Skip to main content

REVIEW article

Front. Pediatr., 01 May 2024
Sec. Neonatology
Volume 12 - 2024 | https://doi.org/10.3389/fped.2024.1371710

Oxygen in the neonatal ICU: a complicated history and where are we now?

  • Division of Neonatology, Oregon Health & Science University, Portland, OR, United States

Despite major advances in neonatal care, oxygen remains the most commonly used medication in the neonatal intensive care unit (NICU). Supplemental oxygen can be life-saving for term and preterm neonates in the resuscitation period and beyond, however use of oxygen in the neonatal period must be judicious as there can be toxic effects. Newborns experience substantial hemodynamic changes at birth, rapid energy consumption, and decreased antioxidant capacity, which requires a delicate balance of sufficient oxygen while mitigating reactive oxygen species causing oxidative stress. In this review, we will discuss the physiology of neonates in relation to hypoxia and hyperoxic injury, the history of supplemental oxygen in the delivery room and beyond, supporting clinical research guiding trends for oxygen therapy in neonatal care, current practices, and future directions.

1 Introduction

The transition from fetus to newborn is a complex adaptation that results in a low oxygen tension fetal environment to rapidly increasing oxygen exposure postnatally. One study in the 1950s found the oxygen tension of amniotic fluid in preterm infants to be very low, mostly between 6 and 14 mm Hg (1). The low resistance placenta is the source of fetal gas exchange and nutrition; however at birth, with the removal of the placenta, major changes in the cardiovascular system occur including decrease in pulmonary vascular resistance and subsequent increase in oxygen tension (2, 3). Furthermore, neonates have high-energy needs as rapid growth and development demand high oxygen delivery and consumption (2, 4). The brain in particular is undergoing rapid post-natal expansion and requires efficiently functioning mitochondria with a steady supply of oxygen to meet energy demands. Mitochondria coupled with oxygen are also responsible for immune response, calcium buffering, and regulating reactive oxygen species (ROS) production, which can trigger programmed cell death (58). A delicate balance exists between adequate oxygen for growth and development yet avoiding excess oxygen that can lead to toxicity and cell death.

Hypoxia occurs when there is inadequate oxygen at the cellular level and often presents as cyanosis (blue discoloration of skin and mucous membranes) in the neonate. Insufficient oxygen disrupts cerebral oxidative metabolism and can lead to depletion of energy reserves in tissues. Furthermore, hypoxia triggers neurotoxic biochemical pathways including: changes in membrane potentials and ion distribution, production of nitric-oxide, accumulation of excitatory amino acids in extracellular space, inflammation, and necrosis (911). During the hypoxic state, cells shift to anaerobic metabolism leading to buildup of lactate as a byproduct causing metabolic acidosis that further harms cellular processes. Chronic hypoxia ultimately leads to apoptosis and cell death, particularly in the brain resulting in a decreased number of neurons (9, 12). Hypoxia of the fetus or newborn is one of the main causes of brain injury that can lead to long-term neurodevelopmental impairment, especially in preterm infants (13, 14).

Although preterm and term neonates require high oxygen demand they are particularly susceptible to oxidative stress and injury. Oxygen is a key component in energy metabolism in the mitochondrial respiratory chain by being reduced to water. This process generates ROS and reactive nitrogen species (RNS), some of which are free radicals and highly toxic. They damage the structure of nucleic acids, lipids, and proteins leading to oxidative injury (1519). Other ROS, such as hydrogen peroxide, is not a free radical but important for physiologic cell signaling that regulates pulmonary circulation and blood flow within the ductus arteriosus (19). During normal fetal-to-neonatal transition the increased exposure to oxygen causes a mild physiologic oxidative stress that is crucial for postnatal adaptation. However, oxidative injury results when there is severe oxidative stress due to an imbalance between generated free radicals and inadequate neutralization by antioxidant systems, especially with a burst of free radical generation. Pulmonary capillary endothelial and alveolar epithelial cells are targets for ROS injury causing hypoplasia, lung edema, hemorrhage, and deposits of collagen, elastin, and hyaline membrane (2024). The developing gastrointestinal tract is also affected by hyperoxia which causes thickened ileal mucosa, fewer Paneth, goblet cells and vili, decreased secretory components, increased inflammation and fibrosis, and cell apoptosis (2531). All of these changes are detrimental to the gastrointestinal tract's vital function in mucosal immunity. Hyperoxia exposure is also associated with development of retinopathy of prematurity (ROP), which is an eye disease due to abnormal blood vessel growth that can lead to blindness. Oxidative stress has been linked to dysregulated signaling pathways that cause ROP (32, 33).

Preterm neonates are at a further disadvantage for combating oxidative stress due to decrease antioxidant capacity from insufficient accretion of important minerals and vitamins during gestation (34). Additionally, preterm infants are often born at a developmental stage with inadequate expression of antioxidant defense in the lung (35). Plasma vitamin C concentrations are initially higher in pre-term neonates than term neonates but follow a rapid and sharp decline over the first few days of life (36, 37). Biochemical vitamin D deficiency is seen in both term and preterm neonates but is more profound in preterm neonates (3841). A systematic review and meta-analysis demonstrated preterm and term neonates have vitamin E concentrations below the recommended levels at birth (42). Preterm and term neonates are also known to have vitamin A deficiency with multiple clinical trials evaluating the benefit of supplementation by measuring outcomes such as mortality, oxygen use, and growth, but have shown mixed results (43). The largest randomized controlled trial evaluating vitamin A supplementation (NeoVitA) for preventing bronchopulmonary dysplasia (BPD) or death in extremely low birth weight neonates has yet to publish results (44). There are also clinical trials evaluating supplementation of antioxidant enzymes in preterm and term neonates. Superoxide dismutase (SOD) was a promising enzyme for supplementation however there is conflicting evidence regarding long-term follow up (4548). A Cochrane meta-analysis demonstrated that there is insufficient data to conclude SOD is efficacious in preventing chronic lung disease of prematurity (46).

Oxygen toxicity is not limited to molecular injury but may also cause reprogramming of normal lung development. Alveolar simplification and aberrant, disorganized lung vasculature has been demonstrated in preterm baboons (49, 50) and in term rodents exposed to hyperoxia at birth (51, 52). Early exposure to hyperoxia inhibits normal cell proliferation and angiogenesis, leading to altered lung development. The degree of altered lung development may explain the variation in severity of respiratory outcomes in surviving preterm infants.

Pulse oximetry is the most common noninvasive method of measuring saturation of oxygen in the blood and guides the clinical application of supplemental oxygen based on specific thresholds. When air is breathed into the lungs, oxygen is transported into capillaries that in turn send oxygen rich blood to the heart, which subsequently gets pumped to the rest of the body. It is important to remember that the pulse oximeter detects oxygen saturation of the blood and not the lung itself as the lung likely requires increased oxygen to maintain normal oxygen tension in the blood and delivery to other organs.

2 Use of oxygen in the delivery room

2.1 Historical background

The discovery of oxygen is accredited to Joseph Priestly in 1774, however a Polish alchemist named Michal Sedziwój acknowledged its full significance as early as 1604 (53). Oxygen was used in neonatal resuscitation as early as 1780 with the first publication of oxygen use in newborn resuscitation occurring in 1928 (5456). By the 20th century it became the standard of care for asphyxiated newborns due to an unfounded belief in perinatal brain damage resulting from birth asphyxia (54, 55). The potential toxicity of oxygen was first noted in the late 1940s by the first published description of retrolental fibroplasia (RLF), an eye disease affecting preterm neonates, now known as ROP (57). A husband and wife ophthalmologist team at Johns Hopkins University further confirmed the progressive nature of RLF in 1948 (58). It wasn't until the 1950s that oxygen was identified as the offending agent causing RLF, leading to blindness in children (59, 60). This discovery led to a paradigm shift in limiting oxygen in the delivery room that likely led to increased morbidity and mortality of preterm and sick neonates (61). With the development of oxygen saturation monitoring in the 1980s, there was renewed interest in targeted oxygen therapy. Multiple studies have described normal increases in oxygen saturation that occurs in healthy term neonates and thus current neonatal resuscitation guidelines supports increasing oxygen saturation targets in the first 10 min of life (6266).

2.2 Term neonates

Asphyxiated term neonates were treated with 100% oxygen in the delivery room as the standard of care by the 1960s (67). The International Liaison Committee on Resuscitation (ILCOR) formed the first international guidelines for newborn resuscitation in 1992 in which use of 100% oxygen was recommended (68). Since those initial guidelines it has become clear that many practices at that time were not evidence based (69). By the 1990s, studies in human neonates demonstrated adequate resuscitation of asphyxiated term neonates with use of 21% FiO2 and harm in neonates exposed to 100% FiO2 (70, 71). Ramji et al. demonstrated in a randomized controlled trial (RCT) of 84 asphyxiated neonates that room air is as effective as 100% oxygen. Furthermore, neonates in the room air group had higher Apgar scores and less time to first cry with no neurologic sequela at 28 days follow up (70). Saugstad et al. lead an international, multicenter RCT including 703 asphyxiated neonates that demonstrated they could be adequately resuscitated with room air and recovered quicker than neonates resuscitated with 100% FiO2, assessed by Apgar scores and time to first breath and cry (71). At further follow up, conducted by a pediatrician using a standardized examination at 18–24 months, there was no significant difference in neurodevelopmental outcomes between the two groups (72). Studies in the 2000s further illustrated the harmful effects of resuscitation with 100% FiO2. Vento et al. demonstrated in a RCT of 40 asphyxiated term neonates that neonates resuscitated with 100% FiO2 compared to room air exhibited biochemical findings associated with prolonged oxidative stress that persisted even after 4 weeks of life (73). Vento et al. also published similar findings of 830 neonates treated over 6 years at their institution that further supported their conclusions (74). The most complete systematic review and meta-analysis published in 2008 demonstrated that asphyxiated neonates resuscitated with room air compared to 100% FiO2 had lower mortality with a trend toward decrease risk of hypoxic ischemic encephalopathy (HIE), a type of brain damage caused by lack of oxygen to the brain during or shortly after birth in which diagnosis is made by a combination of physical exam, blood work, and details of labor and delivery (75). Furthermore, long-term follow up did not show significant difference in neurodevelopmental outcome (76). These studies demonstrated a reduction in mortality with resuscitation using room air without evidence of harm. Term infants who are under normal hypoxic state during the fetal-to-neonatal transition and are then exposed to high concentration of oxygen (reoxygenation) during resuscitation likely generate oxygen free radicals that may explain this poorer outcome (77). Additionally, animal and human studies have demonstrated a decrease in cerebral blood flow with hyperoxia, increasing the risk for ischemic injury (78, 79).

Major changes occurred in ILCOR guidelines for newborn resuscitation from 2005 to 2010 including recommendation to start resuscitation with room air rather than 100% FiO2 (80). More recently, a systematic review and meta-analysis in 2019 further supported room air being superior to 100% FiO2 during the initial resuscitation of the term neonates by reducing neonatal mortality (risk ratio [RR] = 0.73; 95% confidence interval [CI]: 0.57–0.94) (81). This growth in evidence solidified the current neonatal resuscitation guidelines that strongly recommend starting with 21% FiO2 and caution against 100% FiO2 in neonates ≥35 weeks' gestation (82, 83).

2.3 Preterm neonates

At birth, fluid in the alveoli is reabsorbed and replaced with air. A combination of mechanical, vasoactive, and neural factors along with increased oxygenation result in an increase in pulmonary blood flow. This allows the lungs to participate in gas exchange and supply oxygenated blood to the heart. Preterm infants are at a high risk of inadequate pulmonary blood flow due to immature architecture of the lung and surfactant deficiency thus they likely require a higher oxygen concentration during resuscitation than term infants. However, evidence for the optimal level of oxygen used in the resuscitation of preterm neonates is much less conclusive than in term neonates. By the 1990s there was evidence that use of high levels of oxygen may be unnecessary and harmful to preterm neonates (79). However multiple studies demonstrate 21% FiO2 is not sufficient for the preterm population (8486). Therefore the optimal oxygen concentration in the initial resuscitation of preterm neonates is somewhere between 21% and 100% which is a broad range and tremendous knowledge gap in the field. The 2010 ILCOR consensus guidelines recommended against use of 100% FiO2 but initiating some supplemental oxygen in the initial resuscitation of preterm neonates and titrating based on oxygen saturation (87). A systematic review and meta-analysis in 2018 demonstrated no difference in mortality with use of lower (<40%) vs. higher (≥40%) FiO2 in the resuscitation of preterm neonates. Only 1 of the 10 included studies reported increased mortality in the lower (21%) vs. higher (100%) oxygen group (88). A RCT of 52 preterm neonates (<30 weeks' gestational age) demonstrated the higher (100%) oxygen group led to improved breathing effort, improved oxygenation, and shorter duration of mask ventilation as compared to the lower (30%) oxygen group without increased oxidative stress markers and no difference in mortality (89). There are plans for a prospective meta-analysis of current ongoing RCTs of neonates <29 weeks' gestation randomized to high (60%) vs. low (30%) oxygen for initial resuscitation at birth. Results projected to be complete by 2025 and neurodevelopmental outcomes by 2027 (90). The largest clinical trial to date of oxygen concentration for the resuscitation of preterm neonates in the delivery room demonstrated higher incidence of bradycardia in the room air group vs. 100% FiO2 (91). Studies further demonstrated oxygen saturation <80% at 5 min after birth was associated with increased mortality (9294). The current consensus for the initial resuscitation of preterm neonates is to target an oxygen saturation of >80% by the first 5 min of life and avoid bradycardia with goal heart rate >100 beats per min.

3 Use of oxygen in the neonatal ICU beyond the initial resuscitation

3.1 Historical background

Hyperoxia (state of excess supply of oxygen in tissues and organs) beyond the initial resuscitation is associated with multisystem morbidity and mortality including lung injury (BPD), ROP, brain injury (intraventricular hemorrhage and periventricular malacia), and intestinal injury. Yet, the optimal concentration of oxygen is still unknown, especially in preterm neonates. When hyperoxia was identified as the cause of RLF and blindness in children in the 1950s, the use of supplemental oxygen was intensely curtailed by the 1960s despite estimate of 16 additional deaths for every case of blindness prevented (59, 95, 96). In 1977, a large prospective observational study could not identify a relation between high partial pressure of arterial oxygen (PaO2 ≥ 100 and 150) and RLF however demonstrated a strong association of RLF with cumulative supplemental oxygen exposure (97). With the advent of transcutaneous PO2 electrodes, oxygen monitoring became more accessible and precise allowing for tighter control by the 1980s and 1990s (98100). In 2007, the American Academy of Pediatrics recommended an oxygen saturation range of 85%–95% in neonates in the first two weeks of life (101).

3.2 Oxygen saturation targets for preterm neonates

There were only a few, small, randomized trials evaluating oxygen saturation targets in preterm neonates until the early 2000s. The Supplemental Therapeutic Oxygen for Prethreshold Retinopathy of Prematurity (STOP-ROP) trial randomized 649 preterm neonates with prethreshold ROP at 35 weeks' postmenstrual age to either 89%–94% or 96%–99% oxygen saturation (SpO2) by pulse oximetry. This study demonstrated no difference in incidence of ROP, however the higher threshold group had more adverse pulmonary events (pneumonia and/or exacerbations of chronic lung disease and need for oxygen, diuretics, and hospitalization at 3 months corrected age) (102). The Benefits of Oxygen Saturation Targeting (BOOST) trial randomized neonates born <30 weeks' gestation who needed supplemental oxygen at 32 weeks' postmenstrual age to either 91%–94% or 95%–98% SpO2. This study demonstrated no difference in ROP, growth, or development; however the higher threshold group required more home oxygen (103). However, both these studies tested oxygen saturation targets that would be considered very high by today's standard.

Following these studies, the Neonatal Oxygenation Prospective Meta-Analysis (NeOProM) collaborative was formed in 2003 to investigate optimal oxygen targets in extremely preterm neonates (born at <28 weeks' of gestational age). This was a prospective multicenter study that included five clinical studies with similar protocols to provide individual participant data at trial completion for inclusion in an independent patient meta-analysis. This study randomized 4,965 preterm neonates <28 weeks' gestational age to either 85%–89% or 91%–95% SpO2 within the first 24 h of birth (104). The studies found no difference in combined outcome of death and/or major neurodevelopmental impairment at 18–24 months however the lower oxygen saturation target group had a significantly increased mortality (105109). The lower target threshold group had reduced risk of ROP requiring treatment however there was significant heterogeneity with only one trial within NeOProM that found significant reduction in ROP treatment. Furthermore, this did not translate to increased severe visual impairment at 18–24 months (105110). The lower oxygen target group also had higher incidence of necrotizing enterocolitis (NEC), a disease of the intestinal tract in which tissue lining becomes inflamed, dies, and sloughs off. However, the study demonstrated no difference between the groups for BPD, intraventricular hemorrhage (IVH), periventricular leukomalacia (PVL), and neurodevelopmental outcomes (108, 110). However, a major limitation to consider is the significant overlap in the actual exposed oxygen saturation between the groups resulting in poor separation of the intervention and comparison groups (110).

4 Methods of oxygen measurement and titration

Improved measurement of blood oxygen levels and more precise titration of supplemental oxygen may decrease the time spent in hypoxia and hyperoxia, thus reducing oxygen toxicity. In this section we will discuss current methods of oxygen measurement and titration and ongoing investigations in this area.

4.1 Pulse oximetry

4.1.1 Principles

Measurement of arterial oxygen saturation by pulse oximetry is based on the attenuation of light to the materials through which the light travels and photoplethysmography, an optical technique used to detect blood volume changes in the microvascular bed (111). Red and infrared lights are used based on the differential absorption of oxyhemoglobin and deoxyhemoglobin. During systole, the transmitted light intensity decreases as light is absorbed by the hemoglobin in the arteries. During diastole, transmitted light intensity increases due to decrease absorption. The ratios of the absorbance of the two wavelengths of light during systole and diastole are compared by the pulse oximeter and a built in algorithm converts the ratios to pulse oximetry oxygen saturation (112). This is measured on the neonate by transmitting the light through a distal extremity with a detector placed on the opposite side of the same extremity. The major advantage of pulse oximetery is the non-invasive ease of application and ability for continuous monitoring, however many limitations exist.

4.1.2 Limitations

Due to the shape of the oxygen-hemoglobin dissociation curve, pulse oximetry has difficulty detecting hyperoxia >90% SpO2. Therefore small increase in SpO2 may indicate large increases in partial pressure of oxygen in arterial blood (PaO2) (113, 114). This is a physiologic limitation that has major implications for oxygen toxicity and stress in preterm neonates. Additionally, data used to calibrate pulse oximetry come from healthy Caucasian adults with adult hemoglobin ranges that are correlated to arterial blood samples. This poses a major limitation for calibration for different skin types, ethnicities, and ages. When calibrating for SpO2 < 80%, manufacturers rely on extrapolation and thus are not as accurate. Most manufacturers claim an accuracy of ±2% (115). Signal averaging time also places a limitation on pulse oximetry by causing response delays. Most conventional pulse oximeters have a default time of 8 s to display average values of SpO2 measured over this time period. Signal averaging time can be adjusted to a lower time however this may increase the frequency of alarms causing “alarm fatigue” to bedside providers. Averaging times can also be increased, however this may prevent detection of important hypoxemic episodes (116, 117). Motion artifacts are a common disturbance with pulse oximetry. This can lead to false alarms by movements of the probes or poor peripheral perfusion. However modern algorithms have decreased this issue. Despite these limitations, pulse oximetry remains the most widely used monitoring system in neonates. However, it is critical to recognize that pulse oximetry is a surrogate marker for tissue oxygenation and only measures peripheral arterial oxygen saturation.

4.1.3 Automatic oxygen control

A major clinical variable affecting oxygen saturation targeting using pulse oximetry is the labor-intensive process of manually titrating FiO2 by bedside staff. Studies have shown despite multiple adjustments per hour, target ranges were achieved only 50% of the time (118). Based on this concern and need for tighter control of oxygen in neonates, automatic oxygen control (AOC) systems held high promise. AOC systems consist of an oxygenation monitoring device, (pulse oximeter), gas delivery system (ventilator or cannula), and an algorithm to determine timing and degree of FiO2 adjustments (119). Multiple studies, spanning the last 40 years, in term and very low birth weight preterm neonates have demonstrated AOC systems were more effective than manual titration in achieving oxygenation targets (120133). There have been some studies that investigated long term effects and found no difference in neurodevelopmental outcome at 2 years of age between the automated oxygen control group and standard of care manual control groups (134).

Despite these findings there are many reasons AOC systems are not in widespread use. As the optimal target saturation for preterm neonates is still unknown a more consistent maintenance of target oxygen saturation may uncover adverse or beneficial effects. Furthermore, the reliability of pulse oximetry is key to a dependable AOC system. Close monitoring of this automated system is necessary to ensure accurate reading by pulse oximetry. Additionally, this automated system is not a replacement for provider response in investigating mechanism of oxygen saturation changes and appropriate intervention needed. Despite the known physiological benefits, there is insufficient data regarding improved clinical benefits to confidently introduce these systems into routine clinical care. There are currently ongoing trials to investigate effect of AOC on clinical outcomes including prolonged ventilation, death, severe ROP, BPD, NEC, and neurodevelopmental impairment (135, 136).

4.2 Near-infrared spectroscopy

Another technology that allows continuous non-invasive monitoring of oxygenation is near-infrared spectroscopy (NIRS). Unlike pulse oximetry, NIRS does not detect pulse waveform. It measures transparency of biological tissue of near-infrared spectrum of light affected by oxygen consumption and oxygen delivery by measuring ratio of oxygenation and deoxygenated hemoglobin (137). Originally developed to assess cerebral oxygenation perioperatively during cardiac surgery and neurosurgery, it has expanded to other clinical settings and ages including neonates. The most frequently measured sites in neonates include cerebral, renal, and splanchnic (intestinal) oxygenation.

4.2.1 Cerebral NIRS

The brain is especially vulnerable to hypoxia during the fetal to neonatal transition and thus cerebral NIRS was considered a potential tool in resuscitation of newborns in the late 1970s (138). However given lack of established standards it has been difficult to include cerebral NIRS in the routine resuscitation of newborns. A systematic review in 2022 investigated whether early NIRS monitoring (<6 h of age) can predict neurodevelopmental outcome at 1–2 years in infants with HIE and demonstrated no significant difference in values of cerebral oxygenation. They further concluded very little data exists and further studies are required with standardized approach for adequate comparison (139). A recent multinational, randomized controlled trial of over 600 preterm neonates demonstrated no difference in survival or cerebral injury when cerebral NIRS was measured in combination with defined treatment guidelines in the immediate resuscitation after birth (140).

4.2.2 Renal NIRS

Acute kidney injury (AKI) can result from low renal blood flow especially in the setting of a patent ductus arteriosus (PDA) in very low birth weight neonates. Renal NIRS has been used to help identify early markers of AKI. A prospective observational study in France demonstrated low renal NIRS values during the first 24 h of life is associated with development of AKI in preterm neonates (141). However conflicting evidence exists as well. High renal NIRS in neonates with HIE was demonstrated to be associated with AKI, whereas lower renal NIRS in neonates undergoing cardiac surgery was associated with AKI (142, 143).

4.2.3 Splanchnic (intestinal) NIRS

There has been robust interest in the use of splanchnic or intestinal NIRS in newborns to detect early markers for development of necrotizing enterocolitis (NEC). A prospective cohort study of 100 preterm neonates demonstrated decreased intestinal NIRS values and increased variability in neonates who developed NEC (144). Another prospective observational study of 10 preterm neonates and 20 matched controls demonstrated intestinal oxygenation is impaired before the onset of clinical NEC (145).

4.2.4 Limitations

NIRS is an indirect measure of oxygenation and therefore has several limitations of its application when there is movement, low arterial oxygen levels, darker skin pigmentation, or changes in weight and/or edema (146). Perhaps the biggest limitation is that there are no established standards and specific guidelines for use and interpretation of NIRS in neonates are lacking. Currently, there is not enough evidence to support the routine use of NIRS in neonates to justify additional monitoring on fragile neonates with limited body surface area.

5 Future directions

Currently, there are no available therapies to combat oxidative injury in neonates. Further research in antioxidant therapy including target specificity, bioavailability, and genetic variability may reveal novel approaches to mitigate oxygen toxicity. Stem cells hold therapeutic potential for hypoxia induced cellular injury and apoptosis by promoting neuronal cellular repair and regeneration. However more clinical studies are needed to determine the stem cell type, patient selection, route and time of administration to achieve standardized products and a refined protocol (147). Additionally the optimal level of positive end-expiratory pressure (PEEP) to minimize hyperoxic lung injury but provide adequate lung recruitment during invasive ventilation is unknown. Future research in ventilatory strategies that can be individualized through oxygenation guided lung recruitment methods may provide a personalized approach to mitigate effects of hypoxia and hyperoxia. Furthermore, noninvasive ventilation in preterm infants is also not standardized. Continuous positive airway pressure (CPAP) is the standard of care for preterm infants with respiratory distress syndrome, however the level and duration of CPAP is highly variable. Future research in this area may minimize lung injury by investigating the optimal CPAP strategies to reduce hypoxia and hyperoxia in preterm infants.

6 Conclusions

Term and preterm neonates face the unique dilemma of potentially needing life-saving oxygen supplementation while also at risk for oxygen toxicity. The rapidly changing physiology at birth, the extensive energy needs for growth and development, and decreased antioxidant capacity put them at this unique risk. Neonatology has had a complicated history with the application of oxygen from using it ubiquitously to fierce restriction due to retinopathy of prematurity, which left our sickest and most preterm neonates at risk of morbidity and mortality. With the advent of pulse oximetry there has been more refined approach to oxygen supplementation and titration however the exact thresholds still remain unknown. Furthermore, it is likely that optimal oxygenation saturation is not binary but varies throughout NICU hospitalization. There are ongoing investigations into automatic oxygen control and near-infrared spectroscopy to shed light on the optimal strategies for oxygen thresholds while avoiding oxygen toxicity and stress. Further research in antioxidant systems, stem cell therapy, optimal PEEP, and optimal CPAP level and duration may also reveal other strategies to improve oxygen therapy in preterm infants.

Author contributions

RM: Conceptualization, Writing – original draft, Writing – review & editing. CM: Conceptualization, Supervision, Writing – review & editing.

Funding

The author(s) declare financial support was received for the research, authorship, and/or publication of this article.

Funding/Support: T32HL083808-11A1 (RM) and R33HL147906 (CM). Role of Funder/Sponsor (if any): The funders had no role in study design, data collection, analysis, decision to publish, or preparation of the manuscript.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher's note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

1. Sjostedt S, Rooth G, Caligara F. The oxygen tension of the amniotic fluid. Am J Obstet Gynecol. (1958) 76(6):1226–30. doi: 10.1016/S0002-9378(16)36937-X

PubMed Abstract | Crossref Full Text | Google Scholar

2. Hillman NH, Kallapur SG, Jobe AH. Physiology of transition from intrauterine to extrauterine life. Clin Perinatol. (2012) 39(4):769–83. doi: 10.1016/j.clp.2012.09.009

PubMed Abstract | Crossref Full Text | Google Scholar

3. Johnson N, Johnson VA, McNamara H, Montague IA, Jongsma HW, Aumeerally Z, et al. Fetal pulse oximetry: a new method of monitoring the Fetus. Aust N Z J Obstet Gynaecol. (1994) 34:428–32. doi: 10.1111/j.1479-828X.1994.tb01263.x

PubMed Abstract | Crossref Full Text | Google Scholar

4. Butte NF, Moon JK, Wong WW, Hopkinson JM, Smith EO. Energy requirements from infancy to adulthood. Am J Clin Nutr. (1995) 62:1047S–52S. doi: 10.1093/ajcn/62.5.1047S

PubMed Abstract | Crossref Full Text | Google Scholar

5. Jones A, Thornton C. Mitochondrial dynamics in the neonatal brain—a potential target following injury? Biosci Rep. (2022) 42(3):BSR20211696. doi: 10.1042/BSR20211696

PubMed Abstract | Crossref Full Text | Google Scholar

6. Andrieux P, Chevillard C, Cunha-Neto E, Nunes JPS. Mitochondria as a Cellular Hub in Infection and Inflammation. Int J Mol Sci. (2021) 22(21):11338. doi: 10.3390/ijms222111338

PubMed Abstract | Crossref Full Text | Google Scholar

7. Godoy JA, Rios JA, Picon-Pages P, Herrera-Fernandez V, Swaby B, Crepin G, et al. Mitostasis, calcium and free radicals in health, aging and neurodegeneration. Biomolecules. (2021) 11:1012. doi: 10.3390/biom11071012

PubMed Abstract | Crossref Full Text | Google Scholar

8. Dadsena S, King LE, Garcia-Saez AJ. Apoptosis regulation at the mitochondria membrane level. Biochim Biophys Acta Biomembr. (2021) 1863:183716. doi: 10.1016/j.bbamem.2021.183716

PubMed Abstract | Crossref Full Text | Google Scholar

9. Piešová M, Mach M. Impact of perinatal hypoxia on the developing brain. Physiol Res. (2020) 69(2):199–213. doi: 10.33549/physiolres.934198

Crossref Full Text | Google Scholar

10. Esih K, Goričar K, Dolžan V, Rener-Primec Z. Antioxidant polymorphisms do not influence the risk of epilepsy or its drug resistance after neonatal hypoxic-ischemic brain injury. Seizure. (2017) 46:38–42. doi: 10.1016/j.seizure.2017.01.005

PubMed Abstract | Crossref Full Text | Google Scholar

11. Sab IM, Ferraz MMD, Amaral TAS, Resende AC, Ferraz MR, Matusra C, et al. Prenatal hypoxia, habituation memory and oxidative stress. Pharmacol Biochem Behav. (2013) 107:24–8. doi: 10.1016/j.pbb.2013.04.004

PubMed Abstract | Crossref Full Text | Google Scholar

12. Guo R, Hou W, Dong Y, Yu Z, Stites J, Weiner CP. Brain injury caused by chronic fetal hypoxemia is mediated by inflammatory cascade activation. Reprod Sci. (2010) 17:540–8. doi: 10.1177/1933719110364061

PubMed Abstract | Crossref Full Text | Google Scholar

13. Alderliesten T, van Bel F, van der Aa NE, Steendijk P, van Haastert IC, de Vries LS, et al. Low cerebral oxygenation in preterm infants is associated with adverse neurodevelopmental outcome. J Pediatr. (2019) 207:109–16.e2. doi: 10.1016/j.jpeds.2018.11.038

PubMed Abstract | Crossref Full Text | Google Scholar

14. Verhagen EA, Van Braeckel KN, van der Veere CN, Groen H, Dijk PH, Hulzebos CV, et al. Cerebral oxygenation is associated with neurodevelopmental outcome of preterm children at age 2 to 3 years. Dev Med Child Neurol. (2015) 57(5):449–55. doi: 10.1111/dmcn.12622

PubMed Abstract | Crossref Full Text | Google Scholar

15. Millán I, Piñero-Ramos JD, Lara I, Parra-Llorca A, Torres-Cuevas I, Vento M. Oxidative stress in the newborn period: useful biomarkers in the clinical setting. Antioxidants (Basel). (2018) 7(12):193. doi: 10.3390/antiox7120193

Crossref Full Text | Google Scholar

16. Lobo V, Patil A, Phatak A, Chandra N. Free radicals, antioxidants and functional foods: impact on human health. Pharmacogn Rev. (2010) 4:118–26. doi: 10.4103/0973-7847.70902

PubMed Abstract | Crossref Full Text | Google Scholar

17. Beharry KD, Cai CL, Valencia GB, Valencia AM, Lazzaro DR, Bany-Mohammed F, et al. Neonatal intermittent hypoxia, reactive oxygen Species, and oxygen-induced retinopathy. React Oxyg Species. (2017) 3:12–25. doi: 10.20455/ros.2017.805

Crossref Full Text | Google Scholar

18. Kalogeris T, Bao Y, Korthuis RJ. Mitochondrial reactive oxygen species: a double edged sword in ischemia/reperfusion vs. preconditioning. Redox Biol. (2014) 2:702–14. doi: 10.1016/j.redox.2014.05.006

PubMed Abstract | Crossref Full Text | Google Scholar

19. Saugstad OD, Oei JL, Lakshminrusimha S, Vento M. Oxygen therapy of the newborn from molecular understanding to clinical practice. Pediatr Res. (2019) 85:20–9. doi: 10.1038/s41390-018-0176-8

PubMed Abstract | Crossref Full Text | Google Scholar

20. Mach WJ, Thimmesch AR, Pierce JT, Pierce JD. Consequences of hyperoxia and the toxicity of oxygen in the lung. Nurs Res Pract. (2011) 2011:260482. doi: 10.1155/2011/260482

PubMed Abstract | Crossref Full Text | Google Scholar

21. Yee M, Vitiello PF, Roper JM, Staversky RJ, Wright TW, McGrath-Morrow SA, et al. Type II epithelial cells are critical target for hyperoxia-mediated impairment of postnatal lung development. Am J Physiol Lung Cell Mol Physiol. (2006) 291(5):L1101–11. doi: 10.1152/ajplung.00126.2006

PubMed Abstract | Crossref Full Text | Google Scholar

22. Kang KW, Lee SJ, Kim SG. Molecular mechanism of Nrf2 activation by oxidative stress. Antioxid Redox Signal. (2005) 7(11–12):1664–73. doi: 10.1089/ars.2005.7.1664

PubMed Abstract | Crossref Full Text | Google Scholar

23. Pagano A, Barazzone-Argiroffo C. Alveolar cell death in hyperoxia-induced lung injury. Ann N Y Acad Sci. (2003) 1010:405–16. doi: 10.1196/annals.1299.074

PubMed Abstract | Crossref Full Text | Google Scholar

24. Ratner V, Starkov A, Matsiukevich D, Polin RA, Ten VS. Mitochondrial dysfunction contributes to alveolar developmental arrest in hyperoxia-exposed mice. Am J Respir Cell Mol Biol. (2009) 40:511–8. doi: 10.1165/rcmb.2008-0341RC

PubMed Abstract | Crossref Full Text | Google Scholar

25. Giannone PJ, Bauer JA, Schanbacher BL, Reber KM. Effects of hyperoxia on postnatal intestinal development. Biotech Histochem. (2007) 82:17–22. doi: 10.1080/10520290701257112

PubMed Abstract | Crossref Full Text | Google Scholar

26. Sherman MP, Bennett SH, Hwang FF, Sherman J, Bevins CL. Paneth cells and antibacterial host defense in neonatal small intestine. Infect Immun. (2005) 73:6143–6. doi: 10.1128/IAI.73.9.6143-6146.2005

PubMed Abstract | Crossref Full Text | Google Scholar

27. Lueschow SR, McElroy SJ. The Paneth cell: the curator and defender of the immature small intestine. Front Immunol. (2020) 11:587. doi: 10.3389/fimmu.2020.00587

PubMed Abstract | Crossref Full Text | Google Scholar

28. Bruno ME, Frantz AL, Rogier EW, Johansen FE, Kaetzel CS. Regulation of the polymeric immunoglobulin receptor by the classical and alternative NF-κB pathways in intestinal epithelial cells. Mucosal Immunol. (2011) 4:468–78. doi: 10.1038/mi.2011.8

PubMed Abstract | Crossref Full Text | Google Scholar

29. Torbati D, Tan GH, Smith S, Frazier KS, Gelvez J, Fakioglu H, et al. Multiple-organ effect of normobaric hyperoxia in neonatal rats. J Crit Care. (2006) 21:85–93. doi: 10.1016/j.jcrc.2005.09.057

PubMed Abstract | Crossref Full Text | Google Scholar

30. Zhao M, Tang S, Xin J, Liu D. Influence of reactive oxygen species on secretory component in the intestinal epithelium during hyperoxia. Exp Med. (2017) 14:4033–40. doi: 10.3892/etm.2017.5027

Crossref Full Text | Google Scholar

31. Li TM, Liu DY. Mechanism of neonatal intestinal injury induced by hyperoxia therapy. J Immunol Res. (2022) 2022:2316368. doi: 10.1155/2022/2316368

PubMed Abstract | Crossref Full Text | Google Scholar

32. Hartnett ME. Pathophysiology and mechanisms of severe retinopathy of prematurity. Ophthalmology. (2015) 122(1):200–10. doi: 10.1016/j.ophtha.2014.07.050

PubMed Abstract | Crossref Full Text | Google Scholar

33. Hartnett ME, Penn JS. Mechanisms and management of retinopathy of prematurity. N Engl J Med. (2012) 367:2515–26. doi: 10.1056/NEJMra1208129

PubMed Abstract | Crossref Full Text | Google Scholar

34. Lee YS, Chou YH. Antioxidant profiles in full term and preterm neonates. Chang Gung Med J. (2005) 28:846–51.16515018

PubMed Abstract | Google Scholar

35. Cannavò L, Perrone S, Viola V, Marseglia L, Di Rosa G, Gitto E. Oxidative stress and respiratory diseases in preterm newborns. Int J Mol Sci. (2021) 22(22):12504. doi: 10.3390/ijms222212504

Crossref Full Text | Google Scholar

36. Darlow BA, Buss H, McGill F, Fletcher L, Graham P, Winterbourn CC. Vitamin C supplementation in very preterm infants: a randomized controlled trial. Arch Dis Child Fetal Neonatal Ed. (2005) 90:F117–22. doi: 10.1136/adc.2004.056440

PubMed Abstract | Crossref Full Text | Google Scholar

37. Berger TM, Rifai N, Avery ME, Frei B. Vitamin C in premature and full-term human neonates. Redox Rep. (1996) 2(4):257–62. doi: 10.1080/13510002.1996.11747058

PubMed Abstract | Crossref Full Text | Google Scholar

38. Adnan M, Wu SY, Khilfeh M, Davis V. Vitamin D status in very low birth weight infants and response to vitamin D intake during their NICU stays: a prospective cohort study. J Perinatol. (2022) 42(2):209–16. doi: 10.1038/s41372-021-01238-9

PubMed Abstract | Crossref Full Text | Google Scholar

39. Burris HH, Van Marter LJ, McElrath TF, Tabatabai P, Litonjua AA, Weiss ST, et al. Vitamin D status among preterm and full-term infants at birth. Pediatr Res. (2014) 75:75–80. doi: 10.1038/pr.2013.174

PubMed Abstract | Crossref Full Text | Google Scholar

40. Monangi N, Slaughter JL, Dawodu A, Smith C, Akinbi HT. Vitamin D status of early preterm infants and the effects of vitamin D intake during hospital stay. Arch Dis Child Fetal Neonatal Ed. (2014) 99:F166–8. doi: 10.1136/archdischild-2013-303999

PubMed Abstract | Crossref Full Text | Google Scholar

41. Fort P, Salas AA, Nicola T, Craig CM, Carlo WA, Ambalavanan N. A comparison of three vitamin D dosing regimens in extremely preterm infants: a randomized controlled trial. J Pediatr. (2016) 174:132–8.e1. doi: 10.1016/j.jpeds.2016.03.028

PubMed Abstract | Crossref Full Text | Google Scholar

42. Assunção DGF, Silva LTPD, Camargo JDAS, Cobucci RN, Ribeiro KDDS. Vitamin E levels in preterm and full-term infants: a systematic review. Nutrients. (2022) 14(11):2257. doi: 10.3390/nu14112257

Crossref Full Text | Google Scholar

43. Kumar A, Anjankar A. A narrative review of vitamin A supplementation in preterm and term infants. Cureus. (2022) 14(10):e30242. doi: 10.7759/cureus.30242

PubMed Abstract | Crossref Full Text | Google Scholar

44. Meyer S, Gortner L, NeoVitaA Trial Investigators. Early postnatal additional high-dose oral vitamin A supplementation versus placebo for 28 days for preventing bronchopulmonary dysplasia or death in extremely low birth weight infants. Neonatology. (2014) 105(3):182–8. doi: 10.1159/000357212

PubMed Abstract | Crossref Full Text | Google Scholar

45. Albertella M, Gentyala RR, Paraskevas T, Ehret D, Bruschettini M, Soll R. Superoxide dismutase for bronchopulmonary dysplasia in preterm infants. Cochrane Database Syst Rev. (2023) 10:CD013232. doi: 10.1002/14651858.CD013232.pub2

PubMed Abstract | Crossref Full Text | Google Scholar

46. Suresh GK, Davis JM, Soll RF. Superoxide dismutase for preventing chronic lung disease in mechanically ventilated preterm infants. Cochrane Database Syst Rev. (2001) 2001(1):CD001968. doi: 10.1002/14651858.CD001968

PubMed Abstract | Crossref Full Text | Google Scholar

47. Davis JM, Parad RB, Michele T, Allred E, Price A, Rosenfeld W, North American Recombinant Human CuZnSOD Study Group. Pulmonary outcome at 1 year corrected age in preterm infants treated at birth with recombinant human CuZn superoxide dismutase. Pediatrics. (2003) 111(3):469–76. doi: 10.1542/peds.111.3.469

PubMed Abstract | Crossref Full Text | Google Scholar

48. Parad RB, Allred EN, Rosenfeld WN, Davis JM. Reduction of retinopathy of prematurity in extremely low gestational age newborns treated with recombinant human Cu/Zn superoxide dismutase. Neonatology. (2012) 102(2):139–44. doi: 10.1159/000336639

PubMed Abstract | Crossref Full Text | Google Scholar

49. Coalson JJ, Winter VT, Gerstmann DR, Idell S, King RJ, Delemos RA. Pathophysiologic, morphometric, and biochemical studies of the premature baboon with bronchopulmonary dysplasia. Am Rev Respir Dis. (1992) 145:872–81. doi: 10.1164/ajrccm/145.4_Pt_1.872

PubMed Abstract | Crossref Full Text | Google Scholar

50. Maniscalco WM, Watkins RH, O’Reilly MA, Shea CP. Increased epithelial cell proliferation in very premature baboons with chronic lung disease. Am J Physiol Lung Cell Mol Physiol. (2002) 283:L991–1001. doi: 10.1152/ajplung.00050.2002

PubMed Abstract | Crossref Full Text | Google Scholar

51. Warner BB, Stuart LA, Papes RA, Wispé JR. Functional and pathological effects of prolonged hyperoxia in neonatal mice. Am J Physiol. (1998) 275:L110–17. doi: 10.1152/ajplung.1998.275.1.L110

PubMed Abstract | Crossref Full Text | Google Scholar

52. Bonikos DS, Bensch KG, Ludwin SK, Northway WH Jr, Oxygen toxicity in the newborn. The effect of prolonged 100 per cent O2 exposure on the lungs of newborn mice. Lab Invest. (1975) 32:619–35.1092910

PubMed Abstract | Google Scholar

53. Richmond S, Goldsmith JP. Refining the role of oxygen administration during delivery room resuscitation: what are the future goals. Semin Fetal Neonatal Med. (2008) 13:368–74. doi: 10.1016/j.siny.2008.04.002

PubMed Abstract | Crossref Full Text | Google Scholar

54. Obladen M. History of neonatal resuscitation—part 2: oxygen and other drugs. Neonatol. (2008) 95(1):91–6. doi: 10.1159/000151761

Crossref Full Text | Google Scholar

55. Kapadia V, Rabi Y, Oei JL. The goldilocks principle. Oxygen in the delivery room: when is it too little, too much, and just right? Semin Fetal Neonatal Med. (2018) 23(5):347–54. doi: 10.1016/j.siny.2018.06.003

PubMed Abstract | Crossref Full Text | Google Scholar

56. Flagg PJ. Treatment of asphyxia in the new-born: preliminary report of the practical application of modern scientific methods. J Am Med Assoc. (1928) 91:788–91. doi: 10.1001/jama.1928.02700110020006

Crossref Full Text | Google Scholar

57. Terry TL. Extreme prematurity and fibroblastic overgrowth of persistent vascular sheath behind each crystalline lens: I. Preliminary report. Am J Ophthalmol. (2018) 192:xxviii. doi: 10.1016/j.ajo.2018.05.024

PubMed Abstract | Crossref Full Text | Google Scholar

58. Owens WC, Owens EU. Retrolental fibroplasia in premature infants. Trans Am Acad Ophthalmol Otolaryngol. (1948) 53:18–41.18891013

PubMed Abstract | Google Scholar

59. Campbell K. Intensive oxygen therapy as a possible cause of retrolental fibroplasia: a clinical approach. Med J Aust. (1951) 2:48–50. doi: 10.5694/j.1326-5377.1951.tb109040.x

PubMed Abstract | Crossref Full Text | Google Scholar

60. Patz A, Hoeck LE, De la Cruz E. Studies on the effect of high oxygen administration in retrolental fibroplasia. I. Nursery observations. Am J Ophthalmol. (1952) 35:1248–53. doi: 10.1016/0002-9394(52)91140-9

PubMed Abstract | Crossref Full Text | Google Scholar

61. Cross KW. Cost of preventing retrolental fibroplasia? Lancet. (1973) 302:954–6. doi: 10.1016/S0140-6736(73)92610-X

Crossref Full Text | Google Scholar

62. Dawson JA, Kamlin CO, Vento M, Wong C, Cole TJ, Donath SM, et al. Defining the reference range for oxygen saturation for infants after birth. Pediatrics. (2010) 125(6):e1340–7. doi: 10.1542/peds.2009-1510

PubMed Abstract | Crossref Full Text | Google Scholar

63. Altuncu E, Ozek E, Bilgen H, Topuzoglu A, Kavuncuoglu S. Percentiles of oxygen saturations in healthy term newborns in the first minutes of life. Eur J Pediatr. (2008) 167(6):687–8. doi: 10.1007/s00431-007-0540-x

PubMed Abstract | Crossref Full Text | Google Scholar

64. Mariani G, Dik PB, Ezquer A, Aguirre A, Esteban ML, Perez C, et al. Pre-ductal and post-ductal O2 saturation in healthy term neonates after birth. J Pediatr. (2007) 150(4):418–21. doi: 10.1016/j.jpeds.2006.12.015

PubMed Abstract | Crossref Full Text | Google Scholar

65. Ankola P, DiAvanti J, Hernandez W, Fernandez L. Pulse oximetry measurements in newborn infants immediately after birth. Presented at the annual meeting of Pediatric Academic Societies; May 2-5, 2009; Baltimore, MD.

66. Kariuki E, Sutton C, Leone TA. Neonatal resuscitation: current evidence and guidelines. BJA Educ. (2021) 21(12):479–85. doi: 10.1016/j.bjae.2021.07.008

PubMed Abstract | Crossref Full Text | Google Scholar

67. Klaus M, Meyer BP. Oxygen therapy for the newborn. Pediatr Clin N Am. (1966) 13:731–52. doi: 10.1016/S0031-3955(16)31880-6

Crossref Full Text | Google Scholar

68. Kerber R, Ornato J, Brown D, Chameides L, Chandra N, Cummins R, et al. Emergency cardiac care committee and subcommittees, American Heart Association. JAMA. (1992) 268:2171. doi: 10.1001/jama.1992.03490160041023

PubMed Abstract | Crossref Full Text | Google Scholar

69. Saugstad OD. New guidelines for newborn resuscitation—a critical evaluation. Acta Paediatr. (2011) 100(8):1058–62. doi: 10.1111/j.1651-2227.2011.02301.x

PubMed Abstract | Crossref Full Text | Google Scholar

70. Ramji S, Ahuja S, Thirupuram S, Rootwelt T, Rooth G, Saugstad OD. Resuscitation of asphyxic newborn infants with room air or 100% oxygen. Pediatr Res. (1993) 34:809. doi: 10.1203/00006450-199312000-00023

PubMed Abstract | Crossref Full Text | Google Scholar

71. Saugstad OD, Rootwelt T, Aalen O. Resuscitation of asphyxiated newborn infants with room air or oxygen: an international controlled trial: the resair 2 study. Pediatrics. (1998) 102:e1. doi: 10.1542/peds.102.1.e1

PubMed Abstract | Crossref Full Text | Google Scholar

72. Saugstad OD, Ramji S, Irani SF, El-Meneza S, Hernandez EA, Vento M, et al. Resuscitation of newborn infants with 21% or 100% oxygen: follow-up at 18 to 24 months. Pediatrics. (2003) 112:296–300. doi: 10.1542/peds.112.2.296

PubMed Abstract | Crossref Full Text | Google Scholar

73. Vento M, Asensi M, Sastre J, Garcıía-Sala F, Pallardó FV, Viña J. Resuscitation with room air instead of 100% oxygen prevents oxidative stress in moderately asphyxiated term neonates. Pediatrics. (2001) 107:642–7. doi: 10.1542/peds.107.4.642

PubMed Abstract | Crossref Full Text | Google Scholar

74. Vento M, Asensi M, Sastre J, García-Sala F, Viña J. Six years of experience with the use of room air for the resuscitation of asphyxiated newly born term infants. Biol Neonate. (2001) 79:261–7. doi: 10.1159/000047103

PubMed Abstract | Crossref Full Text | Google Scholar

75. Saugstad OD, Ramji S, Soll RF, Vento M. Resuscitation of newborn infants with 21% or 100% oxygen: an updated systematic review and meta-analysis. Neonatology. (2008) 94:176–82. doi: 10.1159/000143397

PubMed Abstract | Crossref Full Text | Google Scholar

76. Saugstad OD, Vento M, Ramji S, Howard D, Soll RF. Neurodevelopmental outcome of infants resuscitated with air or 100% oxygen: a systematic review and meta-analysis. Neonatology. (2012) 102:98–103. doi: 10.1159/000333346

PubMed Abstract | Crossref Full Text | Google Scholar

77. Saugstad OD. Mechanisms of tissue injury by oxygen radicals: implications for neonatal disease. Acta Paediatr. (1996) 85(1):1–4. doi: 10.1111/j.1651-2227.1996.tb13880.x

PubMed Abstract | Crossref Full Text | Google Scholar

78. Nijima S, Shortland DB, Levene MI, Evans DH. Transient hyperoxia and cerebral blood flow velocity in infants born prematurely and at full term. Arch Dis Child. (1988) 63:1126–30. doi: 10.1136/adc.63.10_Spec_No.1126

PubMed Abstract | Crossref Full Text | Google Scholar

79. Lundstrom K, Pryds O, Greisen G. Oxygen at birth and prolonged cerebral vasoconstriction in preterm infants. Arch Dis Child. (1995) 73:F81–6. doi: 10.1136/fn.73.2.F81

Crossref Full Text | Google Scholar

80. International Liaison Committee on Resuscitation. 2005 International consensus on cardiopulmonary resuscitation and emergency cardiovascular care science with treatment recommendations. Part 7: neonatal resuscitation. Resuscitation. (2005) 67:293–303. doi: 10.1016/j.resuscitation.2005.09.014

PubMed Abstract | Crossref Full Text | Google Scholar

81. Welsford M, Nishiyama C, Shortt C, Isayama T, Dawson JA, Weiner G, et al. Room air for initiating term newborn resuscitation: a systematic review with meta-analysis. Pediatrics. (2019) 143:e20181825. doi: 10.1542/peds.2018-1825

PubMed Abstract | Crossref Full Text | Google Scholar

82. Escobedo MB, Aziz K, Kapadia VS, Lee HC, Niermeyer S, Schmölzer GM, et al. 2019 American Heart Association focused update on neonatal resuscitation: an update to the American Heart Association guidelines for cardiopulmonary resuscitation and emergency cardiovascular care. Pediatrics. (2020) 145:e922–30. doi: 10.1542/peds.2019-1362

Crossref Full Text | Google Scholar

83. Wyckoff MH, Wyllie J, Aziz K, de Almeida MF, Fabres J, Fawke J, et al. Neonatal life support: 2020 international consensus on cardiopulmonary resuscitation and emergency cardiovascular care science with treatment recommendations. Circulation. (2020) 142:S185–221. doi: 10.1161/CIR.0000000000000895

PubMed Abstract | Crossref Full Text | Google Scholar

84. Wang CL, Anderson C, Leone TA, Rich W, Govindaswami B, Finer NN. Resuscitation of preterm neonates by using room air or 100% oxygen. Pediatrics. (2008) 121:1083–89. doi: 10.1542/peds.2007-1460

PubMed Abstract | Crossref Full Text | Google Scholar

85. Rabi Y, Singhal N, Nettel-Aguirre A. Room-air versus oxygen administration for resuscitation of preterm infants: the ROAR study. Pediatrics. (2011) 128:e374–81. doi: 10.1542/peds.2010-3130

PubMed Abstract | Crossref Full Text | Google Scholar

86. Oei JL, Vento M, Rabi Y, Wright I, Finer N, Rich W, et al. Higher or lower oxygen for delivery room resuscitation of preterm infants below 28 completed weeks gestation: a meta-analysis. Arch Dis Child Fetal Neonatal Ed. (2017) 102(1):F24–30. doi: 10.1136/archdischild-2016-310435

PubMed Abstract | Crossref Full Text | Google Scholar

87. Wyllie J, Perlman JM, Kattwinkel J, Atkins DL, Chameides L, Goldsmith JP, et al. Part 11: neonatal resuscitation: 2010 international consensus on cardiopulmonary resuscitation and emergency cardiovascular care science with treatment recommendations. Resuscitation. (2010) 81(Suppl 1):e260–87. doi: 10.1016/j.resuscitation.2010.08.029

PubMed Abstract | Crossref Full Text | Google Scholar

88. Lui K, Jones LJ, Foster JP, Davis PG, Ching SK, Oei JL, et al. Lower versus higher oxygen concentrations titrated to target oxygen saturations during resuscitation of preterm infants at birth. Cochrane Database Syst Rev. (2018) 2018:CD010239. doi: 10.1002/14651858.CD010239.pub2

Crossref Full Text | Google Scholar

89. Dekker J, Martherus T, Lopriore E, Giera M, McGillick EV, Hutten J, et al. The effect of initial high vs. low FiO2 on breathing effort in preterm infants at birth: a randomized controlled trial. Front Pediatr. (2019) 7:504. doi: 10.3389/fped.2019.00504

PubMed Abstract | Crossref Full Text | Google Scholar

90. Sotiropoulos JX, Schmölzer GM, Oei JL, Libesman S, Hunter KE, Williams JG, et al. PROspective meta-analysis of trials of initial oxygen in preterm newborns (PROMOTION): protocol for a systematic review and prospective meta-analysis with individual participant data on initial oxygen concentration for resuscitation of preterm infants. Acta Paediatr. (2023) 112(3):372–82. doi: 10.1111/apa.16622

PubMed Abstract | Crossref Full Text | Google Scholar

91. Oei JL, Saugstad OD, Lui K, Wright IM, Smyth JP, Craven P, et al. Targeted oxygen in the resuscitation of preterm infants, a randomized clinical trial. Pediatrics. (2017) 139(1):e20161452. doi: 10.1542/peds.2016-1452

PubMed Abstract | Crossref Full Text | Google Scholar

92. Oei JL, Finer NN, Saugstad OD, Wright IM, Rabi Y, Tarnow-Mordi W, et al. Outcomes of oxygen saturation targeting during delivery room stabilisation of preterm infants. Arch Dis Childhood Fetal Neonatal Ed. (2018) 103:F446–54. doi: 10.1136/archdischild-2016-312366

Crossref Full Text | Google Scholar

93. Thamrin V, Saugstad OD, Tarnow-Mordi W, Wang YA, Lui K, Wright IM, et al. Preterm infant outcomes after randomization to initial resuscitation with FiO2 0.21 or 1.0. J. Pediatr. (2018) 201:55–61.e1. doi: 10.1016/j.jpeds.2018.05.053

PubMed Abstract | Crossref Full Text | Google Scholar

94. Kapadia V, Oei JL, Finer N, Rich W, Rabi Y, Wright IM, et al. Outcomes of delivery room resuscitation of bradycardic preterm infants: a retrospective cohort study of randomised trials of high vs. Low initial oxygen concentration and an individual patient data analysis. Resuscitation. (2021) 167:209–17. doi: 10.1016/j.resuscitation.2021.08.023

PubMed Abstract | Crossref Full Text | Google Scholar

95. Silverman WA. Oxygen therapy and retrolental fibroplasia. Am J Public Health Nations Health. (1968) 58:2009–11. doi: 10.2105/AJPH.58.11.2009

PubMed Abstract | Crossref Full Text | Google Scholar

96. Bolton DP, Cross KW. Further observations on cost of preventing retrlolental fibroplasia. Lancet. (1974) 1(7855):445–8. doi: 10.1016/S0140-6736(74)92395-2

PubMed Abstract | Crossref Full Text | Google Scholar

97. Kinsey VE, Arnold HJ, Kalina RE, Stern L, Stahlman M, Odell G, et al. PaO2 levels and retrolental fibroplasia: a report of the cooperative study. Pediatrics. (1977) 60(5):655–68. doi: 10.1542/peds.60.5.655

PubMed Abstract | Crossref Full Text | Google Scholar

98. Duc G, Bucher HU, Micheli JL, Johanson BA. Is transcutaneous PO2, reliable for arterial oxygen monitoring in newborn infants? Pediatrics. (1975) 55(4):566–7. doi: 10.1542/peds.55.4.566c

PubMed Abstract | Crossref Full Text | Google Scholar

99. Deckardt R, Steward DJ. Noninvasive arterial hemoglobin oxygen saturation versus transcutaneous oxygen tension monitoring in the preterm infant. Crit Care Med. (1985) 12(11):935–9. doi: 10.1097/00003246-198411000-00002

Crossref Full Text | Google Scholar

100. Peevy KJ, Hall MW. Transcutaneous oxygen monitoring: economic impact on neonatal care. Pediatrics. (1985) 75(6):1065–7. doi: 10.1542/peds.75.6.1065

PubMed Abstract | Crossref Full Text | Google Scholar

101. American Academy of Pediatrics and the American College of Obstetricians and Gynecologists. Guidelines for Perinatal Care. 6th ed Elk Grove Village, IL: American Academy of Pediatrics (2007).

Google Scholar

102. The STOP-ROP Multicenter Study Group. Supplemental therapeutic oxygen for prethreshold retinopathy of prematurity (STOP-ROP), a randomized, controlled trial. I: primary outcomes. Pediatrics. (2000) 105(2):295–310. doi: 10.1542/peds.105.2.295

PubMed Abstract | Crossref Full Text | Google Scholar

103. Askie LM, Henderson-Smart DJ, Irwig L. Simpson JM: oxygen-saturation targets and outcomes in extremely preterm infants. N Engl J Med. (2003) 349:959–67. doi: 10.1056/NEJMoa023080

PubMed Abstract | Crossref Full Text | Google Scholar

104. Askie LM, Brocklehurst P, Darlow BA, Finer N, Schmidt B, Tarnow-Mordi W, NeOProM collaborative group. NeOProM: neonatal oxygenation prospective meta-analysis collaboration study protocol. BMC Pediatr. (2011) 11:6. doi: 10.1186/1471-2431-11-6

PubMed Abstract | Crossref Full Text | Google Scholar

105. Askie LM, Darlow BA, Finer N, Schmidt B, Stenson B, Tarnow-Mordi W, et al. Association between oxygen saturation targeting and death or disability in extremely preterm infants in the neonatal oxygenation prospective meta-analysis collaboration. JAMA. (2018) 319:2190–201. doi: 10.1001/jama.2018.5725

PubMed Abstract | Crossref Full Text | Google Scholar

106. Schmidt B, Whyte RK. Oxygen saturation target ranges and alarm settings in the NICU: what have we learnt from the neonatal oxygenation prospective meta-analysis (NeOProM)? Semin Fetal Neonatal Med. (2020) 25:101080. doi: 10.1016/j.siny.2020.101080

PubMed Abstract | Crossref Full Text | Google Scholar

107. Askie LM. Meta-Analysis of oxygenation saturation targeting trials: do infant subgroups matter? Clin Perinatol. (2019) 46:579–91. doi: 10.1016/j.clp.2019.05.003

PubMed Abstract | Crossref Full Text | Google Scholar

108. Saugstad OD, Aune D. Optimal oxygenation of extremely low birth weight infants: a meta-analysis and systematic review of the oxygen saturation target studies. Neonatology. (2014) 105:55–63. doi: 10.1159/000356561

PubMed Abstract | Crossref Full Text | Google Scholar

109. Manja V, Saugstad OD, Lakshminrusimha S. Oxygen saturation targets in preterm infants and outcomes at 18–24 months: a systematic review. Pediatrics. (2017) 139:e20161609. doi: 10.1542/peds.2016-1609

PubMed Abstract | Crossref Full Text | Google Scholar

110. Manja V, Lakshminrusimha S, Cook DJ. Oxygen saturation target range for extremely preterm infants: a systematic review and meta-analysis. JAMA Pediatr. (2015) 169:332–40. doi: 10.1001/jamapediatrics.2014.3307

PubMed Abstract | Crossref Full Text | Google Scholar

111. Allen J. Photoplethysmography and its application in clinical physiological measurement. Physiol Meas. (2007) 28:R1–R39. doi: 10.1088/0967-3334/28/3/R01

PubMed Abstract | Crossref Full Text | Google Scholar

112. Tin W, Lal M. Principles of pulse oximetry and its clinical application in neonatal medicine. Semin Fetal Neonatal Med. (2015) 20(3):192–7. doi: 10.1016/j.siny.2015.01.006

PubMed Abstract | Crossref Full Text | Google Scholar

113. Hay WW Jr. The uses, benefits, and limitations of pulse oximetry in neonatal medicine: consensus on key issues. J Perinatol. (1987) 7(4):347–9.2463351

PubMed Abstract | Google Scholar

114. Wassuna A, Whitelaw AG. Pulse oximetry in preterm infants. Arch Dis Child. (1987) 62:957–8. doi: 10.1136/adc.62.9.957

PubMed Abstract | Crossref Full Text | Google Scholar

115. Moyle JTB. Uses and abuses of pulse oximetry. Arch Dis Child. (1996) 74:77–80. doi: 10.1136/adc.74.1.77

PubMed Abstract | Crossref Full Text | Google Scholar

116. Ahmed SJ, Rich W, Finer NN. The effect of averaging time on oximetry values in the premature infant. Pediatrics. (2010) 125(1):e115–21. doi: 10.1542/peds.2008-1749

PubMed Abstract | Crossref Full Text | Google Scholar

117. Vagedes J, Poets CF, Dietz K. Averaging time, desaturation level, duration and extent. Arch Dis Child Fetal Neonatal Ed. (2013) 98(3):F265–6. doi: 10.1136/archdischild-2012-302543

PubMed Abstract | Crossref Full Text | Google Scholar

118. van der Eijk AC, Dankelman J, Schutte S, Simonsz HJ, Smit BJ. An observational study to quantify manual adjustments of the inspired oxygen fraction in extremely low birth weight infants. Acta Paediatr. (2012) 101(3):e97–104. doi: 10.1111/j.1651-2227.2011.02506.x

PubMed Abstract | Crossref Full Text | Google Scholar

119. Nelson C, Bancalari E. Automated closed loop control of inspired oxygen concentration. Respir Care. (2013) 58(1):151–9. doi: 10.4187/respcare.01955

PubMed Abstract | Crossref Full Text | Google Scholar

120. Beddis IR, Collins P, Levy NM, Godfrey S, Silverman M. New technique for servo-control of arterial oxygen tension in preterm infants. Arch Dis Child. (1979) 54(4):278–80. doi: 10.1136/adc.54.4.278

PubMed Abstract | Crossref Full Text | Google Scholar

121. Dugdale RE, Cameron RG, Lealman GT. Closed-loop control of the partial pressure of arterial oxygen in neonates. Clin Physics Physiol Meas. (1988) 9(4):291–305. doi: 10.1088/0143-0815/9/4/001

Crossref Full Text | Google Scholar

122. Bhutani VK, Taube JC, Antunes MJ, Delivoria-Papadopoulos M. Adaptive control of the inspired oxygen delivery to the neonate. Pediatr Pulmonol. (1992) 14(2):110–7. doi: 10.1002/ppul.1950140209

PubMed Abstract | Crossref Full Text | Google Scholar

123. Morozoff PE, Evans RW. Closed-loop control of SaO2 in the neonate. Biomed Instrum Technol. (1992) 26(2):117–23.1562825

PubMed Abstract | Google Scholar

124. Sun Y, Kohane IS, Stark AR. Computer-assisted adjustment of inspired oxygen concentration improves control of oxygen saturation in newborn infants requiring mechanical ventilation. J Pediatr. (1997) 131(5):754–6. doi: 10.1016/S0022-3476(97)70107-0

PubMed Abstract | Crossref Full Text | Google Scholar

125. Morozoff EP, Smyth JA. Evaluation of three automatic oxygen therapy control algorithms on ventilated low birth weight neonates. Conf Proc IEEE Eng Med Biol Soc. (2009) 2009:3079–82. doi: 10.1109/IEMBS.2009.5332532

Crossref Full Text | Google Scholar

126. Claure N, Gerhardt T, Everett R, Musante G, Herrera C, Bancalari E. Closed-loop controlled inspired oxygen concentration for mechanically ventilated very low birth weight infants with frequent episodes of hypoxemia. Pediatrics. (2001) 107(5):1120–4. doi: 10.1542/peds.107.5.1120

PubMed Abstract | Crossref Full Text | Google Scholar

127. Urschitz MS, Horn W, Seyfang A, Hallenberger A, Herberts T, Miksch S, et al. Automatic control of the inspired oxygen fraction in preterm infants: a randomized crossover trial. Am J Respir Crit Care Med. (2004) 170(10):1095–100. doi: 10.1164/rccm.200407-929OC

PubMed Abstract | Crossref Full Text | Google Scholar

128. Claure N, D'Ugard C, Bancalari E. Automated adjustment of inspired oxygen in preterm infants with frequent fluctuations in oxygenation: a pilot clinical trial. J Pediatr. (2009) 155(5):640–5. doi: 10.1016/j.jpeds.2009.04.057

PubMed Abstract | Crossref Full Text | Google Scholar

129. Claure N, Bancalari E, D'Ugard C, Nelin L, Stein M, Ramanathan R, et al. Multicenter crossover study of automated adjustment of inspired oxygen in mechanically ventilated preterm infants. Pediatrics. (2011) 127(1):e76–83. doi: 10.1542/peds.2010-0939

PubMed Abstract | Crossref Full Text | Google Scholar

130. Dijkman KP, Mohns T, Dieleman JP, van Pul C, Goos TG, Reiss IKM, et al. Predictive intelligent control of oxygenation (PRICO) in preterm infants on high flow nasal cannula support: a randomised cross-over study. Arch Dis Child Fetal Neonatal Ed. (2021) 106:621–6. doi: 10.1136/archdischild-2020-320728

PubMed Abstract | Crossref Full Text | Google Scholar

131. Dargaville PA, Marshall AP, Ladlow OJ, Bannink C, Jayakar R, Eastwood-Sutherland C, et al. Automated control of oxygen titration in preterm infants on non-invasive respiratory support. Arch Dis Child Fetal Neonatal Ed. (2022) 107:39–44. doi: 10.1136/archdischild-2020-321538

PubMed Abstract | Crossref Full Text | Google Scholar

132. Sturrock S, Ambulkar H, Williams EE, Sweeney S, Bednarczuk NF, Dassios T, et al. A randomised crossover trial of closed loop automated oxygen control in preterm, ventilated infants. Acta Paediatr. (2021) 110(3):833–7. doi: 10.1111/apa.15585

PubMed Abstract | Crossref Full Text | Google Scholar

133. Schouten TMR, Abu-Hanna A, van Kaam AH, van den Heuvel MEN, Bachman TE, van Leuteren RW, et al. Prolonged use of closed-loop inspired oxygen support in preterm infants: a randomised controlled trial. Arch Dis Child Fetal Neonatal Ed. (2024) 109(2):221–6. doi: 10.1136/archdischild-2023-325831

PubMed Abstract | Crossref Full Text | Google Scholar

134. Salverda HH, Oldenburger NJ, Rijken M, Ratna R, Tan NGB, te Pas AB, et al. Automated oxygen control for very preterm infants and neurodevelopmental outcome at 2 years−a retrospective cohort study. Eur J Pediatr. (2023) 182:1593–9. doi: 10.1007/s00431-023-04809-4

PubMed Abstract | Crossref Full Text | Google Scholar

135. Maiwald CA, Niemarkt HJ, Poets CF, Urschitz MS, Konig J, Hummler H, et al. Effects of closed-loop automatic control of the inspiratory fraction of oxygen (FiO2-C) on outcome of extremely preterm infants—study protocol of a randomized controlled parallel group multicenter trial for safety and efficacy. BMC Pediatr. (2019) 19(1):363. doi: 10.1186/s12887-019-1735-9

PubMed Abstract | Crossref Full Text | Google Scholar

136. Kaltsogianni O, Dassios T, Greenough A. Does closed-loop automated oxygen control reduce the duration of mechanical ventilation? A randomised controlled trial in ventilated preterm infants. Trials. (2022) 23(1):276. doi: 10.1186/s13063-022-06222-y

PubMed Abstract | Crossref Full Text | Google Scholar

137. Sood BG, McLaughlin K, Cortez J. Near-infrared spectroscopy: applications in neonates. Semin Fetal Neonatal Med. (2015) 20(3):164–72. doi: 10.1016/j.siny.2015.03.008

PubMed Abstract | Crossref Full Text | Google Scholar

138. Jöbsis FF. Noninvasive, infrared monitoring of cerebral and myocardial oxygen sufficiency and circulatory parameters. Science. (1977) 198(4323):1264–7. doi: 10.1126/science.929199

Crossref Full Text | Google Scholar

139. Garvey AA, Pavel AM, Murray DM, Boylan GB, Dempsey EM. Does early cerebral near-infrared spectroscopy monitoring predict outcome in neonates with hypoxic ischaemic encephalopathy? A systematic review of diagnostic test accuracy. Neonatology. (2022) 119(1):1–9. doi: 10.1159/000518687

PubMed Abstract | Crossref Full Text | Google Scholar

140. Pichler G, Goeral K, Hammerl M, Perme T, Dempsey EM, Springer L, et al. Cerebral regional tissue oxygen saturation to guide oxygen delivery in preterm neonates during immediate transition after birth (COSGOD III): multicentre randomised phase 3 clinical trial. Br Med J. (2023) 380:e072313. doi: 10.1136/bmj-2022-072313

Crossref Full Text | Google Scholar

141. Bonsante F, Ramful D, Binquet C, Samperiz S, Daniel S, Gouyon J. Low renal oxygen saturation at near-infrared spectroscopy on the first day of life is associated with developing acute kidney injury in very preterm infants. Neonatology. (2019) 115(3):198. doi: 10.1159/000494462

PubMed Abstract | Crossref Full Text | Google Scholar

142. Chock VY, Frymoyer A, Yeh CG, VanMeurs KP. Renal saturation and acute kidney injury in neonates with hypoxic ischemic encephalopathy undergoing therapeutic hypothermia. J. Pediatr. (2018) 200:232–239.e1. doi: 10.1016/j.jpeds.2018.04.076

PubMed Abstract | Crossref Full Text | Google Scholar

143. Ruf B, Bonelli V, Balling G, Hörer J, Nagdyman N, Braun SL, et al. Intraoperative renal near-infrared spectroscopy indicates developing acute kidney injury in infants undergoing cardiac surgery with cardiopulmonary bypass: a case–control study. Crit Care. (2015) 19:27. doi: 10.1186/s13054-015-0760-9

PubMed Abstract | Crossref Full Text | Google Scholar

144. Patel AK, Lazar DA, Burrin DG, Smith EO, Magliaro TJ, Stark AR, et al. Abdominal near-infrared spectroscopy measurements are lower in preterm infants at risk for necrotizing enterocolitis. Pediatr Crit Care Med. (2014) 15(8):735–41. doi: 10.1097/PCC.0000000000000211

PubMed Abstract | Crossref Full Text | Google Scholar

145. Schat TE, van Zoonen AG, van der Laan ME, Mebius MJ, Bos AF, Hulzebos CV, et al. Early cerebral and intestinal oxygenation in the risk assessment of necrotizing enterocolitis in preterm infants. Early Hum Dev. (2019) 131:75–80. doi: 10.1016/j.earlhumdev.2019.03.001

PubMed Abstract | Crossref Full Text | Google Scholar

146. Tran NN, Tran M, Lopez J, Ogbaa M, Votava-Smith JK, Brady KM. Near-infrared spectroscopy: clinical use in high-risk neonates. Neonatal Netw. (2021) 40(2):73–9. doi: 10.1891/0730-0832/11-T-678

PubMed Abstract | Crossref Full Text | Google Scholar

147. Passera S, Boccazzi M, Bokobza C, Faivre V, Mosca F, Van Steenwinckel J, et al. Therapeutic potential of stem cells for preterm infant brain damage: Can we move from the heterogeneity of preclinical and clinical studies to established therapeutics? Biochem Pharmacol. (2021) 186:114461. doi: 10.1016/j.bcp.2021.114461

PubMed Abstract | Crossref Full Text | Google Scholar

Keywords: hyperoxia, hypoxia, oxygenation, neonate, preterm infant, retinopathy of prematurity (ROP), neonatal intensive care unit (NICU)

Citation: Mamidi RR and McEvoy CT (2024) Oxygen in the neonatal ICU: a complicated history and where are we now?. Front. Pediatr. 12:1371710. doi: 10.3389/fped.2024.1371710

Received: 16 January 2024; Accepted: 17 April 2024;
Published: 1 May 2024.

Edited by:

Mohamed E. Abdel-Latif, The Canberra Hospital, Australia

Reviewed by:

Michael Adam O'Reilly, University of Rochester, United States
Janneke Dekker, Leiden University, Netherlands
Maximo Vento, La Fe Hospital, Spain

© 2024 Mamidi and McEvoy. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Rachna R. Mamidi mamidi@ohsu.edu

Download