al., 1988). T-lymphocyte blastogenesis was also preserved and enhanced in moderately stressed intensive care unit patients given an enteral diet containing large amounts of arginine (Daly et al., 1988). Whether these effects translate into an improvement in clinical outcomes in critically ill patients remains unclear.

Of interest, arginine supplementation following vaccination against Streptococcus pneumoniae in people aged 60 and older resulted in increased neutrophil chemotaxis, natural killer cytotoxicity, and serum concentration of immunoglobulin G. These results suggest that arginine supplementation may enhance the immune response elicited by the pneumococcal vaccine in older people (Moriguti et al., 2005).

Wound Healing

In the United States, chronic wounds affect nearly 6.5 million patients and annually cost in excess of $25 to treat. This burden is expected to grow, in part due to an aging population (Sen et al., 2009). Much research has been focused on the prevention and treatment of wounds. Arginine supplementation has been investigated as a means of enhancing wound healing. Although a lot has been learned, much remains unknown.

Seifter and Levinson first postulated the role of arginine in wound healing in 1978. They demonstrated that arginine-deficient animals subjected to dorsal skin incision and closure had decreased wound breaking strength and wound collagen accumulation when compared to animals fed an normal arginine-containing diet (Seifter et al., 1978). Subsequent experiments revealed that chow-fed rats given a dietary supplement of 1% arginine had enhanced wound healing responses as assessed by wound breaking strength and collagen synthesis when compared to chow-fed controls. Likewise, parenteral fed rats given an amino acid mixture containing high doses of arginine (7.5 g/L) displayed superior wound healing via fresh wound strip breaking strength, fixed breaking strength, and collagen deposition (Barbul et al., 1985). In addition, mature rats fed diets supplemented with arginine and glycine had enhanced wound collagen deposition compared to controls (Chyun and Griminger, 1984).

Using a micromodel described by Goodson and Hunt that allows for studying fibroblastic responses and wound collagen deposition in human (Goodson and Hunt, 1982), 36 healthy human volunteers ages 25–35 years were given daily 17 g or 24.8 g of free arginine supplementation; this significantly increased the amount of hydroxyproline deposition at the wound site (an amino acid-specific for collagen) (Barbul et al., 1990). Subsequently, a study evaluated 30 volunteers >70 years old who were given daily supplements of 17 g of free arginine and demonstrated significantly enhanced wound collagen accumulation. Arginine supplementation had no effect on the rate of epithelialization of a superficial skin defect, indicating that the predominant effect is on wound collagen deposition (Kirk et al., 1993). A more recent study demonstrated that arginine, in combination with other supplements, increases human wound collagen deposition (Williams et al., 2002).

No single theory can account for the observed effects of arginine on wound healing. Several possible mechanisms have been postulated to explain the effects and provide a framework for understanding (Fig. 27.2).

ent Arginine supplementation provides a substrate for collagen synthesis at the wound site. Although free arginine makes up just a tiny amount of the collagen molecule, less than 5%, there could be utilization of arginine as substrate for proline through the following pathway:
arginineornithineglutamic semialdehydeprolineimage

ent Arginine levels are essentially nondetectable within the wound environment during the later phases of wound healing when fibroplasia predominates. While ornithine levels are higher in the wound than in the plasma, tracer isotope studies revealed that the rate of conversion of ornithine to proline in the wound is actually quite low, making this mechanism of arginine utilization unlikely (Albina et al., 2005).

ent Arginine induces collagen synthesis via a pituitary secretagogue mechanism. The beneficial effects of supplemental arginine on wound healing are in many respects similar to those of growth hormone—namely, enhanced wound breaking strength and collagen deposition (Herndon et al., 1990; Jørgensen and Andreassen, 1988; Kowalewski and Yong, 1968). In support of such a mechanism, it has been noted that the effect of arginine on wound healing is abrogated in hypophysectomized animals (Barbul et al., 1983). In addition, arginine supplementation in doses that increase the wound healing response also induces elevations in plasma insulin-like growth factor, the peripheral mediator of growth hormone (Kirk et al., 1993).

ent Arginine stimulates T-cell responses, thereby reducing the inhibitory effect of injury on T-cell function (Barbul et al., 1980a, b; Fabris and Mocchegiani, 1992). T lymphocytes are essential for normal wound healing as evidenced by decreased wound breaking strength in animals treated with monoclonal antibodies against T lymphocytes (Peterson et al., 1987). T lymphocytes are found immunohistochemically throughout the various phases of wound healing in distinct patterns (Fishel et al., 1987) and facilitating normal repair (Agaiby and Dyson, 1999).

ent Arginine is the unique substrate for NO. Several studies suggest that NO plays a critical role in wound healing. Exogenous NO administration has been shown to increase collagen synthesis in cultured dermal fibroblasts (Schäffer et al., 1997). In contrast, inhibitors of NO have been shown to significantly impair healing of cutaneous incisional and colonic anastomotic healing in rodents (Efron, 1999; Schäffer et al., 1999). In models of impaired healing such as diabetes, NO synthesis is impaired together with decreased collagen accumulation, while administration of NO restores wound healing responses toward normal (Shi et al., 2003). Transfection of iNOS DNA into wounds results in supraphysiologic collagen deposition (Thornton et al., 1998). Conversely, mice lacking the iNOS gene have delayed the closure of excisional wounds, an impairment that is remedied by adenoviral transfer of the iNOS gene to the wound bed (Yamasaki et al., 1998). Strongly supporting this mechanism of action are findings that arginine does not stimulate wound healing in iNOS knockout mice, suggesting that the iNOS pathway is at least partially responsible for the enhancement of wound healing observed with the administration of arginine (Shi et al., 2000).

image
Figure 27.2 Wound healing and arginine: postulated mechanisms of action.

Cardiovascular Health

Cardiovascular disease, specifically heart disease, is the leading cause of death in the United States, accounting for more than 600,000 deaths per year (Xu et al., 2016). The role of arginine in cardiovascular health has been extensively studied at the basic science, translational, and clinical levels. Given that cardiovascular disease is a condition of aging, such discussion has implications for the elderly population.

As mentioned, arginine is the precursor for the endogenous synthesis of NO (Palmer et al., 1988; Schmidt et al., 1988). This pathway, although it only accounts for a small portion of arginine metabolism, has attracted much attention given the role of NO in vascular physiology and pathophysiology (Böger et al., 1996; Epstein et al., 1993). Normal plasma arginine concentrations are 80–120 µM, and the Kmimage for arginine as substrate for NOS is ~1–10 µM (Brosnan, 2003; Wu et al., 2007). Given such, it appears that there should be a surplus of substrate. Despite this logic, several studies indicate that administration of exogenous arginine increases the generation of NO, a phenomenon now termed the arginine paradox. When examining vascular endothelial function in a lysinuric protein intolerant patient who had a defect in a dibasic amino acid transporter, there was noted impaired uptake of exogenous arginine, leading to a 79% lower plasma concentration. Assessment of the NO-dependent endothelial function revealed 70% lower serum levels of NO. Moreover, the study revealed decreased flow-mediated brachial artery vasodilator response, elevated plasma fibrin degradation products, increased thrombin–antithrombin III complex, and reduced circulating platelet count. Parenteral infusion of arginine reversed all of these effects (Kamada et al., 2001).

Generated NO is an important messenger molecule with diverse function. Animal models of human disease—including hypercholesterolemic rabbits (Böger et al., 1997; 1995; Cooke et al., 1991), hyperlipidemic monkeys (Quillen et al., 1991), and hypertensive rats (Lüscher and Vanhoutte, 1986)—have shown that the biologic functions of endothelium-derived NO are impaired, leading to dysregulation of endothelial control of vascular tone and blood flow. In all of these models, arginine supplementation helped restore NO production and improve endothelial dysfunction.

Translated to the human population, a multitude of positive results in the prevention and management of cardiovascular disease and its downstream organ system effects have been reported. Examples include the following:

ent 4–24 g/day of enteral arginine supplementation significantly lowered both systolic and diastolic blood pressure (Dong et al., 2011);

ent 6.6 g/day of enteral arginine supplementation in patients with peripheral arterial disease resulted in a 66% increase in pain-free walking distance, 23% increase in total walking distance, and improved emotional and social functions per SF-36—effects that were initially noted at two weeks and were sustained at 10 weeks (Maxwell et al., 2000);

ent 15 g/day of enteral arginine supplementation in patients with congestive heart failure led to improvement in glomerular filtration rate, natriuresis, and plasma endothelin levels at 5 days (Watanabe et al., 2000);

ent 6.6 g/day of enteral arginine supplementation in a patient with type 1 diabetes with debilitating exertional angina pectoris resulted in complete amelioration of angina and normalized exercise capacity at 7 days (Schwartz, 2003);

ent 8.4 g/day of enteral arginine supplementation in hypercholesterolemic humans showed modest attenuation of platelet reactivity and aggregation (Wolf et al., 1997); and

ent 17 g/day of enteral arginine supplementation in a healthy, nonsmoking, elderly population revealed a decreased serum total cholesterol level with a reduction in low-density lipoprotein cholesterol but not high-density lipoprotein cholesterol. Thus, the ratio of low- to high-density lipoprotein fraction was increased (Hurson et al., 1995).

Despite the evidence supporting the use of arginine for cardiovascular health, other studies have failed to show any benefit. In one randomized clinic trial, there was no improvement in vascular stiffness measurements or ejection fraction with the addition of arginine in the acute postinfarction period (Schulman et al., 2006). Moreover, in men with stable angina, oral supplementation with 15 g/day of arginine was not associated with improvement in endothelium-dependent vasodilation, oxidative stress, or exercise performance (Walker et al., 2001). Lastly, oral supplementation of 6 g/day of arginine in patients with coronary artery disease did not affect exercise-induced changes in QT-interval duration, QT dispersion, or the magnitude of ST-segment depression (Bednarz et al., 2000).

Though the clinical and therapeutic benefit of arginine supplementation on mitigating cardiovascular disease varies, arginine supplementation may be favorable in at-risk patient populations. The possible role of arginine in maintaining or improving cardiac function prior to onset of severe disease merits further study.

Neurologic Function

Neurodegeneration is characterized by progressive deterioration in cognitive ability and capacity for independent living. The arising clinical syndrome—dementia—is typically a condition that affects older people and leads to disability and dependence (Prince et al., 2013). Given the rising epidemic secondary to population aging, much research has been undertaken on the topic.

Given its role in NO production via neuronal NOS, the function of arginine supplementation on neurologic function has yet to be fully investigated. However, studies exploring the importance of NO on memory and learning via synaptic plasticity is growing (Böhme et al., 1991; Susswein et al., 2004; Zhou and Zhu, 2009). In vitro experiments have shown that NO serves to increase glutamate release and increase synaptic effectiveness, an occurrence termed potentiation that is thought to improve memory (Epstein et al., 1993). In vivo, the same physiology is likely present, for studies reveal that inhibiting NO impairs learning behavior (Chapman et al., 1992). A few investigations have explored the role of arginine supplementation on age-related degenerative neurologic disorders, specifically senile dementia and Alzheimer’s disease. The former study revealed improved cognitive function and reduced lipid peroxidation in 16 elderly patients following arginine supplementation (Ohtsuka and Nakaya, 2000). The latter concluded that arginine might provide a protective effect through the redox stress and inflammatory process and regulation of synaptic plasticity and neurogenesis (Yi et al., 2009).

Sexual Function

Sexual dysfunction is highly prevalent in both men and women and is a topic of particular importance to the elderly population because sexual dysfunction is clearly associated with advancing age (Laumann et al., 1999). Given the prevalence, a multitude of investigations have focused on sexual dysfunction, including a possible role for arginine supplementation. Thus far, there is no clear evidence of benefit. It is known that NO, through the synthesis of cyclic guanosine 5'-monophosphate, functions to initiate and maintain increased intracavernous pressure, penile vasodilatation, and penile erection. Thus, it is postulated that the production of NO via arginine, through the NOS pathway, may be of importance in sexual physiology.

In the 1950s, researchers first showed that adult men fed an arginine-deficient diet for nine days had an approximately 90% decreased sperm count and a 10-fold increase in the percentage of nonmotile sperm (Albanese, 1952). A subsequent study showed that supplementing infertile men with oral arginine for 6–8 weeks increased sperm counts and motility (Tanimura, 1967). The role of arginine in erectile dysfunction has also been studied. Studies in mice (Moody et al., 1997) and humans (Chen et al., 2001; Zorgniotti and Lizza, 1994) revealed improved erectile response and, in humans, subjective improvement in sexual function following arginine supplementation. A more robust study, however, utilizing a crossover design, failed to show benefit in arginine supplementation versus placebo in the management of impotence (Klotz et al., 1999). Despite the unclear benefit, arginine remains an active ingredient in most over-the-counter aphrodisiacs and sexual enhancement formulas.

Endocrinology

Endocrine disturbances are common in the aging population. A study in 2010 revealed that 9.6% of people 65 and older carried the diagnosis of diabetes mellitus (DM). An additional 9.3% of the population met clinical criteria for the disease but had not yet been diagnosed (Harris, 1990). Given such prevalence, the role of arginine on the endocrine system has been studied and will be discussed in the following.

Intravenous infusions of essential amino acids, specifically arginine, induce the release of insulin in humans (Floyd et al., 1966). Further, enteral intraduodenal arginine infusions result in sustained increases in plasma insulin levels (Dupre et al., 1969) with no observed hypoglycemic events. Later studies revealed that arginine supplementation also increases concentrations of glucagon (Palmer et al., 1976) and growth hormone (Merimee et al., 1965). The role of arginine in growth hormone secretion was further studied and revealed mixed results (Isidori et al., 1981). Arginine has also been found to stimulate the release of prolactin (Rakoff et al., 1973), pancreatic polypeptides, and somatostatin (Weir et al., 1978). Despite evidence of the endocrine-related actions of arginine, the underlying mechanism remains unclear. One proposed method is through arginine-derived NO (Schmidt et al., 1992).

Diabetes is associated with reduced levels of plasma arginine (Pieper et al., 1996) and elevated levels of the NOS inhibitor asymmetric dimethylarginine (Abbasi et al., 2001). Arginine supplementation may be an effective way to improve endothelial function and insulin sensitivity in DM (Giugliano et al., 1997; Wascher et al., 1997). Arginine also reduces microangiopathic complications of DM by counteracting lipid peroxidation (Lubec et al., 1997). A robust double-blind study revealed that oral administration of 9 g/day of arginine significantly improved peripheral and hepatic insulin sensitivity (Piatti et al., 2001). Again, the mechanisms remains unclear, but both NO-dependent (Schmidt et al., 1992) and NO-independent (Thams and Capito, 1999) pathways have been postulated to be operational.

Arginine Dosing and Supplementation

Throughout this chapter, we have discussed many putative effects of arginine based on the growing appreciation of arginine biochemistry and biology since the 1980s. The discovery of arginine as the precursor of nitric oxide led to increasing numbers of studies examining the potential effects in patients with coronary artery disease, peripheral artery disease, and other diseases that may benefit from increased NO production. However, many other conditions may also have therapeutic benefits, including wound healing and immunologic responses. Consumers have many over-the-counter supplements available to them, all of which are various formulations of arginine or arginine in combination with one or more other nutrients or compounds. Overall there is no good systematic data identifying which if any of these supplements might be beneficial. A number of studies examining oral arginine supplementation can be found in Table 27.1, which demonstrates the diversity in dosing and indications for arginine supplementation that have been examined in humans. When considering supplementation with arginine, an arginine-containing supplement, or any nutritional supplement, medical advice should always be obtained.

Table 27.1

Representative Studies Examining Oral Arginine Administration in Human Subjects on Clinical Endpoints

Study Year Subjects Time Course Total Daily Arginine Dose Endpoints
Rector et al. (1996) 1996 Heart failure (n=15); crossover 6 weeks 5.6–12.6 g Increased forearm blood flow
Wolf et al. (1997) 1997 Hypercholesterolemia (n=15 arginine, 8 placebo) 2 weeks 8.4 g Decreased platelet aggregation
Ceremużyński et al. (1997) 1997 Prior MI; stable angina (n=12 arginine, 10 placebo) 3 days 6 g Increased time to maximal ST depression during exercise; decreased maximal ST depression
Lerman et al., (1998) 1998 CAD (n=13 arginine, 13 placebo) 6 months 9 g Increased acetylcholine-mediated coronary blood flow; improved symptom scores
Blum et al. (2000) 2000 CAD patients (n=30); crossover 1 months 9 g Increased plasma arginine; no change in nitrogen oxides, flow-mediated brachial artery dilation, or vascular adhesion molecules
Mullen et al. (2000) 2000 Type 1 diabetes (ages 18–47; n=22 arginine, 22 placebo) 6 weeks 14 g No change in endothelial dysfunction
Maxwell et al. (2000) 2000 PAD (intermittent claudication) (n=12/15 arginine, 14 placebo) 8 weeks 0, 3.3, 6.6 g Increased pain-free walking and walking distance
Maxwell et al. (2002) 2002 CAD (chronic stale angina) (n=36); crossover 2 weeks 6 g (food sup) Improved flow-mediated arterial dilation, increased treadmill exercise time
Lekakis et al. (2002) 2002 Essential hypertension (n=18 arg, 17 placebo) 1.5 h 6 g Flow-mediated dilation of the brachial artery increased
Staff et al. (2004) 2004 Pre-eclampsia (n=15 arginine, 15 placebo) 5 days 12 g No change in diastolic blood pressure
Oka et al. (2005) 2005 PAD (n=18 placebo, 18/17/19 arginine); 12 weeks 0, 9, 18, 27 g Nonsignificant trends for increased speed and distance
VINTAGE MI Trial* (Schulman et al., 2006) 2006 60 years or older (n=78 arginine, 75 placebo); confirmed STEMI 6 months 9 g No effect on vascular stiffness or ejection fraction; *trial halted for increased mortality in arginine group
NO-PAIN Study (Wilson et al., 2007) 2007 PAD; intermittent claudication (n =67 placebo, 66 arginine) 6 months 3 g No changes in nitrogen oxides, vascular reactivity, or walking distance
Ruel et al. (2008) 2008 Surgical three-vessel CAD (n=19 total, divided into 4 groups) 3 months 6 g vs. placebo (all patients receiving VEGF intraop) Increased myocardial perfusion and contractility
Lucotti et al. (2009) 2009 Nondiabetic CAD (n=32 arginine, 32 placebo) 6 months 6.4 g Increased insulin sensitivity; decreased IL-6, MCP-1

Image

CAD, coronary arterial disease; PAD, peripheral arterial disease; STEMI, ST-elevation myocardial infarction; MI, myocardial infarction; MCP-1, monocyte chemotactic protein-1; IL-6, interleukin-6; VEGF, vasculoendothelial growth factor; VINTAGE MI, Vascular Interaction with Age in Myocardial Infarction.

Pharmacokinetics of Arginine and Citrulline

The biologic effects of arginine appear to be influenced by its plasma concentration, and several studies have made kinetic measurements following single or multiple oral doses (Bode-Böger et al., 1998). More recently, the pharmacokinetics of citrulline have been investigated given the close relationship of arginine and citrulline (i.e., the citrulline–NO pathway), as well as the observation that citrulline increases circulating arginine via endogenous synthesis. For example, Moinard and colleagues found that citrulline supplementation increased plasma arginine concentrations without altering urea or urine urea nitrogen excretion, thus increasing nitrogen balance (Rougé et al., 2007).

The average plasma concentration of arginine in fasted humans is approximately 80–100 µM (Bode-Böger et al., 1998; Cynober, 2002; Tangphao et al., 1999), while citrulline concentration is significantly less at approximately 40 µM (Cynober, 2002). Two particular studies have reported both 6 g (Bode-Böger et al., 1998) or 10 g (Tangphao et al., 1999) oral supplementation of arginine leading to maximal concentrations of approximately 300 µM that are reached 60–90 min following ingestion. Maximal concentration is probably in this range for any single tolerated dose, given that increasing arginine concentration appears to plateau at greater administered amounts. When comparing arginine ingestion and citrulline ingestion at increasing doses, citrulline is actually more effective at increasing arginine concentration compared to supplemental arginine itself (Schwedhelm et al., 2008). As one might expect, this is consistent with the endogenous pathway of arginine synthesis mentioned previously in this chapter, which involves intestinal release of citrulline that completely bypasses the liver and is transported to the kidney, where citrulline is used for endogenous synthesis of arginine (Fig. 27.1) (Crenn et al., 2008).

The oral bioavailability of a single oral dose of arginine appears to plateau prior to a 6 g administration, as 6 g orally is associated with approximately 70% bioavailability (Bode-Böger et al., 1998) while 10 g appears to saturate transport and be only 20% bioavailable (Tangphao et al., 1999). In contrast to the bioavailability of arginine, however, citrulline bioavailability appears markedly increased, though there appears to be a lack of studies that fully evaluate the kinetics (Cynober, 2007). Regardless, no experiments appear to have used sufficient doses to saturate citrulline absorption by the gastrointestinal tract. The interesting finding is that citrulline lacks these common gastrointestinal (GI) side effects from similarly structural amino acids, indicating that the transport of citrulline does not appear to be limited (Breuillard et al., 2015; Schwedhelm et al., 2008).

Longevity of supplementation and risk of exposure to excess arginine or citrulline and its effect on long-term maintenance of metabolite concentrations remains unknown. It is important to note that most studies examine either acute (single dose) or subacute (days to weeks of treatment) dosing regimens and not chronic dosing (Schwedhelm et al., 2008). It is unknown how these different lengths of exposure affect the expression of metabolizing enzymes or other compensatory mechanisms given the increased arginine or citrulline delivered via supplementation. Also, it has yet to be elucidated whether the biologic effect could wane or subside with time or prolonged exposure.

Practical Considerations

There are several practical considerations regarding arginine supplementation. The obvious hurdle to overcome with increasingly larger doses of oral arginine supplementation is that arginine has an unpleasant taste, at least at doses of >7 g orally at a single administration (Evans et al., 2004). Larger doses (approximately 9 g per single oral dose and greater) tend to be poorly tolerated in terms of gastrointestinal cramping, bloating, and diarrhea. These GI effects are likely related to the osmotic effects of slow enterocyte uptake (Grimble, 2007), but there are data to suggest that these may be mediated by local increases in NO concentration in the enterocyte milieu as well. Note that supplementation is typically on a background dietary intake of arginine that typically ranges from 2.5 to 5 g/day (Gad, 2010). Despite previous studies reporting supplemental doses of up to 30 g/day (in split doses) being generally well tolerated, more recent studies suggest that nausea and diarrhea become nearly universal when single dosing is escalated—especially amounts of arginine from 15 to 30 g per single oral dose (Gad, 2010; Moinard et al., 2008).

As mentioned, the role of citrulline in arginine homeostasis has been studied, especially whether or not citrulline might be a more effective supplement compared to arginine. Citrulline is better tolerated based on the absence of reported gastrointestinal side effects (Moinard et al., 2008). Oral citrulline increases nitrogen balance and plasma arginine concentrations (Rougé et al., 2007). In fact, when the appearance of both citrulline and arginine in the plasma are measured following oral citrulline dosing, plasma arginine concentration increases in response to citrulline more than arginine supplementation. These results are consistent with the conversion of citrulline to arginine in the kidney being a saturable process (Moinard et al., 2008).

As previously mentioned, the GI side effects of arginine supplementation further make the potential for citrulline more attractive as a long-term, high-dose supplement. The intestinal dibasic amino acid transporter responsible for arginine absorption (as well as other amino acids) is a high-affinity and low-capacity system (Grimble, 2007). Arginine, ornithine, citrulline, and cystine share the same enterocyte transporter (Grimble, 2007), but citrulline remains free from many of these side effects associated with significant arginine intake. What is particularly interesting and currently unknown is why citrulline, but not the other metabolites sharing this transporter, lacks GI side effects.

Arginine Dosing and Aging

Whether or not aging affects how arginine dosing should or can be administered is not well studied. Moreover, the exact mechanisms of how aging affects the absorption of nutrients in general is not completely understood in human or animal models, and many questions remain (Holt, 2007; Keller and Layer, 2014; Schiller, 2009). Aging has a range of effects on macro- and micronutrient absorption, and it is likely that these effects could significantly vary between individuals. Clearly, many questions need to be answered about arginine metabolism in general as well as potentially innovative ways identified to take advantage of biology to augment endogenous responses to amino acids for therapeutic purposes. Could citrulline supplementation be a reasonable way to augment endogenous arginine availability? Could citrulline be advantageous in aged individuals or others with intrinsic intestinal pathology leading to poor GI absorption? These as well as many other questions are important but currently remain unanswered.

References

1. Abbasi F, Asagmi T, Cooke JP, et al. Plasma concentrations of asymmetric dimethylarginine are increased in patients with type 2 diabetes mellitus. Am J Cardiol. 2001;88:1201–1203 <http://dx.doi.org/10.1016/S0002-9149(01)02063-X.

2. Agaiby A, Dyson M. Immuno-inflammatory cell dynamics during cutaneous wound healing. J Anat. 1999;195:531–542 http://dx.doi.org/10.1046/j.1469-7580.1999.19540531.x.

3. Agarwal S, Busse PJ. Innate and adaptive immunosenescence. Ann Allergy Asthma Immunol. 2010;104:183–190 http://dx.doi.org/10.1016/j.anai.2009.11.009.

4. Albanese AA. The effects of amino acid deficiencies in man. Am J Clin Nutr. 1952;1:44–51.

5. Albina JE, Mahoney EJ, Daley JM, Wesche DE, Morris Jr SM, Reichner JS. Macrophage arginase regulation by CCAAT/Enhancer-binding protein beta. Shock. 2005;23:168–172 http://dx.doi.org/10.1097/01.shk.0000148054.74268.e2.

6. Ban H, Shigemitsu K, Yamatsuji T, et al. Arginine and Leucine regulate p70 S6 kinase and 4E-BP1 in intestinal epithelial cells. Int J Mol Med. 2004;13:537–543 http://dx.doi.org/10.3892/ijmm.13.4.537.

7. Barbul A, Wasserkrug HL, Seifter E, Rettura G, Levenson SM, Efron G. Immunostimulatory effects of arginine in normal and injured rats. J Surg Res. 1980a;29:228–235 <http://dx.doi.org/10.1016/0022-4804(80)90165-1>.

8. Barbul A, Wasserkrug HL, Sisto DA, et al. Thymic stimulatory actions of arginine. JPEN J Parenter Enteral Nutr. 1980b;4:446–449 http://dx.doi.org/10.1177/014860718000400502.

9. Barbul A, Rettura G, Levenson SM, Seifter E. Wound healing and thymotropic effects of arginine: a pituitary mechanism of action. Am J Clin Nutr. 1983;37:786–794.

10. Barbul A, Fishel RS, Shimazu S, et al. Intravenous hyperalimentation with high arginine levels improves wound healing and immune function. J Surg Res. 1985;38:328–334 <http://dx.doi.org/10.1016/0022-4804(85)90045-9>.

11. Barbul A, Lazarou SA, Efron DT, Wasserkrug HL, Efron G. Arginine enhances wound healing and lymphocyte immune responses in humans. Surgery. 1990;108:336–337 331–6– discussion.

12. Bauchart-Thevret C, Cui L, Wu G, Burrin DG. Arginine-induced stimulation of protein synthesis and survival in IPEC-J2 cells is mediated by mTOR but not nitric oxide. AJP: Endocrinol Metab. 2010;299:E899–E909 http://dx.doi.org/10.1152/ajpendo.00068.2010.

13. Bednarz B, Wolk R, Chamiec T, Herbaczyńska-Cedro K, Winek D, Ceremużyński L. Effects of oral l-arginine supplementation on exercise-induced QT dispersion and exercise tolerance in stable angina pectoris. Int J Cardiol. 2000;75:205–210 <http://dx.doi.org/10.1016/S0167-5273(00)00324-7>.

14. Bernstein LH. The impact of immunonutrition. Gastroenterology. 1996;110:1677–1678 http://dx.doi.org/10.1053/gast.1996.v110.agast961677.

15. Blum A, Hathaway L, Mincemoyer R, et al. Oral L-arginine in patients with coronary artery disease on medical management. Circulation. 2000;101:2160–2164.

16. Bode-Böger SM, Böger RH, Galland A, Tsikas D, Frölich JC. L‐arginine‐induced vasodilation in healthy humans: pharmacokinetic–pharmacodynamic relationship. Br J Clin Pharmacol. 1998;46:489–497 http://dx.doi.org/10.1046/j.1365-2125.1998.00803.x.

17. Böger RH, Bode-Böger SM, Mügge A, et al. Supplementation of hypercholesterolaemic rabbits with L-arginine reduces the vascular release of superoxide anions and restores NO production. Atherosclerosis. 1995;117:273–284 <http://dx.doi.org/10.1016/0021-9150(95)05582-H>.

18. Böger RH, Bode-Böger SM, Frölich JC. The l-arginine—nitric oxide pathway: role in atherosclerosis and therapeutic implications. Atherosclerosis. 1996;127:1–11 <http://dx.doi.org/10.1016/S0021-9150(96)05953-9>.

19. Böger RH, Bode-Böger SM, Brandes RP, et al. Dietary L-arginine reduces the progression of atherosclerosis in cholesterol-fed rabbits: comparison with lovastatin. Circulation. 1997;96:1282–1290 http://dx.doi.org/10.1161/01.CIR.96.4.1282.

20. Böhme GA, Bon C, Stutzmann J-M, Doble A, Blanchard J-C. Possible involvement of nitric oxide in long-term potentiation. Eur J Pharmacol. 1991;199:379–381 <http://dx.doi.org/10.1016/0014-2999(91)90505-K>.

21. Breuillard C, Cynober L, Moinard C. Citrulline and nitrogen homeostasis: an overview. Amino Acids. 2015;47:685–691 http://dx.doi.org/10.1007/s00726-015-1932-2.

22. Brittenden J, Heys SD, Ross J, Park KGM, Eremin O. Nutritional pharmacology: effects of l-arginine on host defences, response to trauma and tumor growth. Clin Sci. 1994;86:123–132 http://dx.doi.org/10.1042/cs0860123.

23. Bronstein-Sitton N, Wang L, Cohen L, Baniyash M. Expression of the T cell antigen receptor zeta chain following activation is controlled at distinct checkpoints Implications for cell surface receptor down-modulation and re-expression. J Biol Chem. 1999;274:23659–23665 http://dx.doi.org/10.1074/jbc.274.33.23659.

24. Bronte V, Zanovello P. Regulation of immune responses by L-arginine metabolism. Nat Rev Immunol. 2005;5:641–654 http://dx.doi.org/10.1038/nri1668.

25. Brosnan JT. Interorgan amino acid transport and its regulation. J Nutr. 2003;133:2068S–2072S.

26. Brosnan JT, Brosnan ME. Creatine metabolism and the urea cycle. Mol Genet Metab. 2010;100:S49–S52 http://dx.doi.org/10.1016/j.ymgme.2010.02.020.

27. Callery MP, Mangino MJ, Flye MW. Arginine-specific suppression of mixed lymphocyte culture reactivity by kupffer cells - a basis of portal venous tolerance. Transplantation. 1991;51:1076–1079 http://dx.doi.org/10.1097/00007890-199105000-00028.

28. Carraway MS, Piantadosi CA, Jenkinson CP, Huang Y-CT. Differential expression of arginase and iNOS in the lung in sepsis. Exp Lung Res. 2015;24:253–268 http://dx.doi.org/10.3109/01902149809041533.

29. Castillo L. Whole body nitric oxide synthesis in healthy men determined from [15N]arginine-to-[15N]citrulline labeling. Proc Natl Acad Sci USA. 1996;93:11460–11465.

30. Castillo L, Chapman TE, Yu YM, Ajami A, Burke JF, Young VR. Dietary arginine uptake by the splanchnic region in adult humans. Am J Physiol. 1993;265:E532–E539.

31. Castillo L, Sanchez M, Vogt J, et al. Plasma arginine, citrulline, and ornithine kinetics in adults, with observations on nitric oxide synthesis. Am J Physiol. 1995;268:E360–E367.

32. Ceremużyński L, Chamiec T, Herbaczyńska-Cedro K. Effect of Supplemental Oral l-Arginine on Exercise Capacity in Patients With Stable Angina Pectoris. Am J Cardiol. 1997;80:331–333 <http://dx.doi.org/10.1016/S0002-9149(97)00354-8>.

33. Chapman PF, Atkins CM, Allen MT, Haley JE, Steinmetz JE. Inhibition of nitric oxide synthesis impairs two different forms of learning. Neuroreport. 1992;3:567–570 http://dx.doi.org/10.1097/00001756-199207000-00005.

34. Chen W, Chernichovsky I, Sofer M. Effect of oral administration of high-dose nitric oxide donor l-arginine in men with organic erectile dysfunction: results of a double-blind, randomized, placebo-controlled study. BJU Int. 2001;83:269–273 http://dx.doi.org/10.1046/j.1464-410x.1999.00906.x.

35. Chyun JH, Griminger P. Improvement of nitrogen retention by arginine and glycine supplementation and its relation to collagen synthesis in traumatized mature and aged rats. J Nutr. 1984;114:1697–1704.

36. Clark R, Feleke G, Din M, et al. Nutritional treatment for acquired immunodeficiency virus-associated wasting using beta-hydroxy beta-methylbutyrate, glutamine, and arginine: a randomized, double-blind, placebo-controlled study. JPEN J Parenter Enteral Nutr. 2000;24:133–139 http://dx.doi.org/10.1177/0148607100024003133.

37. Cooke JP, Andon NA, Girerd XJ, Hirsch AT, Creager MA. Arginine restores cholinergic relaxation of hypercholesterolemic rabbit thoracic aorta. Circulation. 1991;83:1057–1062 http://dx.doi.org/10.1161/01.CIR.83.3.1057.

38. Corl BA, Odle J, Niu X, et al. Arginine activates intestinal p70(S6k) and protein synthesis in piglet rotavirus enteritis. J Nutr. 2008;138:24–29.

39. Crenn P, Messing B, Cynober L. Citrulline as a biomarker of intestinal failure due to enterocyte mass reductionent. Clin Nutr. 2008;27:328–339 http://dx.doi.org/10.1016/j.clnu.2008.02.005.

40. Curis E, Nicolis I, Moinard C, et al. Almost all about citrulline in mammals. Amino acids. 2005;29:177–205 http://dx.doi.org/10.1007/s00726-005-0235-4.

41. Curis E, Crenn P, Cynober L. Citrulline and the gut. Curr Opin Clin Nutr Metab Care. 2007;10:620–626 http://dx.doi.org/10.1097/MCO.0b013e32829fb38d.

42. Cynober LA. Plasma amino acid levels with a note on membrane transport: characteristics, regulation, and metabolic significance. Nutrition. 2002;18:761–766 <http://dx.doi.org/10.1016/S0899-9007(02)00780-3>.

43. Cynober L. Pharmacokinetics of arginine and related amino acids. J Nutr. 2007;137:1646S–1649S.

44. Daly JM, Reynolds J, Thom A, et al. Immune and metabolic effects of arginine in the surgical patient. Ann Surg. 1988;208:512–523.

45. de Jonge WJ, Hallemeesch MM, Kwikkers KL, et al. Overexpression of arginase I in enterocytes of transgenic mice elicits a selective arginine deficiency and affects skin, muscle, and lymphoid development. Am J Clin Nutr. 2002a;76:128–140 http://dx.doi.org/10.1203/00006450-199804000-00002.

46. de Jonge WJ, Kwikkers KL, Velde te AA, et al. Arginine deficiency affects early B cell maturation and lymphoid organ development in transgenic mice. J Clin Investig. 2002b;110:1539–1548 http://dx.doi.org/10.1172/JCI16143.

47. Dejong CH, Welters CF, Deutz NE, Heineman E, Soeters PB. Renal arginine metabolism in fasted rats with subacute short bowel syndrome. Clin Sci. 1998;95:409–418 http://dx.doi.org/10.1042/cs0950409.

48. Devés R, Boyd CA. Transporters for cationic amino acids in animal cells: discovery, structure, and function. Physiol Rev. 1998;78:487–545 <http://dx.doi.org/10.1016/0022-4804(91)90210-D>.

49. Dong J-Y, Qin L-Q, Zhang Z, et al. Effect of oral l-arginine supplementation on blood pressure: a meta-analysis of randomized, double-blind, placebo-controlled trials. Am Heart J. 2011;162:959–965 http://dx.doi.org/10.1016/j.ahj.2011.09.012.

50. Dupre J, Curtis JD, Unger RH, Waddell RW, Beck JC. Effects of secretin, pancreozymin, or gastrin on the response of the endocrine pancreas to administration of glucose or arginine in man. J Clin Investig. 1969;48:745–757 http://dx.doi.org/10.1172/JCI106032.

51. Durante W. Role of Arginase in Vessel Wall Remodeling. Front Immunol 2013;4 http://dx.doi.org/10.3389/fimmu.2013.00111.

52. Durante W, Johnson FK, Johnson RA. Arginase: a critical regulator of nitric oxide synthesis and vascular function. Clin Exp Pharmacol Physiol. 2007;34:906–911 http://dx.doi.org/10.1111/j.1440-1681.2007.04638.x.

53. Efron D. Expression and function of inducible nitric oxide synthase during rat colon anastomotic healing. J Gastrointest Surg. 1999;3:592–601 <http://dx.doi.org/10.1016/S1091-255X(99)80080-8>.

54. Efron DT, Kirk SJ, Regan MC, Wasserkrug HL, Barbul A. Nitric oxide generation from L-arginine is required for optimal human peripheral blood lymphocyte DNA synthesis. Surgery. 1991;110:327–334.

55. Epstein FH, Moncada S, Higgs A. The L-Arginine-Nitric Oxide Pathway. N Engl J Med. 1993;329:2002–2012 http://dx.doi.org/10.1056/NEJM199312303292706.

56. Evans RW, Fernstrom JD, Thompson J, Morris Jr SM, Kuller LH. Biochemical responses of healthy subjects during dietary supplementation with L-arginine. J Nutr Biochem. 2004;15:534–539 http://dx.doi.org/10.1016/j.jnutbio.2004.03.005.

57. Fabris N, Mocchegiani E. Arginine-containing compounds and thymic endocrine activity. Thymus. 1992;19(Suppl 1):S21–S30.

58. Fafournoux P, Bruhat A, Jousse C. Amino acid regulation of gene expression. Biochem J. 2000;351:1–12.

59. Feldmeyer N, Wabnitz G, Leicht S, et al. Arginine deficiency leads to impaired cofilin dephosphorylation in activated human T lymphocytes. Int Immunol. 2012;24:303–313 http://dx.doi.org/10.1093/intimm/dxs004.

60. Fishel RS, Barbul A, Beschorner WE, Wasserkrug HL, Efron G. Lymphocyte participation in wound healing Morphologic assessment using monoclonal antibodies. Ann Surg. 1987;206:25–29.

61. Flakoll P, Sharp R, Baier S, Levenhagen D, Carr C, Nissen S. Effect of β-hydroxy-β-methylbutyrate, arginine, and lysine supplementation on strength, functionality, body composition, and protein metabolism in elderly women. Nutrition. 2004;20:445–451 http://dx.doi.org/10.1016/j.nut.2004.01.009.

62. Floyd Jr JC, Fajans SS, Conn JW, Knopf RF, Rull J. Stimulation of insulin secretion by amino acids. J Clin Investig. 1966;45:1487–1502 http://dx.doi.org/10.1172/JCI105456.

63. Gad MZ. Anti-aging effects of l-arginine. J Adv Res. 2010;1:169–177 http://dx.doi.org/10.1016/j.jare.2010.05.001.

64. Giugliano D, Marfella R, Verrazzo G, et al. L-arginine for testing endothelium-dependent vascular functions in health and disease. Am J Physiol. 1997;273:E606–E612.

65. Goodson III WH, Hunt TK. Development of a new miniature method for the study of wound healing in human subjects. J Surg Res. 1982;33:394–401 <http://dx.doi.org/10.1016/0022-4804(82)90054-3>.

66. Grimble GK. Adverse gastrointestinal effects of arginine and related amino acids. J Nutr. 2007;137:1693S–1701S.

67. Haraguchi Y, Takiguchi M, Amaya Y, Kawamoto S, Matsuda I, Mori M. Molecular cloning and nucleotide sequence of cDNA for human liver arginase. Proc Natl Acad Sci. 1987;84:412–415.

68. Harris MI. Epidemiology of diabetes mellitus among the elderly in the United States. Clin Geriatr Med. 1990;6:703–719.

69. Häussinger D. Regulation of hepatic ammonia metabolism: the intercellular glutamine cycle. Adv Enzyme Regul. 1986;25:159–180 <http://dx.doi.org/10.1016/0065-2571(86)90013-0>.

70. Häussinger D, Gerok W, Sies H. Hepatic role in pH regulation: role of the intercellular glutamine cycle. Trends Biochem Sci. 1984;9:300–302 <http://dx.doi.org/10.1016/0968-0004(84)90294-9>.

71. Herndon DN, Barrow RE, Kunkel KR, Broemeling L, Rutan RL. Effects of recombinant human growth hormone on donor-site healing in severely burned children. Ann Surg. 1990;212:424–429 – discussion 430–1.

72. Holowatz LA, Kenney WL. Up‐regulation of arginase activity contributes to attenuated reflex cutaneous vasodilatation in hypertensive humans. J Physiol. 2007;581:863–872 http://dx.doi.org/10.1113/jphysiol.2007.128959.

73. Holt PR. Intestinal malabsorption in the elderly. Dig Dis. 2007;25:144–150 http://dx.doi.org/10.1159/000099479.

74. Hu FB, Stampfer MJ, Manson JE, et al. Frequent nut consumption and risk of coronary heart disease in women: prospective cohort study. BMJ. 1998;317:1341–1345 http://dx.doi.org/10.1136/bmj.317.7169.1341.

75. Hurson M, Regan MC, Kirk SJ, Wasserkrug HL, Barbul A. Metabolic effects of arginine in a healthy elderly population. JPEN J Parenter Enteral Nutr. 1995;19:227–230 http://dx.doi.org/10.1177/0148607195019003227.

76. Ikemoto M, Tsunekawa S, Tanaka K, et al. Liver-type arginase in serum during and after liver transplantation: a novel index in monitoring conditions of the liver graft and its clinical significance. Clin Chim Acta. 1998;271:11–23 <http://dx.doi.org/10.1016/S0009-8981(97)00226-X>.

77. Isidori A, Monaco Lo, A, Cappa M. A study of growth hormone release in man after oral administration of amino acids. Curr Med Res Opin. 1981;7:475–481 http://dx.doi.org/10.1185/03007998109114287.

78. Iyer RK, Kim HK, Tsoa RW, Grody WW, Cederbaum SD. Cloning and characterization of human agmatinase. Mol Genet Metab. 2002;75:209–218 http://dx.doi.org/10.1006/mgme.2001.3277.

79. Johnson FK, Johnson RA, Peyton KJ, Durante W. Arginase inhibition restores arteriolar endothelial function in Dahl rats with salt-induced hypertension. AJP: Regul Integr Comp Physiol. 2005;288:R1057–R1062 http://dx.doi.org/10.1152/ajpregu.00758.2004.

80. Jørgensen PH, Andreassen TT. Influence of biosynthetic human growth hormone on biomechanical properties of rat skin incisional wounds. Acta Chir Scand. 1988;154:623–626.

81. Kamada Y, Nagaretani H, Tamura S, et al. Vascular endothelial dysfunction resulting from l-arginine deficiency in a patient with lysinuric protein intolerance. J Clin Investig. 2001;108:717–724 http://dx.doi.org/10.1172/JCI11260.

82. Keller J, Layer P. The Pathophysiology of Malabsorption. Visc Med. 2014;30:150–154 http://dx.doi.org/10.1159/000364794.

83. Kim JH, Bugaj LJ, Oh YJ, et al. Arginase inhibition restores NOS coupling and reverses endothelial dysfunction and vascular stiffness in old rats. J Appl Physiol. 2009;107:1249–1257 http://dx.doi.org/10.1152/japplphysiol.91393.2008.

84. King DE, Mainous III AG, Geesey ME. Variation in L-arginine intake follow demographics and lifestyle factors that may impact cardiovascular disease risk. Nutr Res. 2008;28:21–24 http://dx.doi.org/10.1016/j.nutres.2007.11.003.

85. Kirk SJ, Regan MC, Wasserkrug HL, Sodeyama M, Barbul A. Arginine enhances T-Cell responses in athymic nude mice. JPEN J Parenter Enteral Nutr. 1992;16:429–432 http://dx.doi.org/10.1177/0148607192016005429.

86. Kirk SJ, Hurson M, Regan MC, Holt DR, Wasserkrug HL, Barbul A. Arginine stimulates wound healing and immune function in elderly human beings. Surgery. 1993;114:155–159 – discussion 160.

87. Klotz T, Mathers MJ, Braun M, Bloch W, Engelmann U. Effectiveness of oral L-arginine in first-line treatment of erectile dysfunction in a controlled crossover study. Urol Int. 1999;63:220–223 http://dx.doi.org/10.1159/000030454.

88. Kowalewski K, Yong S. Effect of growth hormone and an anabolic steroid on hydroxyproline in healing dermal wounds in rats. Eur J Endocrinol. 1968;59:53–66 http://dx.doi.org/10.1530/acta.0.0590053.

89. Laumann EO, Paik A, Rosen RC. Sexual dysfunction in the United States: prevalence and predictors. JAMA. 1999;281:537–544 http://dx.doi.org/10.1001/jama.281.6.537.

90. Längle F, Roth E, Steininger R, Winkler S, Mühlbacher F. Arginase release following liver reperfusion Evidence of hemodynamic action of arginase infusions. Transplantation. 1995;59:1542–1549.

91. LeBien TW. Arginine: an unusual dietary requirement of pre-B lymphocytes? J Clin Investig. 2002;110:1411–1413 http://dx.doi.org/10.1172/JCI17210.

92. Lekakis JP, Papathanassiou S, Papaioannou TG, et al. Oral l-arginine improves endothelial dysfunction in patients with essential hypertension. Int J Cardiol. 2002;86:317–323 <http://dx.doi.org/10.1016/S0167-5273(02)00413-8>.

93. Lerman A, Burnett JC, Higano ST, McKinley LJ, Holmes DR. Long-term L-arginine supplementation improves small-vessel coronary endothelial function in humans. Circulation. 1998;97:2123–2128 http://dx.doi.org/10.1161/01.CIR.97.21.2123.

94. Li P, Yin Y-L, Li D, Kim SW, Wu G. Amino acids and immune function. Br J Nutr. 2007;98:237–252 http://dx.doi.org/10.1017/S000711450769936X.

95. Lortie MJ, Novotny WF, Peterson OW, et al. Agmatine, a bioactive metabolite of arginine Production, degradation, and functional effects in the kidney of the rat. J Clin Investig. 1996;97:413–420 http://dx.doi.org/10.1172/JCI118430.

96. Lubec B, Hayn M, Kitzmüller E, Vierhapper H, Lubec G. l-Arginine reduces lipid peroxidation in patients with diabetes mellitus. Free Radic Biol Med. 1997;22:355–357 <http://dx.doi.org/10.1016/S0891-5849(96)00386-3>.

97. Lucotti P, Monti L, Setola E, et al. Oral l-arginine supplementation improves endothelial function and ameliorates insulin sensitivity and inflammation in cardiopathic nondiabetic patients after an aortocoronary bypass. Metabolism. 2009;58:1270–1276 http://dx.doi.org/10.1016/j.metabol.2009.03.029.

98. Luiking YC, Have, Ten GAM, Wolfe RR, Deutz NEP. Arginine de novo and nitric oxide production in disease states. AJP: Endocrinol Metab. 2012;303:E1177–E1189 http://dx.doi.org/10.1152/ajpendo.00284.2012.

99. Lüscher TF, Vanhoutte PM. Endothelium-dependent contractions to acetylcholine in the aorta of the spontaneously hypertensive rat. Hypertension. 1986;8:344–348 http://dx.doi.org/10.1161/01.HYP.8.4.344.

100. Marini JC. Arginine and ornithine are the main precursors for citrulline synthesis in mice. J Nutr. 2012;142:572–580 http://dx.doi.org/10.3945/jn.111.153825.

101. Maxwell AJ, Anderson BE, Cooke JP. Nutritional therapy for peripheral arterial disease: a double-blind, placebo-controlled, randomized trial of HeartBar®. Vasc Med. 2000;5:11–19 http://dx.doi.org/10.1177/1358836X0000500103.

102. Maxwell AJ, Zapien MP, Pearce GL, MacCallum G, Stone PH. Randomized trial of a medical food for the dietary management of chronic, stable angina. J Am Coll Cardiol. 2002;39:37–45 <http://dx.doi.org/10.1016/S0735-1097(01)01708-9>.

103. May PE, Barber A, D’Olimpio JT, Hourihane A, Abumrad NN. Reversal of cancer-related wasting using oral supplementation with a combination of β-hydroxy-β-methylbutyrate, arginine, and glutamine. Am J Surg. 2002;183:471–479 <http://dx.doi.org/10.1016/S0002-9610(02)00823-1>.

104. Merimee T, Lillicrap D, Rabinowitz D. Effect of arginine on serum-levels of human growth-hormone. Lancet. 1965;286:668–670 <http://dx.doi.org/10.1016/S0140-6736(65)90399-5>.

105. Minami Y, Weissman AM, Samelson LE, Klausner RD. Building a multichain receptor: synthesis, degradation, and assembly of the T-cell antigen receptor. Proc Natl Acad Sci USA. 1987;84:2688–2692.

106. Moinard C, Nicolis I, Neveux N, Darquy S, Bénazeth S, Cynober L. Dose-ranging effects of citrulline administration on plasma amino acids and hormonal patterns in healthy subjects: the Citrudose pharmacokinetic study. Br J Nutr. 2008;99:855–862 http://dx.doi.org/10.1017/S0007114507841110.

107. Moinard C, Maccario J, Walrand S, et al. Arginine behavior after arginine or citrulline administration in older subjects. Br J Nutr 2015;1–6 http://dx.doi.org/10.1017/S0007114515004638.

108. Moody JA, Vernet D, Laidlaw S, Rajfer J, Gonzalez-Cadavid NF. Effects of long-term oral administration of L-Arginine on the rat erectile response. J Urol. 1997;158:942–947 <http://dx.doi.org/10.1016/S0022-5347(01)64368-4>.

109. Mori M, Gotoh T. Regulation of nitric oxide production by arginine metabolic enzymes. Biochem Biophys Res Commun. 2000;275:715–719 http://dx.doi.org/10.1006/bbrc.2000.3169.

110. Moriguchi S, Mukai K, Hiraoka I, Kishino Y. Functional changes in human lymphocytes and monocytes after in vitro incubation with arginine. Nutr Res. 1987; <http://dx.doi.org/10.1016/S0271-5317(87)80102-1>.

111. Moriguti JC, Ferriolli E, Donadi EA, Marchini JS. Effects of arginine supplementation on the humoral and innate immune response of older people. Eur J Clin Nutr. 2005;59:1362–1366 http://dx.doi.org/10.1038/sj.ejcn.1602247.

112. Morris SM. Regulation of enzymes of the urea cycle and arginine metabolism. Annu Rev Nutr. 2002;22:87–105 http://dx.doi.org/10.1146/annurev.nutr.22.110801.140547.

113. Morris SM. Enzymes of arginine metabolism. J Nutr. 2004;134:2743S–2747S – discussion 2765S–2767S.

114. Morris Jr SM. Recent advances in arginine metabolism: roles and regulation of the arginases. Br J Pharmacol. 2009;157:922–930 http://dx.doi.org/10.1111/j.1476-5381.2009.00278.x.

115. Morrissey J, McCracken R, Ishidoya S, Klahr S. Partial cloning and characterization of an arginine decarboxylase in the kidney. Kidney Int. 1995;47:1458–1461 http://dx.doi.org/10.1038/ki.1995.204.

116. Mullen MJ, Wright D, Donald AE, Thorne S, Thomson H, Deanfield JE. Atorvastatin but not l-arginine improves endothelial function in type I diabetes mellitus: a double-blind study. J Am Coll Cardiol. 2000;36:410–416 <http://dx.doi.org/10.1016/S0735-1097(00)00743-9>.

117. Munder M, Eichmann K, Modolell M. Alternative metabolic states in murine macrophages reflected by the nitric oxide synthase/arginase balance: competitive regulation by CD4+ T cells correlates with Th1/Th2 phenotype. J Immunol. 1998;160:5347–5354 <http://dx.doi.org/10.1016/0167-5699(96)80606-2>.

118. Nakagawa T, Guarente L. Urea cycle regulation by mitochondrial sirtuin, SIRT5. Aging (Albany NY). 2009;1:578–581.

119. Nakagawa T, Lomb DJ, Haigis MC, Guarente L. SIRT5 deacetylates carbamoyl phosphate synthetase 1 and regulates the urea cycle. Cell. 2009;137:560–570 http://dx.doi.org/10.1016/j.cell.2009.02.026.

120. Ochoa JB, Udekwu AO, Billiar TR, et al. Nitrogen oxide levels in patients after trauma and during sepsis. Ann Surg. 1991;214:621–626.

121. Ochoa J, Strange J, Kearney P, Gellin G, Endean E, Fitzpatrick E. Effects of L-arginine on the proliferation of T lymphocyte subpopulations. JPEN J Parenter Enteral Nutr. 2001a;25:23–29 http://dx.doi.org/10.1177/014860710102500123.

122. Ochoa JB, Bernard AC, O’Brien WE, et al. Arginase I expression and activity in human mononuclear cells after injury. Ann Surg. 2001b;233:393–399.

123. Ohtsuka Y, Nakaya J. Effect of oral administration of L-arginine on senile dementia. Am J Med. 2000;108:439 <http://dx.doi.org/10.1016/S0002-9343(99)00396-4>.

124. Oka RK, Szuba A, Giacomini JC, Cooke JP. A pilot study of l-arginine supplementation on functional capacity in peripheral arterial disease. Vasc Med. 2005;10:265–274 http://dx.doi.org/10.1191/1358863x05vm637oa.

125. Owen OE, Kalhan SC, Hanson RW. The key role of anaplerosis and cataplerosis for citric acid cycle function. J Biol Chem. 2002;277:30409–30412 http://dx.doi.org/10.1074/jbc.R200006200.

126. Palmer JP, Benson JW, Walter RM, Ensinck JW. Arginine-stimulated acute phase of insulin and glucagon secretion in diabetic subjects. J Clin Investig. 1976;58:565–570 http://dx.doi.org/10.1172/JCI108502.

127. Palmer R, Ashton DS, Moncada S. Vascular endothelial cells synthesize nitric oxide from L-arginine. Nature. 1988;333:664–666 http://dx.doi.org/10.1038/333664a0.

128. Peterson JM, Barbul A, Breslin RJ, Wasserkrug HL, Efron G. Significance of T-lymphocytes in wound healing. Surgery. 1987;102:300–305.

129. Piatti P, Monti LD, Valsecchi G, et al. Long-term oral L-Arginine administration improves peripheral and hepatic insulin sensitivity in Type 2 diabetic patients. Diabetes Care. 2001;24:875–880 http://dx.doi.org/10.2337/diacare.24.5.875.

130. Pieper GM, Siebeneich W, Dondlinger LA. Short-term oral administration of l-arginine reverses defective endothelium-dependent relaxation and cGMP generation in diabetes. Eur J Pharmacol. 1996;317:317–320 <http://dx.doi.org/10.1016/S0014-2999(96)00831-X>.

131. Prince M, Bryce R, Albanese E, Wimo A, Ribeiro W, Ferri CP. The global prevalence of dementia: a systematic review and metaanalysis. Alzheimer’s & Dementia. 2013;9:63–75 e2. doi:10.1016/j.jalz.2012.11.007.

132. Quillen JE, Sellke FW, Armstrong ML, Harrison DG. Long-term cholesterol feeding alters the reactivity of primate coronary microvessels to platelet products. Arteriosclerosis Thrombosis Vasc Biol. 1991;11:639–644 http://dx.doi.org/10.1161/01.ATV.11.3.639.

133. Rakoff JS, Siler TM, Sinha YN, Yen SSC. Prolactin and growth hormone release in response to sequential stimulation by arginine and synthetic TRF. J Clin Endocrinol Metab. 1973;37:641–644 http://dx.doi.org/10.1210/jcem-37-5-641.

134. Rector TS, Bank AJ, Mullen KA, et al. Randomized, double-blind, placebo-controlled study of supplemental oral L-arginine in patients with heart failure. Circulation. 1996;93:2135–2141 http://dx.doi.org/10.1161/01.CIR.93.12.2135.

135. Regunathan S, Reis DJ. Characterization of arginine decarboxylase in rat brain and liver: distinction from ornithine decarboxylase. J Neurochem. 2000;74:2201–2208.

136. Rodriguez PC, Zea AH, Culotta KS, Zabaleta J, Ochoa JB, Ochoa AC. Regulation of T cell receptor CD3zeta chain expression by L-arginine. J Biol Chem. 2002;277:21123–21129 http://dx.doi.org/10.1074/jbc.M110675200.

137. Rogers QR, Visek WJ. Metabolic role of urea cycle intermediates: nutritional and clinical aspects Introduction. J Nutr. 1985;115:505–508.

138. Ros E. Nuts and CVD. Br J Nutr. 2015;113(Suppl 2):S111–S120 http://dx.doi.org/10.1017/S0007114514003924.

139. Rougé C, Robert, Des C, et al. Manipulation of citrulline availability in humans. AJP: Gastrointest Liver Physiol. 2007;293:G1061–G1067 http://dx.doi.org/10.1152/ajpgi.00289.2007.

140. Ruel M, Beanlands RS, Lortie M, et al. Concomitant treatment with oral L-arginine improves the efficacy of surgical angiogenesis in patients with severe diffuse coronary artery disease: the Endothelial Modulation in Angiogenic Therapy randomized controlled trial. J Thorac Cardiovasc Surg. 2008;135:762–770 – 770.e1. doi:10.1016/j.jtcvs.2007.09.073.

141. Santhanam L, Lim HK, Lim HK, et al. Inducible NO synthase dependent S-nitrosylation and activation of arginase1 contribute to age-related endothelial dysfunction. Circul Res. 2007;101:692–702 http://dx.doi.org/10.1161/CIRCRESAHA.107.157727.

142. Satriano J. Agmatine: at the crossroads of the arginine pathways. Ann N Y Acad Sci. 2003;1009:34–43 http://dx.doi.org/10.1196/annals.1304.004.

143. Schäffer MR, Efron PA, Thornton FJ, Klingel K, Gross SS, Barbul A. Nitric oxide, an autocrine regulator of wound fibroblast synthetic function. J Immunol. 1997;158:2375–2381.

144. Schäffer MR, Tantry U, Thornton FJ, Barbul A. Inhibition of nitric oxide synthesis in wounds: pharmacology and effect on accumulation of collagen in wounds in mice. Eur J Surg. 1999;165:262–267 http://dx.doi.org/10.1080/110241599750007153.

145. Schiller LR. Diarrhea and malabsorption in the elderly. Gastroenterol Clin N Am. 2009;38:481–502 http://dx.doi.org/10.1016/j.gtc.2009.06.008.

146. Schmidt HHHW, Nau H, Wittfoht W, et al. Arginine is a physiological precursor of endothelium-derived nitric oxide. Eur J Pharmacol. 1988;154:213–216 <http://dx.doi.org/10.1016/0014-2999(88)90101-X.

147. Schmidt H, Warner T, Ishii K, Sheng H, Murad F. Insulin secretion from pancreatic B cells caused by L-arginine-derived nitrogen oxides. Science. 1992;255:721–723 http://dx.doi.org/10.1126/science.1371193.

148. Schrempf Decker GE, Baron DP, Brattig NW, Bockhorn H, Berg PA. Biological and immunological characterization of a human liver immunoregulatory protein. Hepatology. 1983;3:939–946 http://dx.doi.org/10.1002/hep.1840030610.

149. Schulman SP, Becker LC, Kass DA, et al. L-Arginine therapy in acute myocardial infarction: the vascular interaction with age in myocardial infarction (VINTAGE MI) randomized clinical trial. JAMA. 2006;295:58–64 http://dx.doi.org/10.1001/jama.295.1.58.

150. Schwartz L. Amelioration of microvascular angina with arginine supplementation. Ann Intern Med. 2003;138:160 http://dx.doi.org/10.7326/0003-4819-138-2-200301210-00027.

151. Schwedhelm E, Maas R, Freese R, et al. Pharmacokinetic and pharmacodynamic properties of oral L‐citrulline and L‐arginine: impact on nitric oxide metabolism. Br J Clin Pharmacol. 2008;65:51–59 http://dx.doi.org/10.1111/j.1365-2125.2007.02990.x.

152. Seifter E, Rettura G, Barbul A, Levenson SM. Arginine: an essential amino acid for injured rats. Surgery. 1978;84:224–230.

153. Sen CK, Gordillo GM, Roy S, et al. Human skin wounds: a major and snowballing threat to public health and the economy. Wound Repair Regen. 2009;17:763–771 http://dx.doi.org/10.1111/j.1524-475X.2009.00543.x.

154. Shi HP, Efron DT, Most D, Tantry US, Barbul A. Supplemental dietary arginine enhances wound healing in normal but not inducible nitric oxide synthase knockout mice. Surgery. 2000;128:374–378 http://dx.doi.org/10.1067/msy.2000.107372.

155. Shi HP, Most D, Efron DT, Witte MB, Barbul A. Supplemental L‐arginine enhances wound healing in diabetic rats. Wound Repair Regen. 2003;11:198–203 http://dx.doi.org/10.1046/j.1524-475X.2003.11308.x.

156. Shin W, Berkowitz DE, Ryoo S. Increased arginase II activity contributes to endothelial dysfunction through endothelial nitric oxide synthase uncoupling in aged mice. Exp Mol Med. 2012;44:594–602 http://dx.doi.org/10.3858/emm.2012.44.10.068.

157. Sipilä I. Inhibition of arginine-glycine amidinotransferase by ornithine A possible mechanism for the muscular and chorioretinal atrophies in gyrate atrophy of the choroid and retina with hyperornithinemia. Biochim Biophys Acta. 1980;613:79–84.

158. Staff AC, Berge L, Haugen G, Lorentzen B, Mikkelsen B, Henriksen T. Dietary supplementation with l‐arginine or placebo in women with pre‐eclampsia. Acta Obstet Gynecol Scand. 2004;83:103–107 http://dx.doi.org/10.1111/j.1600-0412.2004.00321.x.

159. Stead LM, Au KP, Jacobs RL, Brosnan ME, Brosnan JT. Methylation demand and homocysteine metabolism: effects of dietary provision of creatine and guanidinoacetate. AJP: Endocrinol Metab. 2001;281:E1095–E1100.

160. Susswein AJ, Katzoff A, Miller N, Hurwitz I. Nitric oxide and memory. Neuroscientist. 2004;10:153–162 http://dx.doi.org/10.1177/1073858403261226.

161. Tangphao O, Grossmann M, Chalon S, Hoffman BB, Blaschke TF. Pharmacokinetics of intravenous and oral l-arginine in normal volunteers. Br J Clin Pharmacol. 1999;47:261–266 http://dx.doi.org/10.1046/j.1365-2125.1999.00883.x.

162. Tanimura J. Studies on arginine in human semen II The effects of medication with L-arginine-HCL on male infertility. Bull Osaka Med Sch. 1967;13:84–89.

163. Tasaki T, Kwon YT. The mammalian N-end rule pathway: new insights into its components and physiological roles. Trends Biochem Sci. 2007;32:520–528 http://dx.doi.org/10.1016/j.tibs.2007.08.010.

164. Thams P, Capito K. L-arginine stimulation of glucose-induced insulin secretion through membrane depolarization and independent of nitric oxide. Eur J Endocrinol. 1999;140:87–93 http://dx.doi.org/10.1530/eje.0.1400087.

165. Thornton FJ, Schäffer MR, Witte MB, et al. Enhanced collagen accumulation following direct transfection of the inducible nitric oxide synthase gene in cutaneous wounds. Biochem Biophys Res Commun. 1998;246:654–659 http://dx.doi.org/10.1006/bbrc.1998.8681.

166. Urschel KL, Shoveler AK, Uwiera RRE, Pencharz PB, Ball RO. Citrulline is an effective arginine precursor in enterally fed neonatal piglets. J Nutr. 2006;136:1806–1813.

167. Valitutti S, Müller S, Salio M, Lanzavecchia A. Degradation of T cell receptor (TCR)-CD3-zeta complexes after antigenic stimulation. J Exp Med. 1997;185:1859–1864.

168. Varshavsky A. The N-end rule: functions, mysteries, uses. Proc Natl Acad Sci. 1996;93:12142–12149.

169. Visek WJ. Arginine needs, physiological state and usual diets A reevaluation. J Nutr. 1986;116:36–46.

170. Wakabayashi Y, Yamada E, Yoshida T, Takahashi H. Arginine becomes an essential amino acid after massive resection of rat small intestine. J Biol Chem. 1994;269:32667–32671.

171. Walker HA, McGing E, Fisher I, et al. Endothelium-dependent vasodilation is independent of the plasma L-arginine/ADMA ratio in men with stable angina. J Am Coll Cardiol. 2001;38:499–505 <http://dx.doi.org/10.1016/S0735-1097(01)01380-8>.

172. Wang Y, Zhang L, Zhou G, et al. Dietary L-arginine supplementation improves the intestinal development through increasing mucosal Akt and mammalian target of rapamycin signals in intra-uterine growth retarded piglets. Br J Nutr. 2012;108:1371–1381 http://dx.doi.org/10.1017/S0007114511006763.

173. Wascher TC, Graier WF, Dittrich P, et al. Effects of low‐dose l‐arginine on insulin‐mediated vasodilatation and insulin sensitivity. Eur J Clin Investig. 1997;27:690–695 http://dx.doi.org/10.1046/j.1365-2362.1997.1730718.x.

174. Watanabe G, Tomiyama H, Doba N. Effects of oral administration of L-arginine on renal function in patients with heart failure. J Hypertens. 2000;18:229–234.

175. Weir GC, Samols E, Loo S, Patel YC, Gabbay KH. Somatostatin and pancreatic polypeptide secretion: effects of glucagon, insulin, and arginine. Diabetes. 1978;28:35–40 http://dx.doi.org/10.2337/diab.28.1.35.

176. Williams JZ, Abumrad N, Barbul A. Effect of a specialized amino acid mixture on human collagen deposition. Ann Surg. 2002;236:369–374 discussion 374–5. doi:10.1097/01.SLA.0000027527.01984.00.

177. Wilson AM, Harada R, Nair N, Balasubramanian N, Cooke JP. L-arginine supplementation in peripheral arterial disease: no benefit and possible harm. Circulation. 2007;116:188–195 http://dx.doi.org/10.1161/CIRCULATIONAHA.106.683656.

178. Witte M. Arginase acts as an alternative pathway of L-arginine metabolism in experimental colon anastomosis. J Gastrointest Surg. 2003;7:378–385 <http://dx.doi.org/10.1016/S1091-255X(02)00431-6>.

179. Wolf A, Zalpour C, Theilmeier G, et al. Dietary L-arginine supplementation normalizes platelet aggregation in hypercholesterolemic humans. J Am Coll Cardiol. 1997;29:479–485 <http://dx.doi.org/10.1016/S0735-1097(97)00523-8>.

180. Wu G, Morris Jr SM. Arginine metabolism in mammals. Metabolic and Therapeutic Aspects of Amino Acids in Clinical Nutrition 2003;153–161.

181. Wu G, Bazer FW, Cudd TA, et al. Pharmacokinetics and safety of arginine supplementation in animals. J Nutr. 2007;137:1673S–1680S.

182. Wyss M, Kaddurah-Daouk R. Creatine and creatinine metabolism. Physiol Rev. 2000;80:1107–1213.

183. Xiong Y, Fru MF, Yu Y, Montani J-P, Ming X-F, Yang Z. Long term exposure to L-arginine accelerates endothelial cell senescence through arginase-II and S6K1 signaling. Aging (Albany NY). 2014;6:369–379.

184. Xu, J., Murphy, S.L., Kochanek, K.D., Bastian, B.A., 2016. National Vital Statistics Reports [WWW Document]. cdc.gov. URL <http://www.cdc.gov/nchs/data/nvsr/nvsr64/nvsr64_02.pdf> (accessed 3.30.16).

185. Yamasaki K, Edington HD, McClosky C, et al. Reversal of impaired wound repair in iNOS-deficient mice by topical adenoviral-mediated iNOS gene transfer. J Clin Investig. 1998;101:967–971 http://dx.doi.org/10.1172/JCI2067.

186. Yi J, Horky LL, Friedlich AL, Shi Y, Rogers JT, Huang X. L-arginine and Alzheimer’s disease. Int J Clin Exp Pathol. 2009;2:211–238.

187. Yu H, Iyer RK, Kern RM, Rodriguez WI, Grody WW, Cederbaum SD. Expression of arginase isozymes in mouse brain. Journal of Neurosci Res. 2001;66:406–422 http://dx.doi.org/10.1002/jnr.1233.

188. Yu H, Yoo PK, Aguirre CC, et al. Widespread expression of arginase I in mouse tissues Biochemical and physiological implications. J Histochem Cytochem. 2003;51:1151–1160 http://dx.doi.org/10.1177/002215540305100905.

189. Zhou L, Zhu D-Y. Neuronal nitric oxide synthase: structure, subcellular localization, regulation, and clinical implications. Nitric Oxide. 2009;20:223–230 http://dx.doi.org/10.1016/j.niox.2009.03.001.

190. Zorgniotti AW, Lizza EF. Effect of large doses of the nitric oxide precursor, L-arginine, on erectile dysfunction. Int J Impot Res. 1994;6:33–35 – discussion 36.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.224.59.231