National Academies Press: OpenBook
« Previous: 2 NASA Risk Management and Health Standards
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 45
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 46
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 47
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 48
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 49
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 50
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 51
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 52
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 53
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 54
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 55
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 56
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 57
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 58
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 59
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 60
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 61
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 62
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 63
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 64
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 65
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 66
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 67
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 68
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 69
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 70
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 71
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 72
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 73
Suggested Citation:"3 Health Risks." Institute of Medicine. 2014. Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework. Washington, DC: The National Academies Press. doi: 10.17226/18576.
×
Page 74

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

3 Health Risks In considering the ethics issues that will emerge when making deci- sions about sending humans into harm’s way on long duration and explo- ration spaceflights, the committee decided to examine some of the health risks that illustrate key ethical challenges and tensions in risk and deci- sion making. It is important to note that the stressors and health risks may vary from mission to mission, depending on the remoteness of the destination and many other factors. For example, missions to locations that will require months to years to return to Earth can have fundamental- ly different stressors than destinations that are closer but may involve being in space for long periods of time. Although the phrase “long dura- tion and exploration spaceflights” is used throughout the report, the committee acknowledges that the health risks may vary widely between missions. These examples are meant to illustrate the diversity of risks and, therefore, highlight some of the ethics quandaries arising in decisions regarding uncertain risks, unknowns about the extent of individual varia- tions, and long-term health implications for astronauts. In addition, the examples also demonstrate the range of risk management strategies that are being employed by or considered by the National Aeronautics and Space Administration (NASA), international partners, and commercial companies to prevent, mitigate, or treat these risks to human health. The following sections are brief overviews of the complex issues relevant to: • the risk of loss of life due to mission operations, • vision impairment risks, • behavioral health and performance risks, 45

46 LONG DURATION AND EXPLORATION SPACEFLIGHT • bone demineralization, and • radiation exposure. MISSION OPERATIONS RISKS Although there have been many successful space missions over the past 50 years, the substantial risks undertaken by astronauts are most ev- ident in documented “near misses” and tragic losses of life. During the Apollo 11 mission, Neil Armstrong piloted a lunar module to the Moon’s surface, landing with less than 30 seconds worth of fuel remaining (Garber and Launius, 2005). Apollo 12 was struck by lightning at ap- proximately 36 and 52 seconds after launch, momentarily shutting down the electrical power (Molloy and Petrone, 2013). During the Apollo 13 mission, an oxygen tank exploded while en route to the Moon, and ground crews and astronauts had to improvise to safely return the crew to Earth (Garber and Launius, 2005). The 1997 collision of a resupply vehi- cle, as well as a fire, jeopardized the lives of crew members aboard the Mir Space Station (NASA, 2014b). In 1967, a cabin fire in an Apollo capsule during a launch pad test in Cape Canaveral, Florida, killed three astronauts.1 Over the course of the NASA space program, 24 crew mem- bers have lost their lives in the line of duty, including the 14 individuals aboard the space shuttles Challenger and Columbia, the Apollo 1 astro- nauts, and 5 individuals who died in training-jet crashes (NASA, 2014a). Due to the nature of the propulsion systems needed to launch, the dis- tance from Earth, the immediacy of some inflight catastrophes and the challenges of landing, options for rescue operations are limited. The probability of loss of crew due to a catastrophic vehicle accident is generally higher and accompanied by greater uncertainty in early expe- rience with new crew delivery vehicles. As experience is gained through effective use of continuous risk management systems, the expected re- sults are decreased risk and less uncertainty, although risks remain high. Thus, for example, the probabilistic risk assessment for loss of crew and vehicle on the first shuttle flight was 1/10 versus 1/90 on the 135th shut- 1 On January 27, 1967, the crew of Roger B. Chaffee, Virgil “Gus” Grissom, and Edward H, White, Jr., were killed in a fire that spread quickly in a pure oxygen environ- ment; the astronauts had no opportunity to open the hatch (Garber and Launius, 2005; Williams, 2011).

HEALTH RISKS 47 tle flight (Behnken et al., 2013). The shuttle mission aggregate risk of loss of crew and vehicle was approximately 1/46 (Behnken et al., 2013). The risks of the loss of the crew and vehicle during a spaceflight mission are calculated to allow comparisons of different approaches and platforms using probabilistic risk assessments. NASA provides detailed briefings for astronaut candidates on mission risks. Comparisons are made in this briefing to the risk of death in other high-risk occupations and to catastrophic events and combat. Astronauts who provided infor- mation to the committee emphasized the thorough and ongoing nature of communication throughout their careers regarding risks. VISION IMPAIRMENT Vision impairment represents a newly identified health risk and ex- emplifies how NASA approaches a new risk in the current risk manage- ment framework. The potential for vision changes has both acute and long-term implications for the astronaut and for current and future missions. Overview and Risk Identification Although astronauts have reported vision changes during spaceflight for more than 40 years (Alexander et al., 2012), these were assumed to be transient and isolated. With the Mir Space Station and the Internation- al Space Station (ISS), longer tours of duty became possible. A Mir tech- nical report from 2008 described disc edema in 8 of 16 cosmonauts on landing, with one magnetic resonance imaging (MRI) report showing signs suggestive of intracranial hypertension (as cited in Alexander et al., 2012). The spaceflight environment on Mir was noted to be similar to the ISS. More significant and lasting visual changes have occurred in astro- nauts on the ISS and have been documented in published case reports (Mader et al., 2011). These changes were primarily a shift toward hyper- opia (far-sightedness). Scotomas (blind spots in the visual field) were also reported. One astronaut reported that he needed to shift his head to compensate for scotoma when reading instruments during the mission (Alexander et al., 2012). In a group of astronauts examined pre- and postflight, clinicians observed flattening of the eyeball globe using ultra- sound and MRI, supporting the hyperopic shift, edema of the optic nerve, and changes of the optic fundus noted on examination (Mader et al., 2011). In this discussion of the detailed findings on seven astronauts after

48 LONG DURATION AND EXPLORATION SPACEFLIGHT spaceflight, the results are presented in an “unattributable” manner and do not include dates or missions to preserve privacy and confidentiality (Mader et al., 2011). Refractive changes have persisted after return to Earth for these astronauts. Upon identifying the vision-related issues, NASA rapidly imple- mented pre- and postflight ocular testing protocols and convened a Papilledema Summit in July 2009 to bring together experts in space med- icine and terrestrial analogs and to suggest avenues for research (Watkins and Barr, 2010). The summit examined potential physiological causes of the vision alterations and suggested changes in preflight, inflight, and postflight testing. Because the optic changes bore a striking similarity to those seen with elevated intracranial pressure (i.e., papilledema), this hypothesis was the first to be extensively explored (Alexander et al., 2012). Lumbar puncture has been performed on four U.S. astronauts (at single intervals ranging from 12 to 60 days after flight), and opening pressures were found to be elevated in two individuals and borderline in the other two (Alexander et al., 2012). Other hypotheses regarding causes of vision alterations include elevated carbon dioxide in the ISS atmosphere, in- creased sodium content of the astronaut diet, fluid shifts due to increases in resistance exercises, radiation exposure, and individual susceptibility (Alexander et al., 2012). Since 1989, as part of postflight eye examinations, U.S. astronauts have been asked about improvements or degradations in their distant or near vision. Compilation of that information on 300 U.S. astronauts who had flown in space found that approximately 29 percent of shuttle astro- nauts and 60 percent of ISS astronauts noted a degradation in distant and near visual acuity, and some of these changes remained unresolved years after flight (Mader et al., 2011). With the exception of the report from Mir previously cited, data from international partners are lacking. Risk Management: Clinical Practice Guidelines, Surveillance, and Research NASA is monitoring vision changes and intracranial pressure, and research efforts in this area are fully under way. NASA has developed clinical practice guidelines for treatment of astronauts with postflight refrac- tive changes, which include a classification system from class 0 (least severe) through class 4 (most severe) based on the results of imaging studies

HEALTH RISKS 49 and indicating the follow-up testing and monitoring that are required (Alexander et al., 2012). In 2012, the NASA Human Research Program released the evidence report Risk of Spaceflight-Induced Intracranial Hypertension and Vision Alterations, which summarizes research evidence and raises questions for future study (Alexander et al., 2012). Current research efforts are ex- ploring ground-based analogs and include studies of normal individuals in supine and head-down bed rest, hind-limb-suspended rodent models, and various pathologic situations in which elevated intracranial hyperten- sion and papilledema occur. The evidence report details four major gaps requiring future research: • Etiological mechanisms and contributing risk factors for ocular structural and functional changes seen inflight and postflight; • Validated and minimally obtrusive diagnostic tools to measure and monitor changes in intracranial pressure, ocular structure, and ocular function; • Ground-based analogs and/or models to simulate the spaceflight- associated visual impairment and increased intracranial pres- sures; and • Preventive and treatment measures to mitigate changes in ocular structure and function and intracranial pressure during space- flight (Alexander et al., 2012). Individual Variation Issues As with many health risks, individual factors (including, but not lim- ited to, age and sex) may account for some of the variability in manifes- tation of vision and ocular alterations. The evidence report on this risk concludes with this statement, “In summary, 15 long-duration male as- tronauts ranging in age from 45 to 55 years have experienced confirmed inflight and postflight visual and anatomical changes” (Alexander et al., 2012, p. 87). Health Standards and Risk Profile A NASA health standard has not yet been developed for vision im- pairment and increased intracranial pressure and the extent, if any, to which these are linked is not fully known (IOM, 2014). At present, this

50 LONG DURATION AND EXPLORATION SPACEFLIGHT risk is listed as “unacceptable” on the Human Research Roadmap (see Table 2-2) for missions to the ISS for 12 months, to a near-Earth aster- oid, or to Mars, and “insufficient data” for a lunar mission (NASA, 2013). Issues for Long Duration and Exploration Spaceflights In 2013, Mader and colleagues reported on a single astronaut who was carefully studied after two long-duration missions, documenting progressive vision changes associated with repeat spaceflight, raising concerns that changes occurring during the first flight “may have set the stage for recurrent or additional changes when the astronaut was subject- ed to physiological stress of repeat space flight” (Mader et al., 2013, p. 249). The prevalence and severity of the observed visual changes raised significant concerns about the ability of an affected pilot to successfully land a spacecraft. This is an active area of research with many remaining unknowns that could affect future crew selection, mission success, and astronaut health. Ethics issues raised by this example focus on how best to address uncertainty and unknowns including determinations of long-term respon- sibilities for monitoring, preventing, and treating the health condition moving forward. Issues of individual susceptibility are still being deter- mined, but this example points to the importance of diverse participation, including women, to obtain population-based information (as all exam- ples to date of vision problems have occurred in men). BEHAVIORAL HEALTH AND PERFORMANCE RISKS NASA has identified three categories of behavioral health and perfor- mance risks associated with long duration and exploration spaceflight: (1) adverse behavioral conditions and psychiatric disorders; (2) performance errors due to fatigue resulting from sleep loss, circadian desynchronization, extended wakefulness, and work overload; and (3) performance decre- ments due to inadequate cooperation, coordination, communication, and psychosocial adaptation within a Team Gap (Schmidt et al., 2009; Slack et al., 2009; Whitmire et al., 2009).

HEALTH RISKS 51 Overview and Risk Identification Much of the evidence regarding behavioral health risks associated with long duration spaceflight is derived from anecdotal evidence (Aldrin, 1973; Lebedev, 1988; Burrough, 1998; Linenger, 2000), archiv- al and observational data collected during spaceflight (Kanas et al., 2007; Stuster, 2010), and observational and experimental studies conducted in analog settings, such as polar expeditions (Gunderson, 1974; Sandal et al., 1996; Palinkas and Suedfeld, 2008), submarines (Tansey et al., 1979; Sandal et al., 1999; Thomas et al., 2000), and space simulations (Gushin et al., 1996; Sandal, 2001; Basner et al., 2013). The Mars 500 study, with participants from four nations, was a 17-month isolation experiment in preparation for a mission to Mars (ESA, 2011). While the reported incidence of behavioral health problems encoun- tered during spaceflight has been quite low, the actual incidence may be underestimated due to a reluctance of astronauts to report them (IOM, 2001; Shepanek, 2005). Billica (2000) reported a 2.86 per person-year incidence of such problems among the 508 crew members who flew on 89 space shuttle missions between 1981 and 1989. The most common behavioral symptoms reported by crew members were anxiety and irrita- bility. Data collected for 28.84 person-years of NASA spaceflight identi- fied 24 cases of anxiety, for an incidence rate of 0.832 cases per person- year (Slack et al., 2009). No behavioral emergencies have been reported to date (Slack et al., 2009). Studies in analog settings have reported much higher rates of behav- ioral health problems, reflecting longer periods of isolation and confine- ment and differences in crew member characteristics. For example, the incidence of behavioral health problems after extended stays in Antarcti- ca was estimated in one study at 5.2 percent (Palinkas et al., 2004). A number of factors may contribute to the risk of behavioral health and sleep problems in space, including mission duration (Slack et al., 2009), disruption of sleep and circadian rhythms (Czeisler et al., 1986; Dinges et al., 1997), physiological changes that occur in microgravity, workload, lack of social and environmental stimulation, cultural and organizational factors, family issues, and personality characteristics (Gunderson, 1974; McFadden et al., 1994; Rose et al., 1994; NRC, 1998; Rosnet et al., 2000). For the ISS, both observational (i.e., Stuster, 2010) and anecdotal evidence indicate that the crew has been successful at thriving.

52 LONG DURATION AND EXPLORATION SPACEFLIGHT Risk Management: Countermeasures and Research Analog and other types of studies have identified measures that can prevent or mitigate behavioral health, sleep impairment, and cognition risks. Countermeasures that have been successfully implemented for ISS operations include working with ground control regarding scheduling, support services from operational psychology personnel, multiple layers of accountability when conducting critical tasks, and education for ground crews (Flynn, 2005; Slack et al., 2009). Existing techniques for monitoring crew behavioral health and providing social and psychologi- cal support also have proved effective on the ISS. However, it is un- known whether the effectiveness of these countermeasures can be maintained over longer periods and at greater distances from Earth where delays in communication, delivery of services, and implementation of countermeasures will occur. Policies designed to minimize fatigue, indi- vidual stress, and interpersonal tension, such as allocated “time off,” will be critical to preserving and promoting the behavioral health of astronaut personnel. Astronauts also train in analog environments (including Antarctica, the underwater NASA Extreme Environment Mission Operations [NEEMO], and winter and mountain survival programs). The training environments include both physical capabilities and problem resolution involving group dynamics. The National Research Council (NRC) Committee for the Decadal Survey on Biological and Physical Sciences in Space noted that “contin- ued research is required to identify individual, interpersonal, cultural, and environmental determinants of crew cohesion, crew performance, and ground-crew interaction” (NRC, 2011, p. 89). Such research may lead to greater specificity of behavioral health standards. The NASA Human Research Program is working on addressing research gaps in this area (see Box 3-1). BOX 3-1 NASA Human Research Program’s Research Gaps Behavioral Health and Performance Risks BMed 1: What are the most effective methods to enhance behavioral health and prevent decrements before, during, and after spaceflight missions?

HEALTH RISKS 53 BMed 2: What are the most effective methods to predict, detect, and as- sess decrements in behavioral health (which may negatively affect per- formance) before, during, and after spaceflight missions? BMed 3: What aspects, if any, of cognitive performance change in-flight? If there are changes, do they persist post mission? If so, for how long? BMed 4: What are the most effective methods for detecting and assessing cognitive performance during exploration missions? BMed 5: What individual characteristics predict successful adaptation and performance in an isolated, confined, and extreme environment, espe- cially for long-duration missions? BMed 6: What are the most effective methods for treating the individual to remedy behavioral health problems during spaceflight missions (includ- ing behavioral health meds)? BMed 7: What are the most effective methods for modifying the environ- ment to prevent and remedy behavioral health problems during space- flight missions? BMed 8: How do family, friends, and colleagues affect astronauts’ behav- ioral health and performance before, during, and after spaceflight? Team Gap Risks Team Gap 1: We need to understand the key threats, indicators, and life cycle of the team for autonomous, long duration, and/or distance explo- ration missions. Team Gap 2: We need to identify a set of validated measures, based on the key indicators of team function, to effectively monitor and measure team health and performance fluctuations during autonomous, long du- ration, and/or distance exploration missions. Team Gap 3: We need to identify a set of countermeasures to support team function for all phases of autonomous, long duration, and/or dis- tance exploration missions. Team Gap 4: We need to identify psychological measures that can be used to select individuals most likely to maintain team function for au- tonomous, long duration, and/or distance exploration missions. Team Gap 5: We need to identify validated ground-based training meth- ods that can be both preparatory and continuing to maintain team func- tion in autonomous, long duration, and/or distance exploration missions. Team Gap 6: We need to identify methods to support and enable multiple distributed teams to manage shifting levels of autonomy during long du- ration, and/or distance exploration missions. Team Gap 8: We need to identify psychological and psychosocial factors, measures, and combinations thereof that can be used to compose high- ly effective crews for autonomous, long duration, and/or distance explo- ration missions. Team Gap 9: We need to identify spaceflight acceptable thresholds (or ranges) of team function, based on key indicators, for autonomous, long duration, and/or distance exploration missions.

54 LONG DURATION AND EXPLORATION SPACEFLIGHT Sleep and Cognition Risks Sleep Gap 1: We need to identify a set of validated and minimally obtru- sive tools to monitor and measure sleep-wake activity and associated performance changes for spaceflight. Sleep Gap 2: We need to understand the contribution of sleep loss, circa- dian desynchronization, extended wakefulness and work overload, on individual and team behavioral health and performance (including oper- ational performance), for spaceflight. Sleep Gap 4: We need to identify indicators of individual vulnerabilities and resiliencies to sleep loss and circadian rhythm disruption, to aid with individualized countermeasure regimens, for autonomous, long du- ration, and/or distance exploration missions. Sleep Gap 5: We need to identify environmental specifications and opera- tional regimens for using light to prevent and mitigate health and per- formance decrements due to sleep, circadian, and neurobehavioral disruption, for flight, surface and ground crews, during all phases of spaceflight operations. Sleep Gap 6: We need to identify how individual crew members can most ef- fectively and safely use medications to promote sleep, alertness, and circa- dian entrainment, as needed during all phases of spaceflight operations. Sleep Gap 8: We need to develop individualized scheduling tools that predict the effects of sleep-wake cycles, light and other countermeas- ures on performance, and can be used to identify optimal (and vulnera- ble) performance periods during spaceflight. Sleep Gap 9: We need to identify an integrated, individualized suite of countermeasures and protocols for implementing these countermeas- ures to prevent and/or treat chronic partial sleep loss, work overload, and/or circadian shifting, in spaceflight. Sleep Gap 10: We need to identify the spaceflight environmental and mis- sion factors that contribute to sleep decrements and circadian misa- lignment, and their acceptable levels of risk. NOTE: Team Gap 7 merged with Team Gap 3. Sleep Gap 3 is closed. Sleep Gap 7 merged with Sleep Gap 2. SOURCE: NASA, 2014c. Individual Variation Issues Several studies have pointed to phenotypic and genotypic variations to sleep disruption and its behavioral consequences (Van Dongen et al., 2004; Landholt, 2008; Kuna et al., 2012), suggesting the identification of predictive biomarkers (Czeisler, 2011) that could prove useful in identi- fying astronauts likely to experience sleep-related performance decre- ments and for managing sleep-wake regulation during exploration spaceflight (Goel and Dinges, 2012).

HEALTH RISKS 55 Individual characteristics identified as predictors of social compati- bility in analog studies and surveys of astronaut personnel include low extraversion and high introversion (Palinkas and Suedfeld, 2008), high positive instrumentality (goal oriented, active, self-confident) and ex- pressiveness (kind, aware of others’ feelings), and low negative instru- mentality (arrogant, hostile, boastful, egotistical) and communion (self- subordinating, subservient, unassertive) (McFadden et al., 1994; Rose et al., 1994). Crew heterogeneity with respect to social (e.g., age, sex, cultural background), psychological (need for achievement, aggressiveness, au- tonomy), and other characteristics (e.g., interest in leisure activities) also predicts crew cohesion and conflict, Team Gap decision making, and re- sponse to crisis (NRC, 1998, 2011; Kanas and Manzey, 2008; Kanas et al., 2009). Health Standards and Risk Profile According to NASA’s Human Research Roadmap, the research rat- ings for risk of adverse behavioral conditions and psychiatric disorders has been determined to be “Controlled” for the ISS and lunar missions, “Acceptable” for near-Earth asteroid missions, and “Unacceptable” for the Mars design reference mission (see Table 2-2) (NASA, 2013). The risk of fatigue-related performance errors is considered “Controlled” for all four design reference missions. The research rating for the third risk related to Team Gap issues is currently listed as “Controlled” for the lu- nar design reference mission and “Acceptable” for the ISS, near-Earth asteroid, and Mars missions. For a mission to a near-Earth asteroid or Mars, mission duration and distance from Earth are among the major stressors. A NASA health standard for behavioral health and perfor- mance has been developed that addresses all of these risks (see Box 2-1). Cognitive and psychiatric assessments and screening criteria are used in the initial astronaut selection process and in the annual astronaut re- certification medical examinations. Furthermore, processes have been developed by the ISS partner agencies that detail neurocognitive and be- havioral health baseline assessments and follow-up. Issues for Long Duration and Exploration Spaceflights Few established policies regarding limits to long-duration isolation and confinement exist. The National Science Foundation’s Division of

56 LONG DURATION AND EXPLORATION SPACEFLIGHT Polar Programs requires that all candidates for winter-over duty at Ant- arctica’s Amundsen-Scott South Pole and McMurdo stations undergo a psychiatric evaluation conducted by a civilian contractor. The U.S. Ant- arctic Program also places limits on the number of continuous seasons (e.g., summer and winter) a person can stay at the same station, essential- ly mandating that personnel either spend a subsequent year or part of a year at another station or go home for a season prior to returning. Astronauts need to be fully informed that individual and interperson- al issues that typically are of little concern on Earth or during short dura- tion missions have the potential to become clinically and operationally significant in conditions of prolonged isolation and confinement. While more is known about controlling the risk of fatigue-related performance errors for shorter flights, much remains to be learned about the chronic stressors of remote missions. Those missions will require autonomous operations that will change the ways crews conduct tasks, leaving them potentially more vulnerable to fatigue-related performance errors and more open to challenges regarding crew dynamics and cohesion. Ethics issues raised in considering these risks include appropriate- ness of crew selection criteria (e.g., genetic, sex, and cultural differ- ences), the necessity of informed decision making, and challenges and tensions between astronaut privacy and the need to continuously learn about the impact of spaceflight on behavior and performance. However, as with other areas of human performance, there remains a great detail of uncertainty in the field of clinical psychology related to the ability to predict individual behavior or group behavior on the basis of individual characteristics, especially in isolated and confined extreme environments (Palinkas and Suedfeld, 2008). Clinical assessments have proved to be more successful at “screening out” individuals who are unsuited to living and working in such environments; however, they have been shown to have less utility in “screening in” individuals who are best qualified for such assignments (Grant et al., 2007). Similarly, the adaptation of coun- termeasures found to be effective on Earth or on the ISS to long duration and exploration space missions remains controversial given their limited external validity due to context-specific influences (Shepanek, 2005). As with other health issues, establishing ethics principles for these missions must take into consideration the feasibility and desirability of doing so in the face of uncertainty.

HEALTH RISKS 57 BONE DEMINERALIZATION Bone demineralization during exposure to microgravity illustrates the multiple parameters of a set of related health risks; how these risks are assessed, studied, and managed by NASA; countermeasure develop- ment; and interaction with engineering systems. Overview and Risk Identification Bone demineralization in microgravity is a well-recognized and well-studied phenomenon that still is not completely understood. It over- laps significantly with areas of highly active research for terrestrial med- icine, but it is not fully known in what ways microgravity-induced bone loss might be similar to, or different from, osteoporosis. Bone normally remodels in response to physical load, becoming stronger where that load is stronger and weaker where there is less load. Approximately a 10th of the human skeleton is normally renewed annu- ally (Sibonga et al., 2008b). In microgravity, bone and muscle experience less loading, and bone mineral density loss rates of approximately 1 to 1.5 percent per month have been observed. This is compared to the 2 to 3 percent loss in bone mineral density per year in postmenopausal females during the first decade after the onset of menopause, which is considered a time of rapid bone loss (Sibonga et al., 2008b). Bone loss is primarily due to increased resorption of bone with subsequent release of calcium into the bloodstream and excretion by the kidneys. Risk Management: Countermeasures and Research Because of significant overlap with important clinical questions for terrestrial health care (e.g., osteoporosis and metabolic bone diseases), human and animal models, monitoring methods (e.g., biochemical mark- ers, bone ultrasound, and dual-energy x-ray absorptiometry scan), and a variety of countermeasures (pharmacologic, dietary, and exercise) have been developed and are being implemented for astronauts (Sibonga et al., 2008a,b). However, much remains to be learned about the effects of mi- crogravity, particularly for long duration exposures. Bone loss in microgravity is exacerbated by dietary factors. Docu- mented undernutrition (“space anorexia”), lack of adequate vitamin D, a diet relatively high in sodium (which increases kidney urine calcium concentration), and the particular relative composition of amino acids in

58 LONG DURATION AND EXPLORATION SPACEFLIGHT the diet have all been implicated as contributing factors and are all ame- nable to careful nutritional adjustment (Smith et al., 2012). Potential countermeasures include a combination of aerobic and re- sistive exercise, nutrition, vitamin D supplementation, and pharmacologic agents such as bisphosphonates. Most recent studies in a small number of astronauts on the ISS have combined optimal nutrition (adequate caloric intake, low dietary sodium, and appropriate amino acid balances) and vitamin D supplementation with an aggressive exercise program (Smith et al., 2012). These relatively successful countermeasures increase the rate of new bone formation, so bone mass is maintained, but there is in- creased bone-cell turnover. Pharmacologic agents that are used to combat bone loss associated with menopause (in women) and aging (in both men and women) have been studied to a very limited extent in the space envi- ronment. Both bone mass and bone architecture contribute to bone strength and fracture risk. New bone that is deposited as a result of coun- termeasures appears to exhibit altered architecture compared with the bone that has been lost (Sibonga et al., 2008b). Moreover, the increased calcium turnover may increase the risk of kidney stones.2 Drinking more water dilutes the calcium concentration and decreases this risk. Thus, demineralization poses risk to an individual astronaut during and after the mission, and risk to the mission if bone loss leads to frac- ture. A combination of countermeasures was tested on 13 crew members during several ISS missions and was associated with increased formation of new bone (without a corresponding decrease in resorption) and overall maintenance of bone mass (Smith et al., 2012). Although multiple knowledge gaps remain, one of particular interest is development of an easy, noninvasive way to monitor bone health in microgravity. Research gaps are identified in the NASA Human Re- search Program’s evidence reports on this health concern (Sibonga et al., 2008a,b). Individual Variation Issues The significant individual variation in bone demineralization is so wide as to obscure differences due to biological sex characteristics or any 2 An unexpected environmental support system hazard occurred during the summer of 2009 when a new type of urine processing assembly (a critical unit that allows water reclamation) broke down shortly after being brought on line in the ISS. The cause was eventually traced to precipitation of calcium from urine in a mechanism strikingly similar to the biological process of renal stone formation (Smith et al., 2012).

HEALTH RISKS 59 other associated predisposing factors (Ploutz-Snyder, 2013). No predic- tive or explanatory model exists for understanding this range in individu- al variation. Health Standards and Risk Profile The risk of fracture due to bone demineralization is classified as “Controlled” for missions to the ISS or to a near-Earth asteroid, and “Ac- ceptable” for lunar and Mars design reference missions (see Table 2-2). The risk of early-onset osteoporosis due to spaceflight is ranked “Ac- ceptable” for all four design reference missions (see Table 2-2). A NASA health standard for bone mineral loss has been developed that addresses both of these risks (see Box 2-1). Issues for Long Duration and Exploration Spaceflights An unknown variable is whether the rate of bone loss for astronauts during prolonged missions will remain consistent with rates observed in shorter (6 months or less) missions, whether it will plateau, or whether it may accelerate as space environment exposure lengthens. The long-term effects of bone loss are relatively well known for Earth-based popula- tions. Both menopause-related and senile osteoporosis are associated with an increased risk of fracture (primarily hip, vertebral, and wrist). While bone demineralization theoretically could increase the risk of ac- tual fracture during a long duration spaceflight, this has never occurred during a mission, although obviously, such an occurrence could jeopard- ize successful completion of the mission or lead to morbidity affecting astronaut well-being after the mission. Bone density is closely monitored in the active astronaut corps as well as in those astronauts who participate in the Longitudinal Study of Astronaut Health. During the ISS flights, astronauts exercise extensively, and their diet is tailored for maintaining bone density. After each flight, bone scans are done to examine the degree of bone loss, and protocols (both pharmaceutical and physical) are recommended for bone recovery. Postflight rehabilitation aims to return bone mass levels to preflight base- lines (NASA, 2007). Long duration and exploration spaceflights with landings, such as a mission to Mars, could expose astronauts to additional risks beyond those associated with a long stay on the ISS. Such an exploration-class mission, as currently envisioned, would include a period of exposure to

60 LONG DURATION AND EXPLORATION SPACEFLIGHT microgravity, the physical stress of landing, exposure to and activities in the fractional gravity of Mars (0.375 of the gravity of Earth; NASA, 2014d), physical stress during takeoff from Mars, exposure to micrograv- ity again, and finally physical stress during landing on Earth. Essentially nothing is known about how exposure to fractional gravity environment on Mars might influence the time course of microgravity-induced bone loss or fracture during the mission. Ethics issues raised in consideration of this health risk include poten- tial impact on crew selection, responsibilities for long-term monitoring and health of astronauts, balancing the unknowns regarding this risk with other risks and benefits, and ensuring informed decision making. RADIATION EXPOSURE For space missions in low Earth orbit (LEO), the major source of ra- diation exposure results from solar storms. For exploration-class mis- sions beyond LEO, it is the exposure to galactic cosmic radiation that is the significant health concern for both acute and long-term health conse- quences. High levels of radiation exposure (e.g., during solar storms) can lead to acute effects, including fatigue, nausea, and vomiting. Chronic exposure increases the risk of cancer, tissue degeneration, development of cataracts, and potential effects on the central nervous system, cardio- vascular system, immune function, and vision. Overview and Risk Identification NASA recognized the potential radiation risk from the beginning of its efforts to send astronauts into space. An extensive radiation research program has been under way and input has been received from a number of independent organizations (see Box 3-2). BOX 3-2 Overview of Selected Reports on Radiation Exposure Limits National Research Council (NRC, 1967), Radiobiological Factors in Manned Space Flight. The working group focused on identifying “immedi- ate or early performance decrement (early responses) occurring within a few hours to 1 month of major exposure;” “progressively increasing per- formance decrement or serious loss of performance over longer period of

HEALTH RISKS 61 flight as a result of an accumulating exposure (progressive injury to the blood-forming system);” and “probability of late radiation response” (p. 244). National Research Council (1970), Radiation Protection Guides and Con- straints for Space-Mission and Vehicle-Design Studies Involving Nuclear Systems. The NRC committee noted that the risk-benefit decisions depend on “a wide range of general and specific scientific and subjective judg- ments and should be the responsibility of those most informed about the aims and goals of the nation’s space program” (p. 3). The report noted the value of specific limits to spacecraft design and mission planning. The final recommendations of the NRC committee included a set of short-term guidelines to limit acute and degenerative effects, along with the concept of a “primary reference risk” to limit the exposure to that which “corre- sponds to an added probability of radiation-induced neoplasia over a peri- od of about 20 years that is equal to the natural probability for the specific population at risk” (p.16). They estimated this exposure to be 400 REM, which they believed corresponded to about a 2.3 percent risk of develop- ing cancer. However, they noted that acceptance of a higher risk for plane- tary missions than for space station missions “would seem both realistic and practical” (p. 16). National Council on Radiation Protection and Measurements (NCRP) Re- port 98, Guidance on Radiation Received in Space Activities (1989). In addition to recommending short-term limits to minimize acute effects, this study noted that cancer was the principal risk, and career limits were set to limit the risk of fatal cancer. The report limited the risk to 3 percent excess lifetime cancer mortality, based on comparison with other hazardous oc- cupations. The study was also the first to take age and sex into account. NCRP Report 132, Radiation Protection Guidance for Activities in Low- Earth Orbit (2000). NCRP 132 questioned the use of mortality data from hazardous occupations as the basis for space-related radiation limits, but endorsed the 3 percent lifetime risk of cancer death, as consistent with guidelines for terrestrial radiation workers. The report also briefly but ex- plicitly addressed the challenge of uncertainties in quantifying radiation limits: “It is well known that risk estimation is a difficult field in which there are many sources of potential error and therefore uncertainty…. Given the magnitude of these uncertainties and the problems of dose specification estimates or risk on which dose limits for astronauts are based should be recognized as very conservative and possibly subject to modified values when more precise information becomes available” (pp. 146-147). Report 132 clarified that the NCRP had only been tasked to assess risk to low Earth orbit (LEO) missions: “In NCRP Report No. 98 (NCRP, 1989) a sec- tion on mission scenarios with estimates of radiation exposure during mis- sions to the moon and Mars was included. Perhaps because of that inclusion, some have assumed that the guidance on dose limits in that re- port applied not only to missions in LEO but to all space missions. That was not the intention since the guidance provided was limited to expo- sures in LEO. In this Report, the guidance is also only intended to be ap-

62 LONG DURATION AND EXPLORATION SPACEFLIGHT plied for radiation exposures incurred during missions in LEO. Further NCRP reports will deal with other space situations” (pp. 2-3). NCRP Report 153, Information Needed to Make Radiation Protection Recommendations for Space Missions Beyond Low-Earth Orbit (2006). “The purpose of this Report is to identify and describe information needed to make radiation protection recommendations for space missions beyond low-Earth orbit (LEO). Current space radiation guidelines pertain only to missions in LEO and are not considered relevant for mission beyond LEO” (NCRP, 2006, p. 1). NRC (2012a), Technical Evaluation of the NASA Model for Cancer Risk to Astronauts Due to Space Radiation. The NRC committee did not address the underlying assumption of whether the permissible exposure limits were appropriate, rather they looked in detail at the constituent elements of the model used to calculate the risk. They found the model to be consistent with general radiation community approaches to quantify the radiation risk. SOURCES: NRC, 1967, 1970, 2012a; NCRP, 1989, 2000, 2006. Current career limits in the NASA standards are designed to keep the astronaut’s lifetime risk of exposure-induced death from cancer to no more than 3 percent (i.e., limiting the additional risk of cancer posed by mission-related radiation exposure). The risk and lifetime exposure limit can be explained as follows: If 100 astronauts were exposed to the upper bounds of the radiation limits, 3 would die of cancer attributable to that exposure. It would be anticipated that life expectancy for astronauts with radiation-induced cancer would be reduced by an average of 12 to 16 years compared to those without radiation-induced cancer (NASA, 2007). In 1990, NASA agreed to accept the recommendation of NCRP Re- port 98 to limit the lifetime risk to astronauts to 3 percent and formalized it in the 1995 NASA Health Standards (NASA, 1995). The 1995 stand- ards stated the permissible exposure limit in terms of sex- and age- dependent, dose-equivalent limits expressed in REMs,3 which were de- veloped to be consistent with the 3 percent risk of exposure-induced death (REID) (e.g., 237.5 REM for a 35-year-old male, 177.5 REM for a 35-year-old female) (NASA, 1995). These estimated limits did not in- clude a margin to represent the uncertainty in the calculations. As a result 3 REM (rad equivalent man) was used as a unit of dose equivalent in the 1995 NASA standard, but NASA now uses the sievert (Sv), the international standard for dose equiva- lent (NRC, 2012a). 1 REM is equal to 0.01 Sv.

HEALTH RISKS 63 of NCRP Report 132, published in 2000, NASA proposed to add a 95 percent confidence interval to the 3 percent limit. The current radiation permissible exposure limits with the 95 percent confidence interval was formally accepted in March 2007 (NASA, 2007) (see Box 2-1). The pro- cedure used to estimate the confidence interval is based on extrapolation from a wide range of physical and biological research, much of it con- ducted by NASA in a peer-reviewed research program open to the wider radiation research community. The model and the underlying methodol- ogy and assumptions are periodically reviewed by independent commit- tees, most recently by the NRC Space Studies Board (NRC, 2012a). Risk Management: Research, Exposure Limits, Monitoring, and Countermeasures NASA has undertaken a research program to quantify the estimates of the risk of radiation-related cancers and other acute and chronic ad- verse health effects (Cucinotta and Durante, 2009; Cucinotta et al., 2009; Huff and Cucinotta, 2009; Wu et al., 2009). Progress based on a peer- reviewed, science-based approach has significantly reduced the uncer- tainties associated with the estimates of the risk, and the long-term pro- gram goal is to reduce it further. Radiation levels are monitored in real time on the ISS by a suite of sensors distributed throughout the station. The reports from these sensors are used to estimate astronaut exposure throughout the mission. When the rate of exposure exceeds threshold levels, mission control is in- formed. For short periods, when radiation levels are significantly above normal, astronauts may be instructed to stay in better shielded areas of the ISS. In addition, each astronaut has a personal dosimeter that is read after the mission by the NASA Space Radiation Analysis Group to con- firm and record exposure. For exploration missions, research is under way to develop real-time personal dosimeters. NASA’s Space Radiation Analysis Group works with the National Oceanic and Atmospheric Administration’s Space Weather Prediction Center to monitor the radiation levels near Earth and monitor for evi- dence of solar storms. The center issues alerts when a radiation storm is under way, and the Space Radiation Analysis Group alerts ISS mission operations. Countermeasures for radiation exposure are limited. The primary countermeasure is shielding built into the spacecraft. However, due to the highly penetrating nature of galactic cosmic rays, shielding is only mar-

64 LONG DURATION AND EXPLORATION SPACEFLIGHT ginally effective at reasonable thicknesses; increasing the thickness of the shielding adds substantial mass with minimal additional reduction in exposure (NRC, 2008). Innovative approaches to reduce the risks also are being investigated, including pharmacological countermeasures. To date, these methods have not been shown to be effective. Thirty-day exposure limits have been established to limit the poten- tial for acute health outcomes due to radiation exposure, and operational processes and procedures are implemented to ensure that astronauts do not approach these limits. The greatest likely source of acute exposure is solar storms. Providing access to modest shielding and a system for time- ly warning of a pending or ongoing storm are critical elements of a risk management strategy to mitigate this risk (NRC, 2008). Pre-mission planning includes estimates of the radiation exposure to minimize the risk that the crew will exceed the career limits (with consideration of the crew member’s prior exposure). Radiation exposures also may include risks of chronic, degenerative effects that could affect health during a lengthy mission. Only recently has quantitative progress been made in understanding these risks, par- ticularly those affecting the central nervous and circulatory systems (Cucinotta et al., 2013b). An ongoing research effort is under way to im- prove this understanding. Health Standards and Risk Profile NASA has long recognized the threat of radiation, and has estab- lished space permissible exposure limits to protect its astronauts (see Box 2-1). The three tiers of radiation exposure limits in the NASA standards address: career limits, designed to limit the lifetime risk of exposure- induced death from cancer; short-term limits, designed to limit the risk of acute affects; and operational processes and procedures which emphasize that “in-flight radiation exposures shall be maintained using the ‘as low as reasonably achievable’ (ALARA) principle” (see Box 2-1; NASA, 2007, p. 20). Radiation limits are significant in planning for long duration and ex- ploration missions. Existing radiation standards, developed for the Space Shuttle and the ISS, would limit missions to durations of 150 to 250 days (Cucinotta et al., 2013a). The radiation standards are written to apply to all NASA human spaceflight missions and are not developed for any specific program (NASA, 2007). However, while some of the existing programs, such as the Space Shuttle and the ISS programs, can be con-

HEALTH RISKS 65 ducted within the standards, these standards impose potential limitations on long-duration missions (e.g., a 1-year stay on the ISS) or missions with architectures and objectives outside of LEO. The standards are based on recommendations to NASA from NCRP in a report that focused on LEO missions (NCRP, 2000). The NCRP report noted, “In this Re- port, the guidance is also only intended to be applied for radiation expo- sures incurred during missions in LEO” (NCRP, 2000, pp. 2-3). Individual Variation Issues A specific individual’s cancer risks vary by age, sex, and race, as well as genetic factors and life-style choices. National databases allow calculation by age, sex, and other factors, emphasizing the role of indi- vidual variability. NASA’s standards limit the additional risk of cancer posed by radiation exposure, not the total risk of dying from cancer. If one assumes a generic baseline lifetime risk of 23 percent for males and 19 percent for females (ACS, 2013), and leaves aside any differences in individual susceptibility, then male astronauts exposed to the 3 percent radiation exposure limit would have an estimated lifetime risk of death from cancer of 26 percent, while female astronauts would have an esti- mated risk of 22 percent. In general, for females the effective-dose radiation exposure is about three quarters that of males before reaching career limits (NASA, 2007). Females, on average, are more susceptible to radiation-induced cancer, in part because of the baseline higher risk of breast cancer in females; body size also is a factor as individuals with smaller builds have less body self-shielding. However, an individual’s risk of cancer is dominated by a number of factors (including genetic susceptibility, lifestyle choices, and prior exposures, both natural and medical), most of which may be highly individualized or unknown and, therefore, not currently possible to factor in calculations of cancer risk. Issues for Long Duration and Exploration Spaceflights As noted above, the number of days that astronauts could be exposed and stay within the current permissible exposure limits (safe days) will depend on assumptions about shielding and the radiation environment. The exposure varies with the 11-year solar cycle and with the number and intensity of solar storms (Cucinotta et al., 2013a). The existing radia- tion exposure limits would limit long duration or exploration missions to

66 LONG DURATION AND EXPLORATION SPACEFLIGHT 150 to 250 days (Cucinotta et al., 2013a). The estimated time in the zero- g interplanetary environment (“deep space”)4 for a Mars mission with current propulsion systems is estimated at 400 to 600 days (NASA, 2009). The design reference missions to Mars that are being considered include those that would include a long stay on the Martian surface (500 or more days) with a shorter stay in deep space, as well as those that would have a longer deep space duration and shorter Mars surface stays, nominally 30 to 90 days (NASA, 2009). Technology research into alter- native propulsion systems could significantly reduce the Mars transit time (NRC, 2012b). Other factors also could affect the number of safe days in space with regard to radiation exposure. For example, noting that astronauts, in gen- eral, are healthier than the average U.S. population, NASA recently cal- culated the number of safe days in space using the data on cancer risks for non-smoking, normal weight Americans (see Table 3-1). This change in the population data increases the estimates of safe days in space by 30 to 90 percent, depending on an astronaut’s age and sex (Cucinotta et al., 2013a). The NASA current precautionary decision to set exposure limits to protect astronauts with 95 percent probability (given the level of uncer- tainty) itself limits the number of safe days in space. If a less precaution- ary approach were taken by NASA using a lower confidence interval, the number of allowable days in space would be greater; alternatively, if greater precaution were exercised by using a higher confidence interval, fewer days would be permissible. More precise information about the dose of radiation-inducing fatal cancers could also change the number of permissible days in space that would not exceed the health standard. Ethics issues raised in the consideration of radiation exposure in- clude informed decision making by astronauts, individual variation and other factors that might impact crew selection, and assessing and balanc- ing the unknown risks of radiation exposure with other risks and benefits of the mission. While NASA has analyzed health risks of radiation based on age and sex (see Table 3-1), other factors influencing individual varia- tion and radiation risk must also be considered. Existing data on such variation should be factored into analyses, and new data should be col- lected to better predict individual risk from exposure to radiation. 4 Calculations regarding radiation exposure factor in the higher exposures to radiation in interplanetary space outside Earth’s magnetosphere and the relatively lower exposure levels on a planetary or lunar surface.

HEALTH RISKS 67 TABLE 3-1 Estimates of Safe Days in Deep Spacea Average solar maximum GCR and one significant solar storm (similar to that which occurred Average solar minimum GCR in August 1972) Age at NASA 2012 NASA 2012 NASA 2012 Exposure U.S. Average NASA 2012 U.S. Average Never- (years) Population Never-smokers Population smokers Males 35 209 (205) 271 (256) 306 (357) 395 (458) 45 232 (227) 308 (291) 344 (397) 456 (526) 55 274 (256) 351 (335) 367 (460) 500 (615) Females 35 106 (95) 187 (180) 144 (187) 276 (325) 45 139 (125) 227 (212) 187 (232) 319 (394) 55 161 (159) 277 (246) 227 (282) 383 (472) NOTE: Solar minimum is a 2- to 3-year period of low solar activity in the 10- to 11-year solar cycle. Solar maximum is a corresponding 5- to 7-year period of enhanced solar activity. Galactic cosmic radiation is at a peak during solar minimum and somewhat reduced during solar maximum. Values in parentheses for solar minimum are for the deep solar minimum of 2009. Values in paren- theses for solar maximum are for the case where a storm shelter is available to reduce the solar storm exposure to a negligible amount. GCR = galactic cosmic radiation; REID = risk of exposure-induced death. a Safe days in deep space (zero-g interplanetary environment) is a concept de- fined as the maximum number of days with 95 percent confidence interval to be below the NASA 3 percent REID limit, assuming nominal shielding of 20 g/cm2 aluminum. SOURCE: Adapted from Cucinotta et al., 2013a. DISCUSSION: Because the highly penetrating GCR dominates the effective dose, and because GCR is at a peak at solar minimum, the safe days are lowest at solar minimum. The deep solar minimum of 2009 had a significantly higher GCR than previously experienced, thus further reducing the safe days. There are more safe days near solar maximum because the GCR is reduced relative to solar minimum, and because solar storm radiation is significantly attenuated by nominal shielding. If a storm shelter (a smaller volume within the vehicle with additional shielding) is available to the astronauts, the exposure from a solar storm can be reduced to a negligible amount, so the total exposure would be from the GCR alone. This could further increase the number of safe days at solar maximum, assuming adequate solar storm warning.

68 LONG DURATION AND EXPLORATION SPACEFLIGHT SUMMARY The examples provided in this chapter illustrate the range of issues that are faced as NASA makes decisions about health standards for long duration and exploration spaceflights. The potential short- and long-term health impacts encompass many systems of the human body as well as behavior and performance issues. Decisions about health standards are complicated by the depth of uncertainty regarding what will happen with extended stays in space, high exposures to galactic cosmic radiation, and other risks and challenges. Additionally, data are minimal or non-existent for variations in individual susceptibility based on factors such as eth- nicity, sex, age, etc. A major challenge is the lack of information about the interaction of risks and the extent to which these interactions alter the overall level of risk. The following chapter outlines the ethics principles and responsibilities that can be brought to bear on decisions regarding health risks. REFERENCES ACS (American Cancer Society). 2013. Lifetime risk of developing or dying from cancer. http://www.cancer.org/cancer/cancerbasics/lifetime-probability-of- developing-or-dying-from-cancer (accessed January 15, 2014). Aldrin, B. 1973. Return to Earth. New York: Random House. Alexander, D. J., C. R. Gibson, D. R. Hamilton, S. M. C. Lee, T. H. Mader, C. Otto, C. M. Oubre, A. F. Pass, S. H. Platts, J. M. Scott, S. M. Smith, M. B. Stenger, C. M. Westby, and S. B. Zanello. 2012. Evidence report: Risk of spaceflight-induced intracranial hypertension and vision alterations. http://humanresearchroadmap.nasa.gov/Evidence/reports/VIIP.pdf (accessed November 8, 2013). Basner, M., D. F. Dinges, D. Mollicone, A. Ecker, C. W. Jones, E. C. Hyder, A. D. Antonio, I. Savelev, K. Kan, N. Goel, B. V. Morukov, and J. P. Sutton. 2013. Mars 320-d mission simulation reveals protracted crew hypokinesis and alterations of sleep duration and timing. Proceedings of the National Academy of Sciences of the United States of America 110(7):2635-2640.

HEALTH RISKS 69 Behnken, R., M. Barratt, S. Walker, and P. Whitson. 2013. Presentation to the Institute of Medicine, Ethics Principles and Guidelines for Health Standards for Long Duration and Exploration Spaceflights: Astronaut Office. PowerPoint presented at the second meeting of the Committee on Ethics Principles and Guidelines for Health Standards for Long Duration and Exploration Spaceflights. Washington, DC, July 25. http://www.iom.edu /~/media/Files/Activity%20Files/Research/HealthStandardsSpaceflight/201 3-JUL-25/Panel%202%20Astronaut%20Corp%20Final%20IOM_presentation final2.pdf (accessed November 8, 2013). Billica, R. 2000. Inflight medical events for U.S. astronauts during space shuttle programs STS-1 through STS-89, April 1981–January 1998. Presentation to the Institute of Medicine Committee on Creating a Vision for Space Medicine During Travel Beyond Earth Orbit. NASA Johnson Space Center, Houston, February 22. As cited in: IOM, 2001. Safe passage: Astronaut care for exploration missions. Washington, DC: National Academy Press. Burrough, B. 1998. Dragonfly: NASA and the crisis aboard Mir. New York: HarperCollins. Cucinotta, F. A., and M. Durante. 2009. Evidence report: Risk of radiation carcinogenesis. http://humanresearchroadmap.nasa.gov/Evidence/reports/ Carcinogenesis.pdf (accessed January 8, 2014). Cucinotta, F. A., H. Wang, and J. L. Huff. 2009. Evidence report: Risk of acute or late central nervous system effects from radiation exposure. http://humanresearchroadmap.nasa.gov/Evidence/reports/CNS.pdf (access- ed January 8, 2014). Cucinotta, F. A., M.-H. Y. Kim, and L. J. Chappell. 2013a. Space radiation cancer risk projections and uncertainties—2012. http://ston.jsc.nasa. gov/collections/TRS/_techrep/TP-2013-217375.pdf (accessed November 11, 2013). Cucinotta, F. A., M.-H. Y. Kim, L. J. Chappell, and J. L. Huff. 2013b. How safe is safe enough? Radiation risk for a human mission to Mars. PLoS ONE 8(10):e74988. Czeisler, C. A. 2011. Impact of sleepiness and sleep deficiency on public health—utility of biomarkers. Journal of Clinical Sleep Medicine 7(5):s6-s8. Czeisler, C. A., J. S. Allan, S. H. Strogatz, J. M. Ronda, R. Sanchez, C. D. Rios, W. O. Freitag, G. S. Richardson, and R. E. Kronauer. 1986. Bright light resets the human circadian pacemaker independent of the timing of the sleep-wake cycle. Science 233(4764):667-671. Dinges, D. F., F. Pack, K. Williams, K. A. Gillen, J. W. Powell, G. E. Ott, C. Aptowicz, and A. I. Pack. 1997. Cumulative sleepiness, mood disturbance, and psychomotor vigilance performance decrements during a week of sleep restricted to 4-5 hours per night. Sleep 20(4):267-277. ESA (European Space Agency). 2011. Mars500: Study overview. http://www. http://www.esa.int/Our_Activities/Human_Spaceflight/Mars500/Mars500_ study_overview (accessed January 8, 2014).

70 LONG DURATION AND EXPLORATION SPACEFLIGHT Flynn, C. F. 2005. An operational approach to long-duration mission behavioral health and performance factors. Aviation, Space, and Environmental Medicine 233(Suppl. 6):B42-B51. Garber, S., and R. Launius. 2005. A brief history of NASA. http://history.nasa. gov/factsheet.htm (accessed December 4, 2013). Goel, N., and D. F. Dinges. 2012. Predicting risk in space: Genetic markers for differential vulnerability to sleep restriction. Acta Astronautica 77:207-213. Grant I., H. R. Eriksen, P. Marquis, I. J. Orre, L. Palinkas, P. Suedfeld, E. Svensen, and H. Ursin. 2007. Psychological selection of Antarctic personnel: The “SOAP” instrument. Aviation, Space, and Environmental Medicine 78:793- 800. Gunderson, E. K. E. (Ed.). 1974. Psychological studies in Antarctica. In Human adaptability to Antarctic conditions. Washington, DC: American Geophysical Union. Pp. 115-131. Gushin, V. I., T. B. Kolinitchenko, V. A. Efimov, and C. Davies. 1996. Chapter 16: Psychological evaluation and support during exemsi. In Advances in Space Biology and Medicine. Vol. 5, edited by S. L. Bonting. Greenwich, CT: JAI Press. Pp. 283-295. Huff, J. L., and F. A. Cucinotta. 2009. Evidence report: Risk of degenerative tissue or other health effects from radiation exposure. http://humanresearch roadmap.nasa.gov/Evidence/reports/Degen.pdf (accessed January 8, 2014). IOM (Institute of Medicine). 2001. Safe passage: Astronaut care for exploration missions. Washington, DC: National Academy Press. IOM. 2014. Review of NASA’s evidence reports on human health risks: 2013 letter report. Washington, DC: The National Academies Press. Kanas, N., and D. Manzey. 2008. Space psychology and psychiatry. 2nd ed. El Segundo, CA: Microcosm Press. Kanas, N. A., V. P. Salnitskiy, J. E. Boyd, V. I. Gushin, D. S. Weis, S. A. Saylor, O. P. Kozerenko, and C. R. Marmar. 2007. Crewmember and mission control personnel interactions during International Space Station missions. Aviation, Space, and Environmental Medicine 78(6):601-607. Kanas, N., G. Sandal, J. E. Boyd, V. I. Gushin, D. Manzey, R. North, G. R. Leon, P. Suedfeld, S. Bishop, E. R. Fiedler, N. Inoue, B. Johannes, D. J. Kealey, N. Kraft, I. Matsuzaki, D. Musson, L. A. Palinkas, V. P. Salnitskiy, W. Sipes, J. Stuster, and J. Wang. 2009. Psychology and culture during long-duration space missions. Acta Astronautica 64(7-8):659-677. Kuna, S. T., G. Maislin, F. M. Pack, B. Staley, R. Hachadoorian, E. F. Coccaro, and A. I. Pack. 2012. Heritability of performance deficit accumulation during acute sleep deprivation in twins. Sleep 35(9):1223-1233. Landholt, H.-P. 2008. Genotype-dependent differences in sleep, vigilance, and response to stimulants. Current Pharmaceutical Design 14(32):3396-3407. Lebedev, V. 1988. Diary of a cosmonaut: 211 days in space. College Station, TX: Phytoresource Research, Inc.

HEALTH RISKS 71 Linenger, J. M. 2000. Off the planet: Surviving five perilous months aboard the space station Mir. New York: McGraw-Hill. Mader, T. H., C. R. Gibson, A. F. Pass, L. A. Kramer, A. G. Lee, J. Fogarty, W. J. Tarver, J. P. Dervay, D. R. Hamilton, A. Sargsyan, J. L. Phillips, D. Tran, W. Lipsky, J. Choi, C. Stern, R. Kuyumjian, and J. D. Polk. 2011. Optic disk edema, globe flattening, choroidal folds, and hyperopic shifts observed in astronauts after long-duration space flight. Ophthalmology 18(10):2058-2069. Mader, T. H., C. R. Gibson, A. F. Pass, A. G. Lee, H. E. Killer, H.-C. Hansen, J. P. Dervay, M. R. Barratt, W. J. Tarver, A. E. Sargsyan, L. A. Kramer, R. Riascos, D. G. Bedi, and D. R. Pettit. 2013. Optic disk edema in an astronaut after repeat long-duration space flight. Journal of Neuro- Ophthalmology 33:249-255. McFadden, T. J., R. L. Helmreich, R. M. Rose, and L. F. Fogg. 1994. Predicting astronauts’ effectiveness: A multivariate approach. Aviation, Space, and Environmental Medicine 65(10):904-909. Molloy, M. W., and R. A. Petrone. 2013. Apollo 12 command and service module (CSM). http://nssdc.gsfc.nasa.gov/nmc/spacecraftDisplay.do?id=1969-099A (accessed December 4, 2013). NASA (National Aeronautics and Space Administration). 1995. Man-systems integration standards. Revision B. NASA-STD-3000. https://standards. nasa.gov/documents/viewdoc/3314902/3314902 (accessed January 8, 2014). NASA. 2007. NASA space flight human system standard. Volume 1: Crew health. NASA-STD-3001. https://standards.nasa.gov/documents/detail/3315622 (acc- essed November 8, 2013). NASA. 2009. Human exploration of Mars: Design reference architecture 5.0. NASA/SP-2009-566. http://www.nasa.gov/pdf/373665main_NASA-SP-2009- 566.pdf (accessed January 8, 2014). NASA. 2013. Human research program requirements document, Revision F. HRP-47052. Johnson Space Center. http://www.nasa.gov/pdf/579466main_ Human_Research_Program_Requirements_DocumentRevF.pdf (accessed October 18, 2013). NASA. 2014a. Kennedy Space Center visitor complex: Astronaut memorial. http://www.kennedyspacecenter.com/the-experience/astronaut-memorial.aspx (accessed January 13, 2014). NASA. 2014b. Mir Space Station. http://history.nasa.gov/SP-4225/mir/mir.htm (accessed January 13, 2014). NASA. 2014c. Human Research Roadmap: Gaps. http://humanresearchroad map.nasa.gov/Gaps (accessed February 7, 2014). NASA. 2014d. All about Mars. http://mars.nasa.gov/allaboutmars/extreme/ quickfacts (accessed February 19, 2014). NCRP (National Council on Radiation Protection and Measurements). 1989. NCRP report no. 98: Guidance on radiation received in space activities. Bethesda, MD: NCRP.

72 LONG DURATION AND EXPLORATION SPACEFLIGHT NCRP. 2000. NCRP report no. 132: Radiation protection guidance for activities in low-Earth orbit. Bethesda, MD: NCRP. NCRP. 2006. NCRP report no. 153: Information needed to make radiation protection recommendations for space missions beyond low-Earth orbit. Bethesda, MD: NCRP. NRC (National Research Council). 1967. Radiobiological factors in manned space flight. Washington, DC: National Academy Press. NRC. 1970. Radiation protection guides and constraints for space-mission and vehicle-design studies involving nuclear systems. Washington, DC: National Academy Press. NRC. 1998. A strategy for research in space biology and medicine into the next century. Washington, DC: National Academy Press. NRC. 2008. Managing space radiation risk in the new era of space exploration. Washington, DC: The National Academies Press. NRC. 2011. Recapturing a future for space exploration: Life and physical sciences research for a new era. Washington, DC: The National Academies Press. NRC. 2012a. Technical evaluation of the NASA model for cancer risk to astronauts due to space radiation. Washington, DC: The National Academies Press. NRC. 2012b. NASA space technology roadmaps and priorities: Restoring NASA’s technological edge and paving the way for a new era in space. Washington, DC: The National Academies Press. Palinkas, L. A., and P. Suedfeld. 2008. Psychological effects of polar expeditions. Lancet 371(9607):153-163. Palinkas, L. A., F. Glogower, M. Dembert, K. Hansen, and R. Smullen. 2004. Incidence of psychiatric disorders after extended residence in Antarctica. International Journal of Circumpolar Health 63(2):157-168. Ploutz-Snyder, L. 2013. Musculoskeletal working group report. PowerPoint presented at NASA and NSBRI Present a Virtual Workshop: The Impact of Sex & Gender on Adaptation to Space. http://www.nasa.gov/sites/ default/files/files/Ploutz-Snyder_Sex-Gender_508.pdf (accessed October 22, 2013). Rose, R. M., L. F. Fogg, R. L. Helmreich, and T. J. McFadden. 1994. Psychological predictors of astronaut effectiveness. Aviation, Space, and Environmental Medicine 65(10):910-915. Rosnet, E., C. Le Scanff, and M. Sagal. 2000. How self-image and personality affect performance in an isolated environment. Environment and Behavior 32(1):18-31. Sandal, G. M. 2001. Crew tension during a space station simulation. Environment and Behavior 33(1):134-150. Sandal, G. M., R. J. Vaernes, T. Bergan, M. Warncke, and H. Ursin. 1996. Psychological reactions during polar expeditions and isolation in hyperbaric chambers. Aviation, Space, and Environmental Medicine 67(3):227-234.

HEALTH RISKS 73 Sandal, G. M., I. M. Endresen, R. Vaernes, and H. Ursin. 1999. Personality and coping strategies during submarine missions. Military Psychology 11(4):381-404. Schmidt, L. L., K. Keeton, K. J. Slack, L. B. Leveton, and C. Shea. 2009. Evidence report: Risk of performance errors due to poor Team Gap cohesion and performance, inadequate selection/Team Gap composition, inadequate training, and poor psychosocial adaptation. http://humanresearchroadmap. nasa.gov/Evidence/reports/Team Gap.pdf (accessed January 2, 2014). Shepanek, M. 2005. Human behavioral research in space: Quandaries for research subjects and researchers. Aviation, Space, and Environmental Medicine 76(Suppl. 6):B25-B30. Sibonga, J. D., J. A. Jones, J. G. Myers, B. E. Lewandowski, T. F. Lang, and J. H. Keyak. 2008a. Evidence report: Risk of bone fracture. http://humanresearch roadmap.nasa.gov/evidence/reports/Bone%20Fracture.pdf (accessed October 22, 2013). Sibonga, J. D., L. C. Shackelford, A. LeBlanc, S. Petak, S. M. Smith, S. A. Bloomfield, J. H. Keyak, T. F. Lang, S. B. Arnaud, S. Amin, B. L. Clarke, R. Wermers, P. R. Cavanagh, A. A. Licata, S. Judex, M. B. Schaffler, and T. A. Bateman. 2008b. Evidence report: Risk of accelerated osteoporosis. http://humanresearchroadmap.nasa.gov/evidence/reports/osteo.pdf (access- ed October 22, 2013). Slack, K. J., C. Shea, L. B. Leveton, A. M. Whitmire, and L. L. Schmidt. 2009. Evidence report: Risk of behavioral and psychiatric conditions. http://humanresearchroadmap.nasa.gov/evidence/reports/BMED.pdf (access- ed Ocotber 18, 2013). Smith, S. M., M. A. Heer, L. C. Shackelford, J. D. Sibonga, L. Ploutz-Snyder, and S. R. Zwart. 2012. Benefits for bone from resistance exercise and nutrition in long-duration spaceflight: Evidence from biochemistry and densitometry. Journal of Bone and Mineral Research 27(9):1896-1906. Stuster, J. 2010. Behavioral issues associated with long-duration space expeditions: Review and analysis of astronaut journals. http://ston.jsc.nasa.gov/collections/trs/_techrep/TM-2010-216130.pdf (access- ed November 8, 2013). Tansey, W. A., J. M. Wilson, and K. E. Schaefer. 1979. Analysis of health data from 10 years of Polaris submarine patrols. Undersea Biomedical Research 6(Suppl.):s217-s246. Thomas, T. L., T. I. Hooper, M. Camarca, J. Murray, D. Sack, D. Molé, R. T. Spiro, W. G. Horn, and F. C. Garland. 2000. A method for monitoring the health of US Navy submarine crewmembers during periods of isolation. Aviation, Space, and Environmental Medicine 71(7):699-705. Van Dongen, H. P. A., M. D. Baynard, G. Maislin, and D. F. Dinges. 2004. Systematic interindividual differences in neurobehavioral impairment from sleep loss: Evidence of trait-like differential vulnerability. Sleep 27(3):423-433.

74 LONG DURATION AND EXPLORATION SPACEFLIGHT Watkins, S. D., and Y. R. Barr. 2010. Papilledema summit: Summary report. TM-2010-216114. http://ston.jsc.nasa.gov/collections/TRS/_techrep/TM- 2010-216114.pdf (accessed October 22, 2013). Whitmire, A. M., L. B. Leveton, L. Barger, G. Brainard, D. F. Dinges, E. Klerman, and C. Shea. 2009. Evidence report: Risk of performance errors due to sleep loss, circadian desynchronization, fatigue, and work overload. http://humanresearchroadmap.nasa.gov/Evidence/reports/Sleep.pdf (accessed January 2, 2014). Williams, D.R. 2011. The Apollo 1 tragedy. http://nssdc.gsfc.nasa.gov/ planetary/lunar/apollo1info.html (accessed January 8, 2014). Wu, H., J. L. Huff, R. Casey, M.-H. Kim, and F. A. Cucinotta. 2009. Risk of acute radiation syndromes due to solar particle events. http://humanresearch roadmap.nasa.gov/Evidence/reports/ARS.pdf (accessed January 8, 2014).

Next: 4 Risk Acceptance and Responsibilities in Human Spaceflight and Terrestrial Activities »
Health Standards for Long Duration and Exploration Spaceflight: Ethics Principles, Responsibilities, and Decision Framework Get This Book
×
Buy Paperback | $45.00 Buy Ebook | $36.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

Since its inception, the U.S. human spaceflight program has grown from launching a single man into orbit to an ongoing space presence involving numerous crewmembers. As the U.S. space program evolves, propelled in part by increasing international and commercial collaborations, long duration or exploration spaceflights - such as extended stays on the International Space Station or missions to Mars - become more realistic. These types of missions will likely expose crews to levels of known risk that are beyond those allowed by current health standards, as well as to a range of risks that are poorly characterized, uncertain, and perhaps unforeseeable. As the National Aeronautics and Space Administration (NASA) and Congress discuss the next generation of NASA's missions and the U.S. role in international space efforts, it is important to understand the ethical factors that drive decision making about health standards and mission design for NASA activities.

NASA asked the Institute of Medicine to outline the ethics principles and practices that should guide the agency's decision making for future long duration or exploration missions that fail to meet existing health standards. Health Standards for Long Duration and Exploration Spaceflight identifies an ethics framework, which builds on the work of NASA and others, and presents a set of recommendations for ethically assessing and responding to the challenges associated with health standards for long duration and exploration spaceflight.As technologies improve and longer and more distant spaceflight becomes feasible, NASA and its international and commercial partners will continue to face complex decisions about risk acceptability. This report provides a roadmap for ethically assessing and responding to the challenges associated with NASA's health standards for long duration and exploration missions. Establishing and maintaining a firmly grounded ethics framework for this inherently risky activity is essential to guide NASA's decisions today and to create a strong foundation for decisions about future challenges and opportunities.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!