You are on page 1of 70

41.

DELIR IUM
N EU ROBIOL O G Y, CH A R AC T E R I S T IC S , A N D M A N AG E M E N T

José R. Maldonado

I N T RODUC T IO N to be demented.” Delirium is derived from the Latin roots


de (meaning “away from”), lira (meaning “furrow in a field”),
Epidemiologically, delirium is likely the most common psy- and ium (Latin for singular), and literally means “a going
chiatric syndrome found in the general hospital setting. Its off the plowed track; a madness.” Celsus described the vari-
occurrence parallels the severity of the underlying medical ous forms of delirium: “Now there are several kinds of it: for
processes. Clinically, delirium presents as an acute or sub- among phrenitic people some are merry, others sad; other
acute organic mental syndrome characterized by disturbance grow outrageous, and do acts of violence. . . . Those, who are
of consciousness, global cognitive impairment, disorienta- more violent in their actions, it is proper to bind, lest they
tion, perceptual disturbance, deficits in attention, changes should hurt either themselves or any other person.” Despite
in psychomotor activity, disordered sleep–wake cycles, and Celsus’s introduction of the term delirium, the syndrome
fluctuations in symptom severity, all representing a signifi- continued to be referred to as “phrenitis.”
cant change over the patient’s baseline level of cognitive and In the second century a.d., Aretaeus classified diseases
behavioral functioning. Etiologically, delirium is a neurobe- as “acute” or “chronic.” Among the known mental disorders
havioral syndrome caused by the transient disruption of of the time, phrenitis and lethargy (lethargus) were the chief
normal neuronal activity secondary to disturbances of sys- acute diseases, and both were believed to be caused by fever
temic physiology. The phenotypes of delirium (i.e., hyperac- or poisons and thus reflecting brain disease (Lipowski, 1991).
tive, hypoactive, mixed) are likely a function of which of the The ancient description of phrenitis typically involved rest-
various neurotransmitter systems predominates, which can lessness, insomnia, and hallucinations, and appears consis-
change due to progression of illness or the influence of envi- tent with our current understanding of hyperactive delirium.
ronmental and/or pharmacological agents. Conversely, lethargy, involving undue quietness and sleepi-
Hippocrates provided a clear description of the syndrome, ness, is more akin to what we now call hypoactive delirium.
but called it phrenitis, meaning an “acute inflammation of It was not until the sixteenth century that we find the
mind and body” (Lipowski, 1991). In fact, phrenitis literally first description of delirium in the English medical litera-
means “inflammation of the mind or diaphragm” (Greek ture, in the textbook The Method of Physick (Barrough, 1583).
phren = mind, diaphragm; itis = inflammation) (Sakai, 1991; Barrough called it “frenesie” and described it involved the
Frederiks, 2000). It was an ancient disease concept or descrip- derangement of three main functions:  imagination, cogita-
tive notion that denoted a disease consisting of a combination tion, and memory; also featuring disturbed sleep.
of a mental disorder and fever (the term phrenes had a double Thomas Willis (1621–1675) described delirium as result-
meaning, indicating both a mental and a somatic domain), ing from “fever, poisoning, hemorrhage, lack of sleep, and
distinguishing phrenitis from conventional madness (Sakai, drunkenness” (Willis, 1672). According to Willis, its core
1991; Frederiks, 2000). features included “incongruous conceptions and confused
The term delirium is reported to have been first used by the thoughts,” distorted visual perceptions, and disturbed behav-
lay Roman writer Celsus (a.d. 1) (Adamis et al., 2007) and ior. Willis referred to phrenitis as “an inflammation of the
described in his compendium De Medicina as “a madness whole sensitive soul and animal spirits” (Bynum, 2000).
which is both acute and happens in a fever; the Greeks call it It was not until the nineteenth century that the term phre-
phrenitis” (Celsus, 1814, Chapter XVIII, p. 115). According nitis was replaced by the word “delirium,” in part due to the
to Celsus, a fever becomes a phrenitis “when the delirium efforts of Gerard van Swieten (1700–1772), Giovanni Battista
begins to continue without interruption; or when the patient, Morgagni (1682–1771), and Philippe Pinel (1745–1826)
though he still have his reason, yet forms to himself some vain (Byl  & Szafran, 1996; Frederiks, 2000). After the works of
images:  it is perfect, when the mind gives itself up to these these luminaries of medicine, the word phrenitis fell into
images” (Celsus, 1814). It was also Celsus who was the first disuse, replaced by the term febrile delirium (Berrios, 1981).
to declare that “not all deliria were reversible, but in some In 1813, the British physician Thomas Sutton introduced
cases, although the causes disappeared, patients continued the term delirium tremens to designate delirium caused by

823
the withdrawal from “spirits, and in consequence of excess 2008b; NICE, 2010). In fact, delirium has been reported to
or habitual indulgence in them” (Sutton, 1813). In modern be one of the six most common preventable conditions in
times, the term is exclusively applied to delirium resulting hospitalized elderly patients (Rothschild & Leape, 2000).
from alcohol withdrawal (2001). In the 1817 German treatise Some specific medical conditions and surgical procedures
“The Dream and the Feverish Insanity” (Der Traum und des are associated with a particularly high incidence of delir-
fieberhafte Irreseyn), Greiner first pioneered the term “clouding ium during the postoperative period and among critical
of consciousness” (Verdunkelung des Bewusstseins) to describe care unit patients (see Table 41.1). Terminally ill cancer
the pathophysiological abnormality responsible for delirium, patients, patients with moderate to severe traumatic brain
and linking disordered consciousness to the phenomenon of injury, frail elders, and the critically ill are clinical popu-
delirium (Greiner, 1817). This relationship was further devel- lations generally recognized to have a high incidence of
oped by Jackson, who viewed delirium as a clinical manifesta- delirium over their course of illness, due to factors such as
tion of disordered consciousness (Jackson, 1932). To Jackson, polypharmacy, comorbid medical conditions, and meta-
delirium represented a state of reduced consciousness that bolic dysfunction.
ranged from the “slightest confusion of thought to coma,” and A recent study at a large general hospital found that
was due to some degree of “dissolution” of the topmost layer delirium had a point prevalence of one in every five inpa-
of the nervous centers and consequent release from inhibition tients, with some variation in certain medico-surgical units
of the evolutionarily lower centers. and populations (Ryan, et al. 2013). The frequency of delir-
Some of the most significant contributions in the field of ium varies from 15% to 60% in hospitalized general medi-
delirium include the works of: cal and surgical patients (Smith & Dimsdale, 1989; Francis,
Martin, & Kapoor, 1990; Levkoff et al., 1992; Schor et al.,
Engel and Romano (1959), who demonstrated that the 1992; Williams-Russo, Urquhart, Sharrock, & Carlson,
behavioral and cognitive manifestations of delirium were 1992; Pompei et al. 1994; Parikh & Chung, 1995; van der
a reflection of an underlying metabolic derangement, Mast & Roest, 1996; Elie, Cole, Primeau, & Bellavance,
and demonstrated the relationship between reduction of 1998; Bucht, Gustafson, & Sandburg, 1999; Fann Alfano,
brain metabolic rate and electroencephalogram (EEG) Roth-Roemer, Katon, & Syrjala, 2007; Aldemir Ozen,
slowing; Kara, Sir, & Bac, 2001; Lepouse, Lautner, Liu, Gomis, &
Leon, 2006; Siddiqi, House, & Holmes, 2006; Hala, 2007;
Blass et al. (1981), who first proposed that impaired cere- Lundstrom et al., 2007; Maldonado, 2008b; Katznelson
bral oxidative metabolism resulted in reduced synthesis et al., 2009; Maldonado et al., 2009; Tognoni et al., 2011;
of various neurotransmitters, particularly acetylcholine, Ryan et al., 2013). The rate of postoperative delirium has
leading to some of the characteristic deficits; and been reported to range from 10% to 74% (Dyer, Ashton, &
Lipowski (1967, 1980, 1987), considered by many the Teasdale, 1995; Vaurio, Sands, Wang, Mullen, & Leung,
contemporary father of the study of delirium, who, in his 2006; Bruce, Ritchie, Blizard, Lai, & Raven, 2007;
early seminal work, summarized the history of the field, Maldonado et al., 2009; Wiesel, Klausner, Soffer & Zold,
highlighted the problem with nomenclature, character- 2011). Studies have demonstrated that up to 87% of criti-
ized its clinical features, summarized known etiological cally ill patients develop delirium during their ICU stay (Ely,
factors, suggested diagnostic criteria and management, Margolin, et al., 2001). Delirium is common among cancer
and influenced the course of delirium research for years. patients (Adams, 1988; Weinrich & Sarna, 1994; Morita,
Tei, Tsunoda, Inoue, & Chihara, 2001; Breitbart, Gibson, &
Unfortunately, one of the factors contributing to confu- Tremblay, 2002; Centeno, Sanz, & Bruera, 2004; Gaudreau,
sion about delirium is that there are many other terms used Gagnon, Harel, Tremblay, & Roy, 2005, Gaudreau, Gagnon,
by various medical disciplines to describe the same phenome- Roy, Harel & Tremblay, 2007).
non: for example, “ICU (intensive care unit) psychosis,” “sun- Among the elderly, studies have found that between
downing,” “acute confusional state,” “toxic confusional state,” 14% and 24% of elderly patients are admitted to the hos-
“post-anesthetic excitement,” “acute postoperative psychosis,” pital with delirium (Inouye, 1998). It is estimated that
“post-cardiotomy delirium,” and “toxic-metabolic encepha- new cases arise in 6%–46% of hospitalized elderly patients
lopathy.” The author personally prefers the term acute brain (Inouye, 1993). In the emergency department, the preva-
failure, as this term is compatible with multiple etiologies and lence in elderly patients is 9.6% (Elie et al., 2000). This data
settings, describes the acuteness of the onset, implies that the is consistent with more recent studies demonstrating that
most important organ in the body, the brain, is failing, and among those aged 85+ years, the prevalence of delirium is
this may be associated with significant cognitive decline, mor- about 10%, rising up to 22% in populations with higher
bidity, and mortality. percentages of demented elders and as high as 70% for those
in long-term care facilities (de Lange, Verhaak, & van der
Meer, 2013).
E PI DE M IOL O G Y OF   DE L I R I U M The occurrence of delirium seems to be highest among
the very sick. Various studies have repeatedly demonstrated
Delirium’s prevalence surpasses that of all other psychiat- that up to 87% of those critically ill develop delirium during
ric syndromes in the general medical setting (Maldonado, their stay in critical care units (Ely, Margolin, et al., 2001).

824 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Table 41.1 A COMPAR ISON OF THE INCIDENCE OF PSYCHIATR IC DISOR DER
IN THE GENER AL POPULATION AND DELIR IUM AMONG MEDICALLY ILL PATIENTS

MEDICAL POPULATIONS
UNDER STUDY PR EVALENCE OF DELIR IUM (%)

General Medicine Wards 9–24% (Ritchie et al., 1996; Valdes et al., 2000; Maldonado, Dhami, & Wise,
2003; Gonzalez, de Pablo, et al., 2004; Speed et al., 2007)

HIV/AIDS 30%–40% (Fernandez, Levy, & Mansell, 1989; Uldall et al., 2000)

Medical-ICU 60%–80% (Ely et al., 2001)

Post-Stroke 13%–48% (McManus et al., 2007)

Surgery

Postoperative Delirium 10%–74% (Vaurio et al., 2006; Wiesel, Klausner, et al., 2011)

General Surgical Wards 7%–52% (Dyer & Teasdale, 1995; Bucht, Gustafson, & Sandburgh, 1999)

Spine Surgery 12.5% (Kawaguchi et al., 2006)

Post-CABG 25%–32% (Nevin, Colchester, et al., 1989; Ebert & Herrmann 2001)

Post-Cardiotomy 50%–67% (Smith & Dimsdale, 1989; van der Mast & Roest, 1996; Maldonado
et al., 2009)

Abdominal Aneurysm Repair 33% (Benoit et al., 2005)

Geriatrics

Outpatient Minor (Cataract) 4.4% (Milstein et al., 2002)


Surgery

At Time of Hospitalization 10%–15%

In Nursing Homes 15%–60%

Frail Elderly Patient 60% (Francis et al., 1990)

Elective Hip or Knee Replacement 25% (Gustafson, Berggren, et al., 1988; Marcantonio, Flacker, et al., 2000)

Bilateral Knee Replacement 41% (Williams-Russo et al., 1992)

Femoral Neck Fracture Repair 65% (Gustafson, Bucht, et al., 1988; Marcantonio et al., 2000)

Oncology

General Prevalence 25%–40% (Weinrich & Sarna, 1994; Olofsson et al., 1996; Tuma & DeAngelis,
2000)

Palliative Care Units 26%–44% (Lawlor, Pereira, et al., 2000; Morita, Tei, et al., 2001; Lawlor, 2002;
Centeno, Sanz, & Bruera, 2004)

Bone Marrow Transplantation 73% (Fann et al., 2007)

Advanced Cancer Up to 85% (Lawlor et al., 2000)

Psychiatry
Among Psychiatric Patients 14.6% (Ritchie et al., 1996)

In addition, delirium is common among cancer patients et al., 2000). Similar to patients in critical care units, termi-
(Adams, 1988; Weinrich & Sarna, 1994; Morita et al., 2001; nal delirium occurs in up to 88% of patients dying of cancer
Breitbart et al., 2002; Centeno, Sanz, & Bruera, 2004; (Lawlor et al., 2000).
Gaudreau et al., 2005b). Among those with advanced can- The large variation of incidence and prevalence data
cer, delirium was diagnosed in 42% of patients on admission, reported by the various studies reflects differences in patient
while it developed in 45% of hospitalized cancer patients populations studied, the severity of their medical conditions,
who were not delirious at the time of admission (Lawlor and the diagnostic criteria used.

41. D E L I R I U M • 825
I M PAC T OF   DE L I R I U M cardiopulmonary edema, and cardiac arrhythmia (OR:  6.5;
95% CI:  2.7–15.6); longer length of stay in both the criti-
cal care unit (weighted mean difference [WMD]: 7.32 days;
MOR BI DI T Y A N D MORTA L I T Y R E L AT E D
95% CI:  4.63–10.01) and hospital (WMD:  6.53  days; 95%
TO DE L I R I U M
CI:  3.03–10.03); spent more time on mechanical ventila-
Across many hospital settings (e.g., emergency department, tion (WMD: 7.22 days; 95% CI: 5.15–9.29); and were more
general medical ward, postoperative, critical care units) and likely to be discharged to skilled placement (OR: 2.59; 95%
after controlling for multiple variables (e.g., demographics, CI: 1.59–4.21) (Zhang et al., 2013).
age, gender, illness severity, and comorbid medical condi- Delirium’s impact among the elderly is even greater. The
tions), patients who develop delirium fare much worse than mortality rate for elderly patients in acute care hospitals is
comparable nondelirious cohorts (Vasilevskis, Han, Hughes, much higher among those with delirium than those with-
Ely, 2012). A  recent meta-analysis, including 16 studies out delirium: 8% versus 1% in one study (Francis, Martin, &
(n  =  6410)  among critically ill patients, demonstrated that, Kapoor, 1990). Among hospitalized, elderly, medically ill
despite modern advances, when compared to their nonde- patients, incident delirium was associated with an excess
lirious counterparts, hospitalized patients with delirium had stay after diagnosis of 7.78  days, even after controlling for
higher mortality rates (odds ratio [OR]: 3.22; 95% confidence covariates (McCusker, Cole, et  al., 2003). Similarly, studies
interval [CI]: 2.30–4.52); had longer lengths of stay in both have demonstrated that delirium in the emergency depart-
the ICU (weighted mean difference [WMD]: 7.32 days; 95% ment was an independent predictor of prolonged hospi-
CI: 4.63–10.01) and in the general hospital (WMD: 6.53 days; tal length of stay (twice as long), compared to nondelirious
95% CI: 3.03–10.03); spent more time on mechanical ventila- patients, even after adjusting for confounding factors (Han
tion (WMD:  7.22  days; 95% CI:  5.15–9.29); experienced a et  al., 2011). Furthermore, the number of days of delirium
higher rate (6X) of complications (e.g., acute respiratory extu- older patients experience while in the critical care unit is
bation, removal of catheter, cardiac arrhythmia; OR: 6.5; 95% significantly associated with time to death within 1  year
CI: 2.7–15.6); and were more likely to be discharged to skilled post-ICU admission, after controlling for age, severity of ill-
placement (OR:  2.59; 95% CI:  1.59–4.21) (Zhang, Pan,  & ness, comorbid conditions, psychoactive medication use, and
Ni, 2013). baseline cognitive and functional status (hazard p = 0.001;
Delirious patients, when compared with patients suf- hazard ratio, 1.10; 95% CI, 1.02–1.18); see Thoracic Society
fering from the same medical problem who do not develop Journals, http://www.atsjournals.org/doi/abs/10.1164/rccm.
delirium as a complication, experience prolonged hospital 2 0 0 9 0 4 - 0 537O C? u rl _ver =Z 39.8 8 -2 0 03 & r f r_ id=
stays—five to ten days longer on average (Francis, Martin, & ori:rid:crossref.org&rfr_dat=cr_ pub%3dpubmed&#.
Kapoor, 1990; O’Keeffe & Lavan, 1997; Ely, Gautam, et al. VB9cea10wfg – Figure 2 Pisani, Kong, et al., 2009.
2001; Maldonado, Dhami, & Wise, 2003; Ely, Shintani et al., These bleak prognoses apply not only to severe cases; even
2004). Significantly more patients who develop delirium in patients with prevalent subsyndromal delirium (SSD) have
the hospital ultimately require institutional post-acute care, been shown to experience longer acute-care hospital stay,
such as placement in a skilled nursing facility (e.g., 16% for increased post-discharge mortality, more symptoms of delir-
delirium patients versus 3% for those without delirium), ium, and a lower cognitive and functional level at follow-up
even after controlling for illness severity, activities of daily than patients with no SSD (Cole, McCusker, Dendukuri, &
living (ADL) status, prior cognitive impairment, and fever Han, 2003). Among patients admitted to an intensive care
(Francis, Martin, & Kapoor, 1990; O’Keeffe & Lavan, 1997). unit, when comparing no delirium versus SSD, the mortality
Psychiatric hospital inpatients with delirium had hospital rate goes up from 2.4% to 10.6% (p < 0.001), ICU length of stay
stays that were 62.1% longer than those of patients without goes up from 2.5 to 5.2 days (p < 0.001), and the overall hos-
delirium (Ritchie, Steiner, & Abrahamowicz, 1996). pital length of stay goes up from 12.9 to 16.7 days (p < 0.001)
Among critically ill patients there is an independent (Ouimet, Kavanaugh, Gottfried, & Skrobik, 2007).
association between delirium present within 24 hours
after admission to the critical care unit with an increased
C O G N I T I V E A N D BE H AV IOR A L S E QU E L A E
in-hospital mortality (i.e., 16.2% for delirious patients vs.
5.7% for nondelirious; van den Boogaard et al., 2010). Others There appears to be a reciprocal relationship between delir-
have reported that patients who developed delirium in criti- ium and cognitive decline, with evidence demonstrating that
cal care units experience higher mortality, both at 90  days the presence of baseline cognitive deficits, including demen-
(11% vs. 3%; Pompei et al., 1994) and at 6 months (34% vs. tia, lowers the threshold to develop delirium (Inouye, 1998;
15%; Ely et  al., 2004; (See http://jama.jamanetwork.com/ Wahlund & Bjorlin, 1999; Litaker, Locala, Franco, Bronson, &
article.aspx?articleid=198503)). A  recent meta-analysis of Tannous, 2001; McNicoll et  al., 2003; Benoit et  al., 2005;
clinical observational studies in critically ill patients revealed Kalisvaart et al., 2006; Wacker, Nunes, Cabrita, & Forlenza,
that, when compared with their nondelirious counterparts, 2006; Smith, Attix, Weldon, Greene, & Monk, 2009; Franco,
delirious patients had higher mortality rate (OR:  3.22; Valencia, et al., 2010; Tognoni et al., 2011). Data suggests that
95% CI:  2.30–4.52); a higher rate of complications includ- baseline dementia is the strongest risk factor for delirium
ing removal of catheter, self-extubation, reintubation, acute among older patients (Elie et al., 1998; McCusker et al., 2003;
respiratory distress syndrome, nosocomial pneumonia, McAvay et al., 2006; Inouye et al., 2007). On the other hand,

826 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
the development of delirium appears to increase the risk of follow-ups, 79% and 71% of survivors had cognitive impair-
cognitive decline, including long-term cognitive impairment ment, respectively, with 62% and 36% being severely impaired
(LTCI) and dementia (Rockwood et al., 1999; Morandi et al., (Girard et al., 2010). These studies demonstrate that delirium
2012). A number of pathways and mechanisms have been pro- duration was independently associated with long-term cogni-
posed, yet, despite of some suggestive evidence, none has been tive outcome (i.e., the longer the duration of delirium, the more
conclusively proven as causative (see Box 41.1). significant the cognitive deficits). For example, an increase
Both clinical experience and research data suggests that from 1 day of delirium to 5 days was independently associated
many patients do not fully recover (cognitively) from delir- with nearly a 5-point decline (i.e., a one-half standard devia-
ium, particularly the elderly. The proportion of patients with tion [SD] decline) in the cognitive battery mean score (95%
residual cognitive impairment reported in various studies CI, –9.2 to –0.1).In fact, more than 20 prospective studies
depends on the clinical population under study, the cause (> 5,000 patients) during the last three decades demonstrate
of the delirium (when it is known), and the threshold and a significant association between delirium and long-term
criteria used for diagnosing. In fact, a substantial number cognitive dysfunction (MacLullich et al., 2009; Witlox et al.,
of patients who survive delirium are left with post-delirium 2010). The studies suggest that about 40% of patients with
cognitive impairment (Wacker et al., 2006; Griffiths & Jones, delirium develop some form of cognitive impairment when
2007; Bickel, Gradinger, Kochs, & Forstl, 2008; Kat et al., followed up from about 3 months to 5 years after an episode
2008; Maldonado, 2008b; Fong et al., 2009; MacLullich, of delirium (Levkoff et al. 1992, McCusker, Cole, Dendukuri,
Beaglehole, Hall, & Meagher, 2009; Girard et al., 2010). Data Belzile, & Primeau, 2001; Jackson et al., 2004; Witlox et al.,
demonstrates that, when compared to controls, patients diag- 2010). A comprehensive meta-analysis concluded that there is
nosed with delirium are much more likely to experience LTCI greater cognitive impairment among patients with delirium
(Macdonald, 1999; Rockwood et al., 1999; Jackson, Gordon, than matched controls; there is a higher incidence of dementia
Hart, Hopkins, & Ely, 2004; Gunther, Jackson, & Ely, 2007). in patients with a history of delirium; and that one of three
A study of mechanically ventilated patients who experienced survivors of critical illness with delirium developed cogni-
in-hospital delirium found that at 3-month and 12-month tive impairment (Jackson et al., 2004). A systematic search

Box 41.1 MECHANISMS MEDIATING DELIRIUM AND COGNITIVE IMPAIR MENT

1. A number of factors and mechanisms leading to delirium may also directly cause CNS damage and neuronal dysfunction, and thus
mediate both the manifestations of delirium and long-term cognitive impairment, including:
a) cytokine release and other neuroinflammatory mediators;
b) decrease perfusion and oxygenation, leading to decreased cerebral oxidative metabolism;
c) changes in blood–brain barrier permeability;
d) hyper-catabolic states;
e) water and electrolyte imbalances;
f) excessive glucocorticoid levels and other HPA-axis dysfunctions;
g) melatonin and sleep-wake cycle abnormalities
h) genetic factors (e.g., some have described an association between APOE4 and prolonged delirium duration)
i) Amyloid-β deposition (e.g., increased amyloid-β levels have been associated with subjective cognitive dysfunction in critical illness
survivors)
2. Pharmacological agents used to treat the underlying causes of the delirium (e.g., steroids, calcineurin inhibitors, other immunosup-
pressants, dopamine) or the agents used to treat delirium (e.g., dopamine-blocking agents, benzodiazepines) may themselves lead to
neuronal damage in a fragile brain.
3. Any of the mechanisms listed above may themselves lead to alterations in neurotransmitter concentration or receptor sensitivity,
which itself may underlie the different symptoms and clinical presentations of delirium and/or long-term cognitive dysfunction.
Thus, the same mechanisms that cause the substrate for delirium may mediate the cognitive impairments observed after the acute
presentation of delirium has resolved.
4. It is possible that instead of causing cognitive deficits or dementia, delirium (and its underlying causes) only serves as a catabolic
agent leading to an acceleration of normal, physiological, cerebral aging mechanisms leading to dementia.
5. It is also possible that an episode of delirium simply unmasks subtle cognitive deficits already present, although not yet identified.

41. D E L I R I U M • 827
of studies published between January 1981 and April 2010 Another study found that only 4% of delirious patients
found that delirium is associated with an increased risk of experienced full resolution of all symptoms of delirium
death compared with controls (38.0% vs. 27.5%); that patients before discharge from the hospital (Levkoff et  al., 1992).
who experienced delirium were at increased risk of institu- In this case, after following this sample longitudinally, an
tionalization (33.4% vs. 10.7%); and that patients who expe- additional 20.8% experienced resolution of symptoms by
rienced delirium were at increased risk of dementia (62.5% vs. the third month after hospital discharge, with an additional
8.1%) (Witlox et al., 2010). 17.7% by the sixth month after discharge from the hospital.
Some have postulated four possible putative mechanisms Among elderly hip surgery patients, delirium was a strong
or explanations for the relationship between delirium and independent predictor of cognitive impairment and the
long-term cognitive impairment: (1) that delirium is a marker occurrence of severe dependency in ADL. In fact, 38 months
of chronic progressive pathology but unrelated to any progres- after discharge from the hospital, 53.8% of the surviving
sion; (2) that delirium is a consequence of acute brain dam- patients with postoperative delirium continued to experi-
age, which is also responsible for a “single hit” or triggering of ence cognitive impairment, compared to only 4.4% of the
active processes causing LTCI; (3) that delirium itself is the nondelirious subjects (Bickel et  al., 2008). Similarly, a pro-
cause of the LTCI; and/or (4)  that pharmacological agents spective, matched, controlled cohort study of elderly hip sur-
used for the treatment of delirium or other conditions are gery patients demonstrated that the risk of dementia or mild
the cause of the LTCI associated with delirium (MacLullich cognitive impairment (MCI) over a 30-month follow-up was
et al., 2009). almost twice as high in patients with postoperative delirium
There appears to be a “dose effect” of delirium on cognition, as in those without (Kat et al., 2008).
with a longer duration of delirium found to be associated with A large systematic evidence review demonstrated that the
worse average performance on neuropsychological testing at 3 proportion of persistent delirium in older hospital patients at
and 12 months follow-up (p = 0.02 and p = 0.03, respectively) discharge, 1, 3, and 6 months was 44.7%, 32.8%, 25.6%, and
even after adjusting for age, education, preexisting cognitive 21% respectively; one study reported a proportion of 41% at
function, severity of illness, severe sepsis, and exposure to sed- 12 months after discharge (Cole, Ciampi, Belzile, & Zhong,
ative medications in the critical care unit (Girard et al., 2010). 2009)  (see Figure 41.1). The data suggest that many older
An increase from 1 to 5 days of delirium was independently hospital patients do not recover from delirium and that the
associated with a 7-point decline in the cognitive battery persistence of delirium is associated with adverse outcomes.
mean score at 12 months follow-up (p = 0.03) (Girard et al., In some of the studies, the long-term outcomes (mortality,
2010). Various factors may be associated with this dose effect: nursing home placement, cognition, function) of patients
neuroinflammatory changes; changes in brain–blood barrier with persistent delirium were consistently worse than the out-
permeability; the effect of psychoactive medications; anemia, comes of patients who had recovered from delirium.
hypoxia, anoxia, and other low-perfusion states; or systemic Furthermore, the data shows that SSD (associated with crit-
organ failure (Maldonado, 2008a, 2013). ical illness), in the absence of full-blown delirium, has also been
Yet, whatever the cause, there appears to be an association found to result in long-term cognitive dysfunction 2  months
between longer duration of delirium and greater brain atro- to 6 years following critical illness (e.g., 46%–70% of patients
phy as measured by a larger ventricle-to-brain ratio at hospi- showed signs of cognitive dysfunction at 1  year and 25% at
tal discharge (p = 0.03) and at 3-month follow-up (p = 0.05) 6 years) (Cole et al., 2003; Morandi, McCurley, et al., 2012).
(Gunther et  al., 2012). As expected, greater brain atrophy On the other hand, the data suggests that among elderly
(higher ventricle-to-brain ratio) at 3  months was associated patients there is a significant acceleration in the slope of cogni-
with worse cognitive performance (executive functioning tive decline in those with Alzheimer’s disease (AD) following
and visual attention) at 12 months (p = 0.04) (Gunther et al., an episode of delirium (Fong et al., 2009). In fact, prospective
2012). Others have found that the longer the duration of data from hospitalized patients with AD (n = 263; median
delirium, the greater the degree of white matter disruption follow-up duration 3.2  years) found that after adjusting for
in the genu of the corpus callosum and anterior limb of the dementia severity, comorbidity, and demographic character-
internal capsule (p = .01), up to 3 months after hospital dis- istics, patients who had developed in-hospital delirium expe-
charge (Morandi, Rogers, et al., 2012). rienced greater cognitive deterioration in the year following
A prospective study followed older hospitalized delirious hospitalization relative to patients who had not developed
patients who suffered from prevalent or incident delirium delirium (Gross et al., 2012). Cognitive deterioration follow-
detected during the first week of hospitalization for up to a ing delirium in the year after hospitalization accelerated to
year and found that 39%, 38.5%, and 48.9% of patients with twice the rate of patients who did not develop delirium. This
dementia met delirium criteria at discharge, 6-month, and rate of cognitive deterioration among delirious patients con-
12-month follow-ups, respectively, compared to 11.1%, 8.8%, tinued throughout a 5-year period following hospitalization
and 14.8% of patients without dementia (McCusker, Cole, (Gross et al., 2012). The Vantaa 85+ study (population-based
et al., 2003). However, cognitive improvement during hospi- cohort; n = 553 patients aged ≥85 years at baseline) followed
talization and long-term changes in the number of symptoms examined individuals at 3, 5, 8, and 10  years after delirium
were remarkably similar in the two groups, with inattention, and found that delirium increased the risk of incident demen-
disorientation, and impaired memory being the most com- tia (OR 8.7, 95% CI, 2.1–3.5) and was associated with the loss
mon symptoms. of an additional 1 point per year in the Mini-Mental State

828 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Discharge 1 month
Rockwood_89 Hodkinson_73
Levkoff_92
Williams_92
Rockwood_93 Jitapunkul_92
Gaudet_93
O’Keefe_97 Rudberg_97
Rudberg_97
Kelly_01
McCusker_03 Kelly_01
Pitkala_04
Lundstrom_05
McAvay_06 Kiely_04
Marcantonio_06
Inouye_07 Marcantonio_06
Combined
Combined* Combined

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Proportion Proportion

3 months 6 months
Levkoff_92 Levkoff_92

Treloar_97 Markup Area

McCusker_03

Kelly_01

Combined Combined
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Proportion Proportion

Individual and combined proportions (and 95% confidence intervals) for persistent delirium at discharge and one, three, and six months
Figure 41.1
after enrollment. Source: (Cole, Ciampi et al., 2009)

Examination compared to those with no history of delirium a delirious state. The incidence of PTSD related to postop-
(95% CI, 0.11–1.89) (Davis et al., 2012). erative delirium has been shown to be related to the nature
A prospective cohort study of 53 patients admitted for and level of the sedation and analgesia, degree of factual
hip fracture surgery aged 75 years and older found that at the memory recall, incidence of delirium, and underlying preva-
3 months preoperative assessment, patients who had experi- lence of preexisting psychiatric morbidity (Blank  & Perry,
enced postoperative delirium showed poorer performance 1984; Bourgon, 1985; Stukas et al., 1999; Dew et al., 2001;
on tests of global cognition and episodic memory, even after Jones, Griffiths, Humphris, Skirrow, 2001; Breitbart et  al.,
adjusting for age, gender, and baseline cognitive impairment 2002; DiMartini, Dew, Kormos, McCurry, & Fontes, 2007;
(Basinski, Alfano, Katon, Syrjala, & Fann, 2010). No differ- Griffiths  & Jones, 2007; Roberts, Rickard, Rajbhandari,  &
ence was found on tests of attention between patients with Reynolds, 2007; O’Malley, Leonard, Meagher,  & O'Keeffe,
and without history of delirium after adjusting for relevant 2008; Basinski et al., 2010).
confounders. Another study of elderly patients (≥60) under- Finally, a number of factors associated with the medical
going cardiac surgery (i.e., coronary-artery bypass grafting setting may exacerbate the experience of hypnagogic halluci-
or valve replacement) found that at 6 months post-surgery, a nations, such as medication use, physical restraints, and social
higher percentage of patients with delirium than those with- isolation (Jones et  al., 2001). Patients describe memories of
out delirium had not returned to their preoperative baseline these experiences as being very vivid, detailed, and frighten-
level of functioning (40% vs. 24%, p = 0.01), although that ing. A frightening “hallucinated” memory of a nurse trying
difference was not significant at 12  months (31% vs. 20%, to kill the patient or removing his organs, without a balanc-
p = 0.055) (Saczynski et al., 2012). ing memory of their medical condition and the care provided
A different type of residual syndrome is post-traumatic by the same nurse, is likely to contribute to the development
stress disorder (PTSD) secondary to the dramatic and bizarre of pseudo-memories with great emotional valence for that
delusional and hallucinatory experiences that occur during patient.

41. D E L I R I U M • 829
Studies have found that patients with delirium within the due to the greater need for long-term care or additional home
4 weeks after myeloablative hematopoietic cell transplanta- health care, rehabilitation services, and informal caregiving.
tion had significantly more distress and fatigue at 6 months A recent study looking at costs over one year following an epi-
(P < .004) and at 1 year (P < .03) and had worse symptoms sode of delirium estimated that delirium is responsible for an
of depression and post-traumatic stress at 1  year (P < .03), additional cost of $60,000 to $64,000 per patient for the year
compared with patients without delirium (Basinski et  al., following the index hospitalization (Leslie et al., 2008).
2010). Others have found that family members of delirious
ICU patients can also develop PTSD. Both family caregivers
(76%) and nurses (73%) have reported severe emotional dis- E T IOL O G Y OF   DE L I R I U M
tress related to a patient’s delirium (Breitbart et al., 2002). In
that study, family members were most distressed by patients’ Many factors potentially contribute to the development of
hyperactivity and functional impairment; nurses were most delirium (see Table 41.2). I use the mnemonic “End Acute
distressed by the severity of the patients’ symptoms and by the Brain Failure,” to help recall the 20 of the most clinically rele-
presence of perceptual disturbances. vant risk factors. The reason to use this term is to convey to all
Data suggest that the impact of delirium on outcome may members of the medical team the seriousness of delirium—
be changed by therapeutic interventions. (Please see the sec- indeed, during delirium, the brain is experiencing acute fail-
tion on delirium management later in this chapter.) ure, with potentially disastrous consequences. Only some
of the risk factors are discussed here; readers are referred to
my comprehensive review of delirium risk factors for further
T H E C O S T OF DE L I R I U M
details (Maldonado, 2008b). In general, risk factors can be
The increased morbidity and extended hospital care associ- grouped as non-modifiable and potentially modifiable.
ated with delirium have been associated with greater care
costs (Inouye, 2000; Ely et al., 2004; Milbrandt et al., 2004).
NON-MODI F I A BL E R I S K FAC TOR S
Delirious hospital inpatients require more nursing time per
patient and have higher per-diem hospital costs, in addition to Recognition of the non-modifiable factors can help identify
their increased lengths of stay (Siddiqi, Stockdale, Britton, & patients at high risk for delirium in order to provide enhanced
Holmes, 2007). The economic impact of delirium is sub- surveillance and implement preventive measures. (Table 41.3).
stantial, rivaling the health care costs of falls and myocardial Four important non-modifiable factors are older age, baseline
infarction (Hall et al., 1988; Rizzo, McAvay & Tinetti, 1996; cognitive impairment, severity of underlying medical illness,
Inouye, 2000). and preexisting mental disorders.
A retrospective chart review of patients who received
step-down critical care at a large university hospital showed
Age
that the 14% of the ICU population who developed delirium
utilized 22% of total inpatient days (Maldonado, Dhami, & Old age is likely to be a contributor due to a greater number
Wise, 2003). The treated group consisted of the patients for of medical comorbidities and overall frailty. The aging pro-
whom a psychosomatic medicine consult was obtained and cess itself is associated with some degree of cognitive deficits
the team implemented at least 80% of the recommended regi- and increased risk of dementia. Finally, aging is associated
men (as detailed in the treatment section at the end of this with age-related cerebral changes in stress-regulating neu-
chapter). The average number of days from symptomatic onset rotransmitter and intracellular signal transduction systems.
to resolution was 10.8 days for untreated patients and 6.3 days Chronic neurodegeneration is accompanied by an inflamma-
for treated patients. As a group, delirious patients were older tory response characterized by a chronic, but selective, acti-
(71.3 vs. 63.6  years), remained hospitalized longer (16.4 vs. vation of central nervous system (CNS) microglial cells that
6.6  days), and had greater total costs per case ($63,900 vs. are “primed” to produce exaggerated inflammatory responses
$30,800). Another study of ICU delirium demonstrated that to immunological challenges. This inflammation is assumed
even after adjusting for age, comorbidity, severity of illness, to contribute to disease progression through the produc-
degree of organ dysfunction, nosocomial infection, hospital tion of inflammatory mediators, including cytokines and
mortality, and other potential confounders, delirium was acute phase proteins (Cunningham, Wilcockson, Campion,
associated with 39% higher ICU and 31% higher hospital Lunnon, & Perry, 2005,; Cunningham et al., 2009). Aging
costs (Milbrandt et al., 2004). They also found that the sever- is also associated with decreased volume of acetylcholine
ity and duration of delirium were associated with incremental (Ach)-producing cells and decreased cerebral oxidative
greater care cost. Similarly, a study of hospitalized medically metabolism. Thus, the decline in cognitive functioning asso-
ill elderly patients the average cost per day was 2.5-fold greater ciated with the normal aging process is aggravated by the
in those with delirium, and the total excess cost attributable to presence of even mild hypoxia, which further inhibits ACh
delirium ranged from $16,303 to $64,421 per patient (Leslie, synthesis and its release.
Marcantonio, Zhang, & Leo-Summers, 2008). Among elderly patients in the ICU, the probability of
The total annual direct healthcare costs attributable to developing delirium increases by 2% per year of age for each
delirium in the United States might be as high as $152 billion year after age 65 (Figure 41.2) (Pandharipande et al., 2006).
(Leslie et al., 2008). These costs accrue after hospital discharge Many other studies confirm that older age is an independent

830 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Table 41.2 DELIR IUM: PR EDISPOSING AND PR ECIPITATING R ISK FACTORS—“END ACUTE BR AIN FAILUR E”

R ISK FACTOR EX A MPLES

Electrolyte imbalance and Electrolyte disturbances (e.g., hyperammonemia, hypercalcemia, hypo/hyperkalemia,


dehydration hypomagnesemia, hypo/hypernatremia)

Neurological disorder and injury All neurological disorders: e.g., CNS malignancies, abscesses, cerebrovascular accident (CVA),
vasculitis, multiple sclerosis (MS), epilepsy, Parkinson’s disease, normal pressure hydrocephalus
(NPH), traumatic brain injury (TBI), diffuse axonal injury (DAI), limbic encephalitis (both
non-paraneoplastic and paraneoplastic syndrome).
Of the various forms of sensory impairment, only visual impairment has been shown to contribute
to delirium. Visual impairment can increase the risk of delirium 3.5-fold.

Deficiencies (nutritional) Nutritional deficiencies (e.g., malnutrition, low serum protein/albumin, low caloric intake,
“failure to thrive”), malabsorption disorders (e.g., celiac disease), and hypovitaminosis; specifically
deficiencies in cobalamine (B12), folate (B9), niacin (B3; leading to pellagra), thiamine (B1; leading to
beriberi and Wernicke’s disorder).

Age Age [>65] and Gender [m>f]

Cognition Baseline cognitive functioning, including dementia and other cognitive disorders and/or a past
history of delirium have all been shown to increase the likelihood of delirium.

U-Tox (intoxication and withdrawal) Substances of abuse—Acute illicit substance intoxication (e.g., cocaine, PCP, LSD, hallucinogens),
as well as poisons, pesticides, solvents, and heavy metals (i.e., lead, manganese, mercury)—and
substances withdrawal.

Trauma Physical trauma and injury; heat stroke, hyperthermia, hypothermia, severe burns, including
Toxins trauma of surgical procedures.
Various toxins, including bio-toxins (animal poison); heavy metals (lead, manganese, mercury);
insecticides; poisons; carbon dioxide; toxic effect of pharmacological agents (i.e., serotonin
syndrome, neuroleptic malignant syndrome, anticholinergic states); blood levels (toxic levels of
various therapeutic substances; e.g., lithium, VPA, carbamazepine, immunosuppressant agents).

Endocrine disturbance Endocrinopathies such as hyper/hypo-adrenal corticoid; hyper/hypoglycemia; hyper/


hypothyroidism.

Behavioral—psychiatric Certain psychiatric diagnoses, including undue emotional distress; a history of alcohol and other
substance abuse, as well as depression, schizophrenia, and bipolar disorder have been associated
with a higher incidence of delirium.

Rx and other toxins Several pharmacological agents have been identified, especially those with high anticholinergic
activity, including prescribed agents (especially narcotics and GABA-ergic agents) and various OTC
agents (especially anticholinergic substances; polypharmacy).

Anemia, anoxia, hypoxia, and low Any state that may contribute to decreased oxygenation (e.g., pulmonary or cardiac failure,
perfusion states hypotension, anemia, hypoperfusion, intraoperative complications, hypoxia, anoxia, carbon
monoxide poisoning, shock).

Infectious Pneumonia, urinary tract infections, sepsis, encephalitis, meningitis, HIV/AIDS.

Noxious stimuli (pain) Data suggest that both pain and medications used for the treatment of pain have been associated
with the development of delirium. Studies have demonstrated that the presence of postoperative
pain is an independent predictor of delirium after surgery. On the other hand, the use of opioid
agents has also been implicated in the development of delirium.

Failure (organ) Organ and systemic failure; usually cardiac, hepatic, pulmonary, and renal failure.

APACHE score (severity of illness) Evidence shows that the probability of transitioning to delirium increases dramatically for each
additional point in the Acute Physiology and Chronic Health Evaluation (APACHE II) severity of
illness score.

Intracranial processes Stroke (especially non-dominant hemispheric); intracranial bleed; meningitis; encephalitis;
neoplasms.

Light, sleep, and circadian rhythm Sleep deprivation and insomnia, sleep disorders (e.g., obstructive sleep apnea) and disturbances/
reversal in sleep-wake cycle.
(continued)
Table 41.2 (CONTINUED)

R ISK FACTOR EX A MPLES

Uremia and other metabolic Acidosis, alkalosis, hyperammonemia, hypersensitivity reactions; glucose, acid–base disturbances.
disorders

Restraints and any factors causing The use of restraints, including endotracheal tubes (ventilator), soft and leather restraints,
immobility intravenous lines, bladder catheters, and intermittent pneumatic leg-compression devices, casts, and
traction devices all have been associated with an increased incidence of delirium.
Emergence delirium Emergence from medication-induced sedation, coma, or paralysis; which may be associated with
CNS depressant withdrawal, opioid withdrawal, REM-rebound, sleep deprivation.

risk factor among medically ill and surgical patients, with the 2006). A  study of nondemented elderly patients undergo-
increase in rate per year dependent on the specific context in ing elective orthopedic surgery demonstrated that subtle
which incidence is measured (Milstein, Pollack, Kleinman, & preoperative attention deficits, as tested by digit vigilance
Barak, 2002; Khurana, Gambhir, & Kishore, 2011; Lee et al., and reaction time testing, were closely associated with post-
2011; Rudolph et al., 2011; Srinonprasert et al., 2011). operative delirium (Lowery, & Ballard, 2007). In fact, these
subtle changes predicted a 4-fold to 5-fold increased risk of
postoperative delirium for subjects >1 SD above the sample
Baseline Cognitive Impairment
means on these variables. Of note, in this study none of the
Baseline cognitive deficits, even those not rising to the level previously identified risk factors for delirium was significantly
of dementia, significantly increase the risk of developing associated with delirium (i.e., urea/creatinine ratio, visual
delirium. It is well established that individuals with compro- impairment, blood pressure, sodium levels, hematocrit, bili-
mised cognitive ability preoperatively (e.g., dementia) are at rubin). Finally, a study of nearly 1,000 surgical patients found
greater risk of delirium (Franco et  al., 2010; Inouye, 1998; that preoperative executive dysfunction was associated with
Litaker et al., 2001; McNicoll et al., 2003; Benoit et al., 2005). a greater incidence of postoperative delirium, independent of
But recent evidence suggests that decrements in higher order other risk factors (Smith et al., 2009). Furthermore, patients
cognitive functions, such as executive function (e.g., problem exhibiting both executive dysfunction and clinically signifi-
solving, processing speed, planning, complex sequencing, cant levels of depression were at greatest risk for developing
and reasoning), may also predict postoperative delirium in delirium postoperatively.
the absence of frank cognitive impairment (Rudolph et  al., A study of elderly patients undergoing hip surgery found that
preoperative Mini-Mental State Exam (MMSE) scores were an
independent predictor of postoperative delirium (Kalisvaart
Table 41.3 DELIR IUM R ISK FACTORS et al., 2006). Similarly, a study of elderly subjects undergoing
NON-MODIFI ABLE
hip or knee replacement demonstrated that the presence of
MODIFI ABLE FACTORS FACTORS

• Various pharmacological agents, • Older age


especially GABA-ergic and • Baseline cognitive
opioid agents, and medications impairment 0.85
with anticholinergic effects • Severity of underlying
0.80
Probability of Transitioning

• Prolonged and/or uninterrupted medical illness


sedation • Preexisting mental 0.75
Immobility disorders
to Delirium


• Acute substance intoxication 0.70
• Substance withdrawal states
• Use of physical restraints 0.65
• Water and electrolyte imbalances
• Nutritional deficiencies 0.60
• Metabolic disturbances and p = 0.004
endocrinopathies (primarily 0.55
deficiency or excess of cortisol) 30 40 50 60 70 80
• Poor oxygenation states (e.g.,
hypo-perfusion, hypoxemia, Age (years)
anemia)
• Disruption of the Age and the probability of transitioning to delirium. The most
Figure 41.2

sleep-wake cycle notable finding related to age was that probability of transitioning


• Uncontrolled pain to delirium increased dramatically for each year of life after 65 years.
• Emergence delirium Adjusted Odds Ratio –1.01 (1.00, 1.02) p = 0.03. Abbreviations: Y-axis =
probability; X-axis = age in years. Source: (Pandharipande et al., 2006)

832 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
dementia increased the incidence of postoperative delirium study of elderly patients undergoing hip surgery (Kalisvaart
from 32% to 100% (Wacker et al., 2006). Approximately 70% et  al., 2006) and another, of critically ill patients (Tsuruta
of elderly patients admitted to a specialized delirium ward had et al., 2010), where APACHE II scores were identified as an
a preexisting cognitive disorder—either dementia or MCI independent predictor of delirium when controlling for the
(Wahlund & Bjorlin, 1999). A recent study of elderly patients use of mechanical ventilation and length of ICU stay.
undergoing urological surgery confirmed that age, baseline
cognitive impairment, and previous history of delirium all
were associated with a higher rate of postoperative delirium Preexisting Mental Disorders
(Tognoni et al., 2011). The relationship between cognitive defi- A 1996 study retrospectively investigated incidence rates
cits and dementia seems to be reciprocal. So, while it is clear and risk factors for delirium among hospitalized psychiat-
from the evidence presented above that the presence of baseline ric patients (Ritchie, Steiner,  & Abrahamowicz, 1996). The
cognitive deficits, including dementia, lowers the threshold to overall incidence of delirium in the study sample was 15%.
develop delirium, data suggest that among elderly patients there Patients with bipolar disorder had the highest incidence of
is a significant acceleration in the slope of cognitive decline in delirium (36%). Only 48% of delirious patients were actually
patients with AD following an episode of delirium (Fong et al., recognized as having delirium at the time it occurred. The
2009). This raises the question of whether prevention of delir- hospital stays of patients with delirium were 62% longer than
ium might ameliorate or delay cognitive decline in patients at those without.
risk, particularly those with AD. Delirious mania, also known as “Bell’s mania,” is defined
as “a syndrome of acute onset of excitement, grandiosity,
Severity of Medical Illness emotional lability, delusions and insomnia characteristic of
mania, and the disorientation and altered consciousness char-
Studies have also shown that the severity of the patient’s acteristic of delirium. Almost all patients exhibit signs of cata-
underlying medical problems has a significant effect on the tonia” (Bell, 1849; Fink, 1999). The syndrome consists of a
development and progression of delirium (Ouimet et al., constellation of symptoms that can arise from both psychotic
2007; Girard et al., 2008; Maldonado, 2008b; Khan, Kahn, and affective psychiatric diseases, as well as from many medi-
& Bourgeois, 2009; Pisani, Kong, et al., 2009; Maldonado, cal diseases. It has been described as a subtype of the catatonia
2011; Srinonprasert et al., 2011; Ryan et al., 2013). A study syndrome, having a rapid onset with signs of mania, delirium,
of adult, mechanically ventilated ICU patients found that and catatonia (Friedman, Mufson, Eisenberg, & Patel, 2003).
increased severity of illness, as measured by the modified Among patients undergoing cardiac surgery—after adjust-
Acute Physiology and Chronic Health Evaluation (APACHE ment for sex, older age, cross-clamp time, hemoglobin, and
II; a modified-APACHE II is reached by subtracting the psychotropic drug use—major depression was significantly
Glasgow Coma Scale score from the usual calculation of the associated with delirium (OR  =  3.86, 95% CI 1.42–10.52,
APACHE II total score) was associated with a greater prob- p = 0.001) (Tully, Baker, Winefield, & Turnbull, 2010). A more
ability of developing delirium (Pandharipande et al., 2006). recent study of cardiac surgery patients found that patients
The incremental risk increased with APACHE score until the with a preoperative diagnosis of depression had an indepen-
latter reached 18 (see Figure 41.3), suggesting that for each dently associated increased risk of developing postsurgical
additional point on the APACHE II scale, the probability of delirium (Kazmierski, Banys, Latek, Bourke,  & Jaszewski,
delirium increased by 6%. These findings were replicated in a 2013). Thus, major depressive disorder has been associated
with incident delirium after various types of surgical proce-
dures, as well as among general hospital inpatients, indepen-
0.70
dent of other risk factors (Schneider et al., 2002; Dasgupta &
Dumbrell, 2006; Kazmierski et  al., 2006; McAvay et  al.,
Probability of Transitioning

0.65 2007; Smith et al., 2009). A study of geriatric surgical subjects


showed that patients with a greater number of preoperative
to Delirium

0.60 depressive symptoms were more likely to develop postop-


erative delirium and experience a longer duration of postop-
0.55
erative delirium, even after adjusting for covariates (i.e., age,
0.50
educational level, functional status, and preoperative alcohol
use) (Leung, Sands, Mullen, Wang, & Vaurio, 2005).
0.45 p = 0.004
Long-term benzodiazepine use is usually associated with
the presence of an underlying psychiatric disorder (e.g., anxiety,
10 15 20 25 30 depression, insomnia, alcohol abuse, benzodiazepine abuse), and
APACHE II Score thus it may be considered a marker of underlying psychiatric
comorbidity. Studies have looked at the presence of postoperative
Severity of Illness and the Probability of Transitioning to
Figure 41.3
delirium between long-term benzodiazepine users and nonusers
Delirium. The probability of transitioning to delirium increased dra-
matically for each additional point in the Acute Physiology and Chronic and found that the incidence of delirium after orthopedic sur-
Health Evaluation II (APACHE II) severity of illness score, until reach- gery in the elderly was significantly more frequent in long-term
ing a plateau APACHE score of 18. (Pandharipande et al., 2006) benzodiazepine users (26%) than in short-term users (13%)

41. D E L I R I U M • 833
(Kudoh,Takase, Takahira, & Takazawa, 2004). Benzodiazepines 1989; Wetterling, Kanitz, Veltrup, & Driessen, 1994; Aldemir
have been demonstrated to have a negative effect on memory et  al., 2001; Lawlor, 2002; Sagawa, Akechi, et  al., 2009;
(Lister, 1985; Ghoneim  & Mewaldt, 1990; Ghoneim, Block, Caplan & Chang, 2010). The correction of these electrolyte dis-
Ping, el-Zahaby, & Hinrichs, 1993), and it has been demonstrated turbances has been shown to significantly shorten the duration
that benzodiazepine use induces both nonamnestic and amnes- of delirium (Koizumi, Shiraishi, Ofuku, & Suzuki, 1988). The
tic MCI (Tannenbaum, Paquette, Hilmer, Holroyd-Leduc,  & relationship between nutritional deficiencies and delirium has
Carnahan, 2012). Furthermore, chronic BZDP users had a sig- long been documented, and it is the clearest in patients suffering
nificantly higher risk of cognitive decline in global cognitive and from Wernicke’s encephalopathy (B1 or thiamine deficiency)
attention tests compared with non-BZDP users (Mura et  al., (Hoes, 1979; Newman, Grocott, et al., 2001; Onishi, Sugimasa,
2013). It appears that even though there may be an improvement Kawanishi, & Onose, 2005). In addition, vitamin B6 and B12
after BZDP discontinuation, there remains a significant impair- deficiency have all been associated with delirium (Peters  &
ment in most areas of cognition when compared to controls Neumann, 1960; Hoes, 1979; Buchman, Mendelsson, et  al.,
(Barker, Greenwood, Jackson, & Crowe, 2004, 2005). 1999; Lerner & Kanevsky, 2002). Some data suggest that vita-
As we discussed before, there is a relationship between min B12 supplementation (in patients with documented defi-
dementia and development of delirium. New data sug- ciencies) reduced the duration of delirium in elderly demented
gest that the use of benzodiazepine is associated with an subjects. And, others have found a relationship between low
increased risk of Alzheimer’s disease (adjusted OR 1.51, 95% albumin levels and the development of delirium among elderly
CI, 1.36–1.69) (Billioti de Gage et  al., 2014). Furthermore, demented patients (Culp & Cacchione, 2008).
the risk increases with cumulative exposure:  1.32 (1.01 to
1.74) for 3 to 6  months, and 1.84 (1.62 to 2.08) for more
than 6 months (Billioti de Gage et al., 2014). Similarly, in a Metabolic Disturbances and Endocrinopathies
large nested-control study, after controlling for multiple vari- Metabolic disturbances and various endocrinopathies have
ables (i.e., age, gender, education level, living alone, alcohol been associated with the development of delirium, primar-
consumption, psychiatric history, and depressive symptom- ily deficiency or excess of cortisol (Thiele  & Hohmann,
atology), the ongoing use of benzodiazepines was associated 1961; Millerowa, 1966; McIntosh et al., 1985; Basavaraju &
with a significantly increased risk of dementia (adjusted OR, Phillips, 1989; Olsson, Astrom, Eriksson,  & Forssell, 1989;
1.7; 95% CI, 1.2–2.4), while former use was associated with Fassbender, Schmidt, Mossner, Daffertshofer,  & Hennerici,
a significantly increased risk of dementia (adjusted OR, 2.3; 1994; O’Keeffe & Devlin, 1994; Johansson, Olsson, Carlberg,
95% CI,1.2–4.5) (Lagnaoui et al., 2002). Others have found Karlsson,  & Fagerlund, 1997; Olsson, 1999; Robertsson
similar results, with a marked increased incidence of dementia et al., 2001; Abildstrom, Christiansen, Sirsma, & Rasmussen,
(OR = 3.50, 95% CI 1.57 to 7.79, p = 0.002) among those on 2004; Nemoto, Kawanishi, Suzuki, Mizukami,  & Asada,
long-term BZDP use (Gallacher et al., 2012). Finally, among 2007; Weng, Chang,  & Weng, 2008)  or thyroid hormone
subjects seeking acute treatment for alcohol withdrawal, 6.9% (Vidal  & Vidal, 1961; Goldfarb, Varma,  & Roginsky, 1980;
developed alcohol withdrawal delirium despite adequate Nibuya, Suda, Mori, & Ishiguro, 2000; El-Kaissi, Kotowicz,
treatment with benzodiazepine agents (Palmstierna, 2001). Berk, & Wall, 2005), and either high or low blood glucose lev-
These findings raise the possibility that the use of benzodiaze- els (Fishbain & Rotundo, 1988; Aldemir, Ozen, et al., 2001;
pines, when combined with other risk factors, may contribute Miller, 2008). There have been case reports of Cushing’s
or accelerate the rate of cognitive decline following delirium. disease caused by adrenocorticotropic hormone (ACTH)–
producing pituitary adenomas, in patients who lacked the
phenotypical signs of hypercortisolism (e.g., facial plethora,
P OT E N T I A L LY MODI F I A BL E R I S K FAC TOR S
supraclavicular fat pads, buffalo hump, truncal obesity, and
Among the significant factors that are amenable to early purple striae) and whose most prominent and distressing
intervention or prevention are electrolyte imbalances and symptoms were severe myopathy and altered mental status
nutritional deficiencies, metabolic disturbances and endo- (Tran  & Elias, 2003). Similarly, there have been reports of
crinopathies, low oxygenation states, environmental factors hypoactive delirium as the primary presentation of a patient
disrupting the sleep-wake cycle and impeding adequate rest, with tonsillar abscess and associated panhypopituitarism
pain and its treatment, the use of drugs with anticholinergic and secondary hyponatremia (Umekawa Yoshida, Sakane, &
effects, substance intoxication and withdrawal states, the use Kondo, 1996)  and of a patient experiencing delirium sec-
of physical restraints, and the development of “emergence ondary to serendipitously diagnosed hypocortisolemia, who
delirium” (i.e., an altered mental status occurring as a patient presented none of the usual signs of adrenal insufficiency or
“emerges” from deep sedation or medication-induced coma). endocrinopathy (Fang & Jaspan, 1989). Others have demon-
strated that, among demented inpatients with no acute medi-
cal illness, there was a strong relationship between delirium
Electrolyte Imbalances and Nutritional Deficiencies
and dexamethasone suppression test (DST) pathology, irre-
Irrespective of the etiology, a water and electrolyte imbalance spective of age and the severity of dementia; suggesting the
(primarily magnesium, phosphate, potassium, and chloride) possibility that an impaired hypothalamus–pituitary–adrenal
provoking a hypo- or hyperosmolar state causes metabolic (HPA) system and a low delirium threshold may lead to delir-
encephalopathy or delirium (Mattle, 1985; Schmickaly et al., ium as a response to stress (Robertsson et al., 2001).

834 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Low Oxygenation States Premorbid Hemoglobin
12
Poor oxygenation (i.e., hypoperfusion, hypoxemia, anemia)
has long been associated with the development of delirium
(Kinney, Duke, et al., 1970; Weissman et al., 1984; McGee,
Veremakis, & Wilson, 1988; Fine, Anderson, Rothstein, 11
Williams, & Gochuico, 1997; Maldonado, 2008a,b).

gm/dl
Inadequate oxidative metabolism may be one of the under-
lying causes of the basic metabolic problems initiating the
10
cascade that leads to the development of delirium; namely,
inability to maintain ionic gradients, causing cortical spread-
ing depression (i.e., spreading of a self-propagating wave of
cellular depolarization in the cerebral cortex) (Moghaddam, 9
Schenk, Stewart, & Hansen, 1987; Somjen et al., 1989, 1990; Delirium Non-delirium
Iijima, Shimase, Iwao, & Sankawa, 1998; Shimizu-Sasamata, (N = 30) (N = 70)

Bosque-Hamilton, Huang, Moskowitz, & Lo, 1998; Basarsky, Note: p<0.004; Student’s t-test with equal variances not assumed
(Levene’s test).
Feighan, & MacVicar, 1999); abnormal neurotransmit-
ter synthesis, metabolism, and release (Globus, Busto et al., Premorbid Hematocrit
1988b, 1989, 1990; Busto et al., 1989; 1990; Globus, Wester, 36
Busto, & Dietrich, 1992; Takagi, Ginsberg, et al., 1994; Busto,
Dietrich, Globus, Alonso, & Dietrich, 1995; Busto, Dietrich,
et al., 1997); and a failure to effectively eliminate neurotoxic
34
byproducts (Busto, Globus, et al., 1989; Globus, Busto, et al.,
1990; Globus, Alonso, et al., 1995).
Percent
Study findings correlate the development of delirium
with findings suggestive of premorbid oxidative metabo- 32
lism impairment. In a study of ICU patients, three measures
of oxygenation (i.e., hemoglobin, hematocrit, and pulse
oximetry) were worse in the patients who later developed
30
delirium (Figure 41.4) (Seaman, Schillerstrom, Carroll,  &
Delirium Non-delirium
Brown, 2006). Similarly, two measures of oxidative stress (N = 30) (N = 70)
(sepsis, pneumonia) occurred more frequently among those Note: p<0.006; Student’s t-test with equal variances not assumed.
diagnosed with delirium. Others have demonstrated that
intraoperative cerebral oxygen desaturation (as measured Premorbid Pulse-Oximetry
by cerebral oximeter) is significantly associated with an 94
increased risk of postoperative delirium and early cognitive
decline and was associated with a nearly threefold increased
risk of prolonged hospital stay after cardiac surgery (CABG)
(Figure 41.5) (Slater et al., 2009). Finally, another large study 91
Percent Saturation

of urgent cardiac surgery demonstrated that preoperative


and intraoperative ScO2 readings were lower in the group
of patients who developed delirium. The binary logistic
regression identified older age, lower MMSE, neurological 88
or psychiatric disease, and lower preoperative ScO2 as inde-
pendent predictors of postoperative delirium (Figure 41.6)
(Schoen et al., 2011).
85
Delirium Non-delirium
(N = 30) (N = 70)
Sleep Disturbances Note: p<0.006; Student’s t-test with equal variances not assumed.

Sleep is another factor that seems to play a significant role in Oxidative Stress Hypothesis. Three measures of oxygenation
Figure 41.4
developing delirium in the ICU. Sleep deprivation has long (Hg, Hct, pulse ox.) were worse in the patients who later developed
been linked to the development of delirium (Lipowski, 1987) delirium.
and psychosis (Berger, Vollmann, et al., 1997). Patients in the
ICU experience severe alterations of sleep, with sleep loss, per 24-hour period) (Aurell & Elmqvist, 1985). Sleep in ICU
sleep fragmentation, and sleep–wake cycle disorganization. patients is characterized by prolonged sleep latencies, sleep
Continuous polysomnographic recordings have demonstrated fragmentation, decreased sleep efficiency, frequent arousals, a
that despite sedative practices, the average amount of sleep in predominance of stage 1 and stage 2 non-rapid eye movement
ICU patients is limited (e.g., as short as 1 hour and 51 minutes (REM) sleep, decreased or absent stage 3 and 4 non-REM sleep,

41. D E L I R I U M • 835
changes/placement), noise, light, and discomfort (Weinhouse
8 & Schwab 2006; Drouot et al., 2008; Friese, 2008b). Another
7 large factor in sleep debt in ventilated patients is continuous
sedation. In fact, studies have demonstrated that propofol and
6 benzodiazepine agents (the primary agents used for continu-
5 ous sedation) have a negative effect on sleep (e.g., severe altera-
tions of sleep with sleep loss, ↓ REM sleep, loss of circadian
4 rhythm, extreme sleep fragmentation, sleep–wake cycle dis-
3 organization) (Drouot et al., 2008; Matthews, 2011; Boyko,
Ording, & Jennum, 2012; Hofhuis, Langevoort, Rommes, &
2
Spronk, 2012; Kamdar, Needham, & Collop, 2012; McKinley
1 et al., 2012; Nakos, 2012; Watson, Ceriana, & Fanfulla, 2012;
Andersen, Boesen, & Olsen, 2013; Brummel & Girard, 2013;
0
Gomez Sanz, 2013).
No Cerebral Cerebral
Desaturation Desaturation

Cerebral Desaturation and Postoperative Delirium. Intra-


Figure 41.5
Pain Management
operative cerebral O2 desaturation was a significant risk factor for As in the case with sleep, both the experience of pain and some
postoperative delirium in CABG patients. (Slater et al., 2009)
of the pharmacological agents used for the treatment of pain
have been associated with the development of delirium. In fact,
and decreased or absent REM sleep (Weinhouse & Schwab, the presence of postoperative pain has been shown to be an
2006; Drouot, Cabello, d’Ortho, & Brochard, 2008; Friese, independent predictor of postoperative delirium (Vaurio et al.,
2008b). 2006). There is also a direct relationship between levels of preop-
There are many factors that may contribute to sleep debt erative pain and the risk of developing postoperative delirium.
in the hospital, and particularly in the ICU. These include fre- Unfortunately, some opioid agents have also been implicated
quent diagnostic and therapeutic interventions and procedures; in the development of delirium (Fong, Sands, & Leung, 2006;
anxiety, fear, pain, and underlying illness and disease severity; Vella-Brincat & Macleod, 2007; Wang, Sands, Vaurio, Mullen,
and environmental factors, such as around-the-clock medi- & Leung, 2007). In fact, opioid agents have been implicated in
cal care (e.g., suctioning, ventilator settings, blood draws, line nearly 60% of the cases of delirium in patients with advanced
cancer (Centeno, Sanz, & Bruera, 2004). Studies among cancer
patients revealed significant associations between opioids and
30
delirium (Gaudreau et al., 2007). The association remained
no Delirium significant after adjustment for corticosteroid, benzodiazepine,
25
Delirium and antipsychotic exposure using generalized estimating equa-
p = 0.024 tion regressions (OR of 1.37; P = .0033). Others have suggested
that patients who used oral opioid analgesic agents as their
20
sole means of postoperative pain control are at decreased risk
of developing delirium compared to subjects using intravenous
Delta ScO2base

(IV) opioid-based, patient-controlled analgesia (PCA) (Vaurio


15 et al., 2006; Wang et al., 2007).
Some opioid agents (e.g., meperidine) may have greater
deliriogenic potential than others (Marcantonio et al., 1994;
10 Morrison et al., 2003; Fong et al., 2006). Studies suggest that
an opioid rotation to less deliriogenic agents like fentanyl or
hydromorphone has been associated with improved pain man-
5 agement and lower delirium rating scores (Morita et al., 2005).
The exact mechanism of delirium causation is unclear, but
agent half-life, the presence of active metabolites, and anticho-
0 linergic potential have all been implicated in opioid-induced
cerebral oxygen saturation ScO2base>59.5% delirium (Slatkin & Rhiner, 2004). Some clinicians prefer the
(ScO2) base<59.5%
use of PCA, as the common wisdom is that patients usually
Preoperative and Intraoperative Cerebral O2 Saturation
Figure 41.6 “self-titrate” and end up receiving lower doses of medication,
(ScO2) Values Influenced the Risk of Post-op Delirium Intraoperative thus lowering delirium risk. This may be true in the cognitively
changes in ScO2 in patients with or without delirium classified by intact individual; but in the cognitively impaired or already
normal or low preoperative ScO2. Delta ScO2base, difference between delirious patient, PCAs are contraindicated. The risk of a PCA
preoperative regional cerebral oxygen saturation with oxygen supple-
mentation and minimal intraoperative regional cerebral oxygen satura- apparatus for a patient who is already confused is that they may
tion; ScO2base, regional cerebral oxygen saturation with supplemental not know how to use it, what to do with the button. Essentially
oxygen. (Schoen et al., 2011) the question to answer is, is the patient cognitively intact and

836 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
competent enough to use PCA appropriately? A simple rule: if Table 41.4 COMMONLY USED MEDICINES THAT
the patient is experiencing any sign of confusion (e.g., disorien- HAVE ANTICHOLINERGIC EFFECTS
tation, finger agnosia), do not use a PCA. This may be resumed
ANTIHISTA MINES
when the patient’s delirium resolves.
Chlorpheniramine
Medications Cyproheptadine
Besides opioid agents, a great number of medications have Dexchlorpheniramine
been associated with an increased risk of delirium (Table
41.4). Some have said that the highest incidence of medica- Diphenhydramine
tion-induced delirium has been observed in patients tak- Dimenhydrinate
ing the greater number of agents (Inouye & Charpentier,
1996), agents with high anticholinergic potential (Tune, Carr, Hydroxyzine
Cooper, Klug, & Golinger, 1993), and medications with signif- Meclizine
icant psychoactive effects (Francis, Martin, & Kapoor, 1990;
Olofsson, Weitzner, Valentine, Baile, & Meyers, 1996; Brauer, Promethazine
Morrison, Silberzweig, & Siu, 2000; Lawlor et al., 2000; ANTIPAR K INSONI AN
Tuma & DeAngelis, 2000; Morita et al., 2001; Breitbart et al.,
2002; Gaudreau & Gagnon, 2005). Among this latter group, Benztropine
opioids, corticosteroids, and benzodiazepines have been iden- Biperiden
tified as major contributors to delirium in several studies (see
Figure 41.7) (Gaudreau, Gagnon, et al., 2005b). In fact, nar- Levodopa
cotics, benzodiazepines, corticosteroids, and other psycho- Trihexyphenidyl
active agents use have been associated with 3-fold to 11-fold
increases in the prevalence of delirium (Morita et al., 2001; CAR DIOVASCULAR

Morrison et al., 2003; Gaudreau & Gagnon, 2005; Gaudreau


Amiodarone
et al., 2005b, 2007; Fong et al., 2006; Pisani, Murphy, et al.,
2009; van Munster et al., 2010). Beta-blockers
There is significant evidence linking a medication’s anticho-
Captopril
linergic potential (as measured by serum anticholinergic activ-
ity) and their incidence of causing delirium (Tune, et al., 1981, Chlorthalidone
1993; Tune, Carr, Hoag, & Cooper, 1992; Golinger al., 1987;
Digoxin
Milusheva et  al., 1990; Inouye  & Charpentier, 1996; Flacker
et al., 1998; Tune & Egeli, 1999; Tune, 2000; Kojima, Terao, & Diltiazem
Yoshimura, 1993; Meyer, Meyer,  & Kressig, 2010). Several
Dipyridamole
studies have identified some of these substances, and we have
become increasingly aware of the importance of understand- Disopyramide
ing the cumulative serum anticholinergic activity (Tune et al.,
Dopamine
1981; Tune & Folstein 1986; Tune et al., 1992, 1993; Tune &
Egeli, 1999; Tune, 2000) (see Table 41.5). Others have demon- Ergotamine
strated that, among medical inpatients, exposure to anticholin-
Furosemide
ergic agents is an independent risk factor for the development
of delirium and is specifically associated with a subsequent Hydrochlorothiazide
increase in delirium symptom severity (Han et al., 2001).
Hydralazine
Despite its apparent usefulness, there are limitations to
the use of the serum anticholinergic activity (SAA) assay in Isosorbide mononitrate
clinical practice: (1) assays of standard drug solutions do not
Lanoxin
account for pharmacokinetic differences among drugs, which
limits the interpretation of such measurements; (2)  emerg- Lidocaine
ing evidence has suggested that anticholinergic medications
Methyldopa
may not be the only cause of elevated SAA (Carnahan, Lund
et  al., 2002). For example, naturally occurring SAAs have Nifedipine
been described in the absence of exposure to known anti-
Procainamide
cholinergic pharmacological agents; their etiology, although
they are postulated to be related to the primary illness state Quinidine
(or aging process), and their clinical significance remain Triamterene
unknown (Flacker & Wei, 2001). Despite these limitations,
elevated SAA has been consistently associated with cognitive Warfarin

(contiuned)
41. D E L I R I U M • 837
Table 41.4 (CONTINUED) Table 41.4 (CONTINUED)
CENTR AL NERVOUS SYSTEM
Propantheline
Barbiturates Scopolamine
Benzodiazepines GENITOUR INARY

Carbamazepine Oxybutinin
Chloral hydrate IMMUNE ACTIVITY/IMMUNOSUPPR ESSION

Dextromethorphan
Azathioprine
Heterocyclic antidepressants (esp. amitriptyline, clomipramine,
Cyclosporine
doxepin, imipramine, trimipramine)
Interleukin-2
Lithium
Interferon
Low-potency neuroleptics (esp. chlorpromazine,
thioridazine) INFECTIOUS

Midazolam Amphotericin-B
Meprobamate Ampicillin
Monoamino oxidase inhibitors (MAOIs) Cefalothin
Phenytoin Cefamandole
Propofol Cefoxitin
SGAs (some: clozapine, olanzapine) Clindamycin
SSRIs (some: paroxetine) Cotrimazole
CORTICOSTEROIDS
Cycloserine
Betamethasone Fluoroquinolones
Cortisone Gentamicin
Dexamethasone Itraconazole
Hydrocortisone Mefloquine
Prednisolone Piperacillin
GASTROINTESTINAL
Tobramycin
Atropine Vancomycin
Belladonna MUSCLE R ELA X ANTS

Clinidum Baclofen
Dicyclomine Carisoprodol
Dimenhydrinate Cyclobenzaprine
Diphenoxylate Metaxalone
H2-antagonists (cimetidine, famotidine, Orphanedrine
nizatidine, ranitidine)
Pancuronium
Hyoscyamine
Tizanidine
Lomotil
ONCOLOGY
Loperamide
Asparaginase
Meclizine
Fluorouracil (5-FU)
Metoclopramide
Tamoxifen
Promethazine
(contiuned)
Table 41.4 (CONTINUED) Table 41.5 ANTICHOLINERGIC DRUG LEVELS IN 25
OTC MEDICATIONS R ANK ED BY THE FR EQUENCY
OF THEIR PR ESCR IPTION FOR ELDER LY PATIENTS
Cold/sinus preparations (antihistamines, pseudoephedrine)
ANTICHOLINERGIC DRUG
Sleep aids (diphenhydramine, alcohol-containing elixirs) LEVEL § (NG/ML OF ATROPINE
MEDICATION† EQUIVALENTS)
Stay-awake preparations
PAIN M ANAGEMENT 1. Furosemide 0.22

2. Digoxin 0.25
Codeine
3. Dyazide 0.08
Meperidine
4. Lanoxin 0.25
NSAIDs (esp. indomethacin and COX2-inhibitors)
5. Hydrochlorothiazide 0.00
Oxycodone
6. Propranolol 0.00
Pentazocine
7. Salicylic acid 0.00
Propoxyphene
8. Dipyridamole 0.11
Tramadol
9. Theophylline anhydrous 0.44
R ESPIR ATORY SYSTEM
10. Nitroglycerin 0.00
Theophylline
11. Insulin 0.00

12. Warfarin 0.12

impairment and delirium in a number of research settings. †= At a 10 M concentration.


-8

In recent years, new data has renewed the concern regard- § =Threshold for delirium = 0.80ng/ml.
ing the relationship between anticholinergic agents and the Source: Tune et al., 1992
neurocognitive disorders, including delirium and demen-
tia. In fact, cumulative evidence has demonstrated a strong
correlation between the use of anticholinergic agents and inci-
dent dementia (Lopez-Alvarez, Zea Sevilla et al. 2014, Gray,
80 Anderson et al. 2015, Mate, Kerr et al. 2015) and an increased
d
risk of hospitalization for confusion (Kalisch Ellett, Pratt
70 et al. 2014).
Among the main offenders, GABA-ergic medications have
60 a been increasingly implicated in the development of delirium
c (Marcantonio et al., 1994; Meuret et al., 2000; Pain, Jeltsch,
50 et  al., 2000; Wang et  al., 2000; Ely, Gautam, et  al., 2001).
Percent of Cases

Several studies have demonstrated a relationship between


40
benzodiazepine use and delirium (Tune  & Bylsma, 1991;
Marcantonio et  al., 1994; Ely, Gautam, et  al., 2001; Kudoh
30 a
c
et al., 2004). In fact, the benzodiazepine agent lorazepam has
b c been found to be an independent risk factor for daily transi-
20 f
tion to delirium (see Figure 41.8) (Pandharipande et al., 2006).
e f
A  recent study found that among ventilated burn patients,
10
e exposure to benzodiazepines was an independent risk factor
for the development of delirium (P <.001) (Agarwal, et  al.,
0
Opioids Cortico- Benzo- 2010). Similarly, the use of various general anesthetics has
steroids diazepines been associated to structural changes leading to disruption
aBreitbart et al. 200226 of blood–brain barrier (BBB) integrity (Forsberg et al., 2014).
bMorita et al. 200129 This will be discussed in further detail under the inflamma-
cTuma and DeAngelis 200030
dLawlor et al. 200014
tory hypothesis of delirium.
eOlofsson et al. 199627 There are many mechanisms by which sedative agents,
fFrancis et al. 19999 particularly GABA-ergic agents, mediate delirium. These
may include interruption of physiological melatonin and
Delirium Potentially Caused by Opioids, Corticosteroids, and
Figure 41.7 sleep patterns, interference with central cholinergic transmis-
Benzodiazepines in Six Case Series. (Gaudreau, Gagnon, et al., 2005b) sion, disruption of thalamic gating, and acute and long-term

41. D E L I R I U M • 839
1.0
5. Long term BZDP use may induce a limitation in cognitive

Probability of Transitioning
reserve capacity, which further reduces a person’s ability to
0.9
cope with early phase brain lesions by soliciting accessory
to Delirium neuronal networks (Stern, 2002).
0.8
6. New evidence suggests that BZDP use may be associated
with an increased cumulative risk of dementia (Billioti de
0.7
Gage et al., 2014).

0.6 p = 0.003 7. BZDP use interferes with central cholinergic function


muscarinic transmission at the level of the basal forebrain
0 10 20 30 40 and hippocampus (i.e., causing a centrally mediated acetyl-
Lorazepam Dose (mg) choline deficient state) (Schneck & Rupreht, 1989; Meuret,
Backman, et al., 2000; Pain et al., 2000; Wang et al., 2000);
Lorazepam and the Probability of Transitioning to Delirium
Figure 41.8
The probability of transitioning to delirium increased with the dose furthermore, BZDP activate GABAA receptors, which
of lorazepam administered in the previous 24 hours. This incremental inhibit glutamate effect on NMDA receptors, thus limiting
risk was large at low doses and plateaued at around 20 mg/day. Ach release (Cervetto & Taccola, 2008).
Abbreviations: Y-axis = delirium risk; X-axis = lorazepam dose (in mg) (Pandharipande et al., 2006)
8. BZDP use interferes with physiological sleep patterns
(e.g., ↓ slow wave sleep→ ↑REM latency→↓REM periods
duration→REM deprivation) (Borbely, Mattmann, et al., 1985;
cognitive disturb ances. Please see Box 41.2 for a comprehen- Achermann & Borbely, 1987; Bastien, LeBlanc, Carrier, &
sive list of the 12 mechanisms by which sedative and analgesic Morin, 2003, Qureshi & Lee-Chiong, 2004; Mazza et al., 2014).
agents mediate their deliriogenic effects. Similar to more clas-
sic anticholinergic agents, acute benzodiazepine use has been 9. BZDP use disrupts the circadian rhythm of melatonin
linked to agitation, confusion and delirium (Daman Willems release (Olofsson et al., 2004).
and Dillon 1986, Short, Forrest et al. 1987, Miller and Gold 10. Long-term use of BZDP increases compensatory
1991, O'Reilly and Smith 1991, Trewin, Lawrence et al. 1992, up-regulation of N-methyl D-aspartate (NMDA) and
Lechin, van der Dijs et al. 1996, Hofmann 2013); and chronic kainate receptors and Ca2+ channels (Chaudieu, St-Pierre,
use is linked to Alzheimer’s disease (Billioti de Gage, Moride Quirion, & Boksa, 1994; Heikkinen, Moykkynen, & Korpi,
et al. 2014, Kmietowicz 2014, Yaffe and Boustani 2014). 2009); enhancing NMDA-induced neuronal damage (Zhu,
Cottrell, & Kass, 1997).
11. BZDP use disrupts thalamic gating function (Gaudreau &
Box 41.2 SEDATIVE AND ANALGESIC AGENTS’ Gagnon, 2005).
DELIRIOGENIC MECHANISMS: 12. Long-term BZDP use may lead to CNS-depressant with-
drawal syndromes, upon abrupt discontinuation.
Sedative agents (mostly GABA-ergic) and opioids may contrib-
ute to the development of delirium by one of twelve possible
mechanisms:
Immobility
1. Acutely, benzodiazepines (BZDP) are associated with the
development of psychomotor retardation, cognitive blunt- Immobility—due to disease conditions, physical restraints,
ing, ataxia, poor balance, and decreased mobility, all known or medical apparatus that effectively limits the patient’s
contributing factors for delirium (Sarasin, Ghoneim, & mobility (e.g., ET tube [ventilator], IV lines, bladder cath-
Block, 1996). eters, intermittent pneumatic leg-compression devices, casts,
traction devices)— has been shown to increase the inci-
2. BZDP have been demonstrated to have a negative effect on
dence of delirium (McCusker, Verdon, Caplan, Meldon, &
memory (Lister, 1985; Ghoneim & Mewaldt, 1990; Ghoneim
Jacobs, 2002; Inouye et al., 2007). Some may include chemi-
et al., 1993).
cal restraints in this group. Data show that patients who are
3. BZDP use induces both non-amnestic and amnestic mild mobilized early (i.e., alerted and provided physical and occu-
cognitive impairment (Tannenbaum et al., 2012) pational therapy) during the course of illness demonstrate
greater return to independent functional status at hospital
4. Chronic administration of BZDP induces a down-regulation
discharge (59% vs. 35%), spend less time in the ICU (2 vs.
of BZDP-receptors (Hutchinson, Smith, & Darlington,
4  days), and experience less delirium (33% vs. 57%) com-
1996), and a reduction in the number of these receptors
pared to a control cohort (Schweickert et al., 2009).
seems to be correlated with cognitive decline (Shimohama,
Both psychoactive medications and immobility may
Taniguchi, Fujiwara, & Kameyama, 1988).
contribute to emergence delirium (ED) or emergence agi-
tation (EA), a postoperative behavior that may occur upon

840 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
awakening in patients exposed to general anesthesia (Bastron this practice and its effects, and better planning, may signifi-
& Moyers, 1967; Gutstein, 1996; Haynes, 1999; Fong et al., cantly reduce delirium’s occurrence.
2006; Maldonado, 2008b) and that has been described in
both children and adults (Albin, Bunegin, Massopust, &
Jannetta, 1974; Lepouse, et al., 2006; Abu-Shahwan, 2008;
Fronapfel, 2008; Hudek, 2009). Even though there is some N E U ROPAT HO G E N E S I S OF   DE L I R I U M
suggestion that the incidence may be increased after the
use of volatile (inhaled) anesthetic agents (Bajwa, Costi, & Many theories have been proposed to explain the etiology of
Cyna, 2010), many other agents commonly used for anesthe- delirium, but Engel and Romano explained it best:
sia and sedation in the ICU have been implicated (as already
discussed above). All forms of conventional sedation may We thus arrive at the proposition that a derangement
be associated with emergence delirium, although some are in functional metabolism underlies all instances of
better than others. As a general rule, propofol is less delirio- delirium and that this is reflected at the clinical level by
genic than midazolam, but dexmedetomidine is less delirio- the characteristic disturbance in cognitive functions.
genic than both. The use of opioids as a sedative agent may
be one of the main contributors of emergence delirium and (Engel & Romano, 1959)
should be avoided. The rule should be: Use sedative agents
for sedation and narcotics for pain management, rather than Thus we will start with the premise that delirium is a neu-
the commonly practiced narcosedation. Another contribu- robehavioral syndrome caused by the transient disruption of
tor to emergence delirium may be CNS depressants (e.g., normal neuronal activity secondary to systemic disturbances
narcotics, benzodiazepines, other GABA-ergic agents) when (Engel & Romano, 1959; Lipowski, 1992; Brown, 2000).
a patient has been exposed to these agents for a long period Over the years, a number of theories have been proposed in an
of time, and the primary team feels “it is time to extubate” attempt to explain the processes leading to the development
so the patient is quickly taken off these agents rather than of delirium (Maldonado, 2008a). Most of these theories are
being carefully weaned off them. Being more mindful of complementary, rather than competing (see Figure 41.9). It is

Precipitants of Delirium
(e.g., infection, trauma, anesthetics, surgery, hypoxia,
hypoglycemia, metabolic derrangements)

Aging/ Melatonin/Tryptophan/
Baseline Cognition Circadian Dysregulation

Neuroinflammation Oxidative Stress Neuroendocrine

Network Neurotransmitter
Disconnectivity availability

Clinical Delirium

Neuropathogenesis of Delirium Schematics of the interrelationship of current theories on the pathophysiology of delirium and how they
Figure 41.9
may relate to each other. Each proposed theory has focused on a specific mechanism or pathologic process (e.g., dopamine excess or acetylcholine
deficiency theories), observational and experiential evidence (e.g., sleep deprivation, aging), or empirical data (e.g., specific pharmacological agents’
association with post-operative delirium; intra-operative hypoxia). Most of these theories are complementary rather than competing, with many
areas of intersection and reciprocal influence. In the end, it is unlikely that any one of these theories is fully capable of explaining the etiology or
phenomenological manifestations of delirium, but rather that their interaction leads to the biochemical derangement and, ultimately, to the com-
plex cognitive and behavioral changes characteristic of delirium. (Adapted from Maldonado, 2013).

41. D E L I R I U M • 841
likely that none of these theories by itself explains the phe- O’Connor, Freund, Johnson, & Kelley, 2008; Cunningham
nomena of delirium, but rather that two or more of these, et al., 2009; Godbout & Johnson, 2009; Cerejeira, Firmino,
if not all, act together to lead to the biochemical derange- Vaz-Serra,  & Mukaetova-Ladinska, 2010; Cunningham,
ment we know as delirium. Here we will highlight four of 2011)  (Figure 41.10). Previous studies have already demon-
them. A detailed review of all has been published elsewhere strated the brain’s ability to monitor the presence of systemic
(Maldonado, 2008a, 2013). inflammatory processes (e.g., outside the BBB) and the devel-
opment of nonspecific physiological (e.g., fever, pain, malaise,
fatigue, and anorexia) and behavioral adaptations (e.g., anhe-
N EU ROI N F L A M M ATORY H Y P OT H E S I S
donia, lethargy, social withdrawal, depressed mood, cognitive
The neuroinflammatory hypothesis suggests that acute impairment) upon exposure to infection or inflammation col-
peripheral inflammatory processes (e.g., infections, surgery, lectively known as “sickness behavior.”
trauma) induce activation of brain parenchymal cells and According to the neuroinflammatory hypothesis, delir-
expression of proinflammatory cytokines and inflamma- ium represents the CNS manifestation of a systemic disease
tory mediators in the central nervous system, which in turn state that has crossed the blood–brain barrier (Maldonado,
induces neuronal and synaptic dysfunction that may serve as 2008b). Indeed, many of the circumstances associated with
the substrate for the neurobehavioral and cognitive symptoms a high occurrence of delirium (e.g., infections, postopera-
characteristic of delirium (Godbout et  al., 2005, Dantzer, tive states) can also be associated with compromise of BBB

CNS Cytokines

Hypothalamus
CRH
AVP
TRH
GnRH

LC

LH/FSH Stressors
ACTH
TSH C2
A2
Adrenal C1
Glands
Cytokines

A1
Vagus n.

T4 T3 SNS
G luc o cor tic

Gonads
oid
s

Immune
Cells
Estrogen PNS
Thymus Spleen

Lymph
Progesterone Bone Nodes
Testosterone Marrow
DHEA
Immune System
(Cells & Organs)

Schematic Illustration of Neural-Immune Interactions Immune signaling of the CNS via systemic routes and the vagus nerve (Vagus n.)
Figure 41.10
and CNS regulation of immunity via the HPA, hypothalamic–pituitary-thyroid, and hypothalamic–pituitary–gonadal axes and the sympathetic
nervous system (SNS) and parasympathetic nervous system (PNS). Cytokine expression within the CNS is represented by asterisks within the
brain. Dotted lines represent negative regulatory pathways, and solid lines represent positive regulatory pathways. CRH: corticotrophin-releasing
hormone; AVP: arginine vasopressin; TRH: thyrotropin-releasing hormone; GnRH: gonadotropin-releasing hormone; ACTH: adrenocortico-
trophin hormone; TSH: thyroid-stimulating hormone; T4: thyroxine; T3: triio-dothyronine; LH: luteinizing hormone; FSH: follicle-stimulating
hormone; LC: locus coeruleus; Al, Cl, A2, C2: brainstem adrenergic nuclei. (Marques-Deak, Cizza, & Sternberg, 2005)

842 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
integrity. It has been postulated that several illness processes IL-1 receptors that are located on perivascular macrophages
(i.e., trauma, infections) and surgical procedures may intro- and endothelial cells of brain venules. Activation of these IL-1
duce triggering factors leading to the activation of the inflam- receptors by circulating cytokines results in the local produc-
matory cascade: extensive tissue trauma, presence of foreign tion of prostaglandin E2.
organisms or substances, elevated hormone levels, blood loss The neuroinflammatory hypothesis proposes that sev-
and anemia, blood transfusions, use of extracorporeal circula- eral conditions associated with delirium are characterized
tion, hypoxia, ischemia and reperfusion, formation of hepa- by activation of the inflammatory cascade with acute release
rineprotamine complexes, and microemboli formation and of inflammatory mediators into the bloodstream (Cerejeira
migration (reviewed by Maldonado, 2008b, 2013). In addi- et  al., 2010). CNS resident cells react to the presence of
tion, other factors associated with illness and surgery may peripheral immune signals (cytokines and lipopolysaccharide
directly lead to compromise of the BBB. For example, recent [LPS]), leading to production of cytokines and other media-
data suggests that the use of various general anesthetics (e.g., tors in the brain, cell proliferation, and activation of the HPA
sevoflurane, isoflurane) can caused marked flattening of the axis (thus linking it with the neuroendocrine hypothesis, dis-
surfaces of brain vascular endothelial cells along with disrup- cussed next) through a complex system of interactions (Figure
tion of BBB-associated tight junctions at cell margins, leading 41.13). These neuroinflammatory changes induce neuronal
to holes in the vascular endothelial lining and increased BBB and synaptic dysfunction (i.e., affecting various neurotrans-
permeability, thus facilitating plasma influx into the brain mitter systems), leading to the neurobehavioral and cognitive
interstitium (Forsberg et al., 2014). The frequency and mag- symptoms characteristic of delirium.
nitude of this effect increases with age, thus potentially serv- Animal studies have shown that the administration of LPS
ing as a mechanism to mediate postoperative delirium and its induces sickness behavior, which requires activation of proin-
increased occurrence among elderly patients (Figure 41.11). flammatory cytokine signaling in the brain (Dantzer et  al.
At least four pathways through which immune signals 1998; Dantzer, 2004). Microglia are the primary recipients of
from the periphery are transduced to the brain have been peripheral inflammatory signals that reach the brain (Figure
proposed (Figure 41.12) (Dantzer et al., 2008). In the neural 41.14) (Miller, Maletic,  & Raison, 2009, Miller, Haroon,
pathway, locally produced cytokines activate primary afferent Raison,  & Felger, 2013). Activated microglia, in turn, initi-
nerves (e.g., vagus). In the humoral pathway, toll-like recep- ate an inflammatory cascade whereby release of relevant cyto-
tors on macrophage-like cells residing in the circumventric- kines, chemokines, inflammatory mediators, reactive nitrogen
ular organs and the choroid plexus respond to circulating species (RNS), and reactive oxygen species (ROS) induces
pathogen-associated molecular patterns by producing proin- mutual activation of astroglia, thereby amplifying inflam-
flammatory cytokines, which enter the brain by diffusion. In matory signals within the CNS. Cytokines, including IL-1,
the cytokine transporters pathway, proinflammatory cyto- IL-6, and TNF-alpha, as well as IFN-alpha and IFN-gamma
kines overflowing in the systemic circulation can gain access (from T cells), induce the enzyme indoleamine 2,3 dioxy-
to the brain through these saturable transport systems at the genase (IDO), which breaks down tryptophan (TRP), the
blood–brain barrier. The prostaglandin pathway involves primary precursor of 5-HT, into quinolinic acid (QUIN), a

Proposed mechanism for link between post-surgical delirium, subsequent cognitive decline
and eventual Alzheimer’s disease

Aβ42
Blood Vessel Leak Cloud Antibody
α7nAChR
Surface
Antigen

Endocytosis Intraneuronal
Aβ42 Accumulation
Cell
Death

Dendritic
Degeneration

Short-term effects Long-term effects


Disruption of brain homeostasis Chronic Abeta42 accumulation
Abnormal synaptic activity Loss of dendrites and synapses

Exposure to anesthetics causes an immediate, short-term breakdown of BBB integrity. This leads to an influx of plasma components
Figure 41.11
into the brain tissue that causes disruption of brain homeostasis and neuronal misfiring, all culminating into the array of symptoms that hallmark
delirium. Failure to completely restore BBB integrity (most common in the elderly) may trigger long-term BBB breakdown, which can drive
chronic plasma influx, more permanent disruption of brain homeostasis, and impairment of neuronal function. Chronic binding of bloodborne
brain-reactive autoantibodies and soluble amyloid peptide (Abeta42) to neurons, and their internalization via endocytosis, are essential features of
subsequent cognitive decline and early Alzheimer’s disease pathology. (Forsberg et al., 2014)

41. D E L I R I U M • 843
(a)

PAG

CEA PB
BNST

PVN NTS

SON
VLM
Vagal nerve

+ Cytokines
PAMPs

(b)

4th ventricle
Macrophage-
like cell

AP TLR

PAMP
BBB

Cytokines
SFO
Volume diffusion
OVLT
ME CP AP NTS

Pathways that Transduce Immune Signals from the Periphery to the Brain The brain and the immune system communicate through dif-
Figure 41.12
ferent pathways. (a) In the neural pathway, peripherally produced pathogen-associated molecular patterns (PAMPs) and cytokines activate primary
afferent nerves (e.g., vagal nerve, trigeminal nerves). (b) The humoral pathway involves circulating PAMPs that reach the brain at the level of the
choroid plexus (CP) and the circumventricular organs, where PAMPs induce the production and release of proinflammatory cytokines, likely reach-
ing the brain by diffusion. (Dantzer et al., 2008)

potent NMDA agonist and stimulator of glutamate (GLU) (Behan & Stone, 2000; Stone, Behan, Jones, Darlington, &
release. Multiple astrocytic functions are compromised due Smith, 2001; Darlington et al., 2007, 2010; Stone, Forrest, &
to excessive exposure to cytokines, QUIN, and RNS/ROS, Darlington, 2012).
ultimately leading to downregulation of glutamate transport- Human studies have confirmed these findings. A study of
ers, impaired glutamate reuptake, and increased glutamate adult patients admitted to inpatient medicine wards showed
release, as well as decreased production of neurotrophic fac- that those who developed delirium had significantly elevated
tors. Of note, oligodendroglia are especially sensitive to the levels (i.e., above the detection limit) of IL-6 (53% versus
CNS inflammatory cascade and suffer damage due to over- 31%) and IL-8 (45% versus 22%), compared with patients
exposure to cytokines such as TNF-alpha, which has a direct who did not develop delirium, even after adjusting for infec-
toxic effect on these cells, potentially contributing to apopto- tion, age, and cognitive impairment (de Rooij, van Munster,
sis and demyelination. The confluence of excessive astrocytic Korevaar, & Levi, 2007). This was the first study to show a
glutamate release, its inadequate reuptake by astrocytes and relationship between peripherally measured cytokine levels
oligodendroglia, activation of NMDA receptors by QUIN, and delirium as a symptom of sickness behavior in acutely
increased glutamate binding and activation of extrasynaptic admitted elderly. It also showed that cognitive function can
NMDA receptors (accessible to glutamate released from glial be impaired by a systemic infection in patients with a neu-
elements and associated with inhibition of brain-derived neu- rodegenerative disorder such as Alzheimer’s disease, and that
rotrophic factor [BDNF] expression), decline in neurotrophic this cognitive decline is preceded by raised serum levels of
support, and oxidative stress ultimately disrupt neural plas- IL-1h. Furthermore, aging and neurodegenerative disorders
ticity through excitotoxicity and apoptosis (Figure 41.15) exaggerate microglial responses following stimulation by

844 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Psychological Stress

Immunological Effects: Neuroendocrine Effects:


IL-1 IL-6 TNF-α CRH ACTH GnRH

Neurochemical Effects: Behavioral Effects:


NE 5-HT DA IL-1 Neuronal δ sleep appetite
Pathways sexcual activity

Liver
Acute phase Circulating
proteins NK Cells
NK activity

Adrenal Gland Adrenal Gland


Cortisol Cortisol
Macrophage /Lymphocyte
IL-1 IL-2
IL-6
Increase TNF-α

Decrease
Stimulatory Physical Stress
influence
Inhibitory
influence

Interlukins and the Stress Response Both psychological stress (e.g., stress, depression) and physical stress (e.g., infection, trauma, surgery)
Figure 41.13
can activate interleukin-1, which will trigger various phenomena both peripherally (e.g., cascade activation of other cytokines, induction of acute
phase proteins) and centrally (e.g., various immunologic, neurochemical, neuroendocrine, and behavioral effects). Feedback mechanisms occur at
several levels and include negative feedback exerted by cortisol. Abbreviations: IL = interleukin, TNF-α = tumor necrosis factor-α, CRH = corti-
cotropinreleasing hormone, GnRH = gonadotropin-releasing hormone, NE = norepinephrine, 5-HT = serotonin, DA = dopamine, NK = natural
killer. (Kronfol & Remick, 2000)

systemic immune stimuli such as peripheral inflammation already demonstrated how stress (induced by illness pro-
and/or infection (as described in the neural aging hypothesis). cesses, such as hypoxia/ischemia, hypoglycemia, or seizures)
can lead to excess glucocorticoid release, which can exac-
erbate cell death by inhibiting glucose transport into cells
N EU ROE N D O C R I N E H Y P OT H E S I S
(Figure 41.16 (Sapolsky & Pulsinelli, 1985; Sapolsky, 1990,
1996; Dinkel, Ogle, & Sapolsky, 2002). But there is grow-
The neuroendocrine hypothesis suggests excessive glucocor- ing evidence showing that GCs may have proinflammatory
ticoid (GC) levels may induce a vulnerable state in neurons. effects in the brain and can even enhance neuroinflamma-
Glucocorticoids are hormones released during the stress tion at multiple levels in the pathway linking LPS exposure
response that are well known for their immunosuppressive to inflammation (Figure  41.17) (Munhoz, Sorrells, Caso,
and antiinflammatory properties; however, recent advances Scavone,  & Sapolsky, 2010; Schiepers, Wichers,  & Maes,
have uncovered situations wherein they have effects in the 2005). In addition, glucocorticoids can increase proinflam-
opposite direction (Sorrells, Caso, Munhoz,  & Sapolsky, matory cell migration, cytokine production, and even tran-
2009). An extensive body of work has substantiated the idea scription factor activity in the brain (Sorrells  & Sapolsky,
that repeated or prolonged exposure to GCs has a deleteri- 2007). This can lead to a number of deleterious effects,
ous impact on brain function, and has also provided evi- including inhibition of glutamate reuptake in the synaptic
dence that GCs probably contribute to age-related decline cleft, inhibition of Ca+ efflux or sequestration, exacerbation
in brain function (Goosens & Sapolsky, 2007). Others have of the breakdown of cytoskeletal proteins (i.e., tau), increase

41. D E L I R I U M • 845
Peripheral
inflammatory
response

TRP
Perivascular
macrophage ↑ Cytokines
↑ Chemokines ↑ IDO ↓ 5ΗΤ
↑ Inflammatory
mediators
↑ QUIN
↑ RNS, ROS Microglia

Astrocyte

Postsynaptic
terminal
↑ GLU
Extrasynpatic
Oligodendrocytr ↓ Trophic NMDA receptor
support ↓ GLU Reuptake activation
↑ GLU Release

TNF-alpha

↓ BDNF
Apoptosis Excitotoxicity
Demyelination Presynaptic
terminal Glutamate

Effects of the CNS Inflammatory Cascade on Neural Plasticity. Microglia are the primary recipients of peripheral inflammatory signals
Figure 41.14
that reach the brain. When activated, microglia initiate an inflammatory cascade (e.g., releasing cytokines, chemokines, inflammatory mediators,
reactive nitrogen species, and reactive oxygen species) leading to the activation of astroglia, thus amplifying inflammatory signals within the CNS.
Cytokines (e.g., IL-1, IL-6, and TNF-alpha, as well as IFN-alpha and IFN-gamma [from T cells]), induce the enzyme indoleamine 2,3 dioxygenase,
affecting tryptophan metabolism and thus increasing quinolinic acid production, a potent NMDA agonist and stimulator of glutamate release.
Eventually, there is astrocytic function compromise leading to downregulation of glutamate transporters, impaired glutamate reuptake, and
increased glutamate release, as well as decreased production of neurotrophic factors. Oligodendroglia are especially sensitive to the CNS inflam-
matory cascade which leads to apoptosis and demyelination. The confluence of these factors leads to a decline in neurotrophic support, oxidative
stress, and, ultimately, disruption of neural plasticity through excitotoxicity and apoptosis. Abbreviations: 5-HT, serotonin; BDNF, brain-derived
neurotrophic factor; CNS, central nervous system; GLU, glutamate; IDO, indolamine 2,3 dioxygenase; IFN, interferon; IL, interleukin; NMDA,
N-methyl-D-aspartate; QUIN, quinolinic acid; RNS, reactive nitrogen species; ROS, reactive oxygen species; TNF, tumor necrosis factor; TRP,
tryptophan. (Miller, Maletic, & Raison, 2009)

in reactive oxygen species, decrease in activity of antioxidant between basal cortisol levels compared to demented, non-
enzymes, decreased release of inhibitory neurotransmitters delirious patients (Robertsson et  al., 2001). But, the most
(i.e., GABA), and decreased production of neurotrophins important finding of the study was a strong relationship
(i.e., BDNF). between delirium and DST pathology irrespective of age and
Disturbances of the HPA system have also been found in severity of dementia, with significant differences in post-DST
several studies on delirium. Studies have found that patients cortisol levels between patients with different degrees of
experiencing postoperative delirium had an impaired delirium:  the greater the intensity of delirium, the greater
stress-regulating system, with significantly elevated mean the level of nonsuppression. Similarly, delirious patients
plasma cortisol levels compared to the preoperative baseline after stroke exhibited significantly increased activation of
and nondelirious patients (McIntosh et  al., 1985). Among the HPA system compared to those without acute confu-
delirious patients triggered by lower respiratory tract infec- sion (Olsson et  al., 1989; Olsson, Marklund, Gustafson,  &
tion, 78% were found to be nonsuppressors on the DST com- Nasman, 1992; Fassbender et  al., 1994; Johansson et  al.,
pared to 14% of the patients without delirium (O’Keeffe & 1997; Slowik et al., 2002).
Devlin, 1994). Among non–medically ill demented patients, The hippocampus is a major target for these effects, with
those who develop delirium exhibited significant differences its dense concentration of glucocorticoid receptors. The

846 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
NH2 Indoleamine NH2
N'-formyl-
2,3-dioxygenase H NH kynurenine
HN O
O OH O OH
L-tryptophan
O
tryptophan
hydroxylase formamidase
OH
HO

NH2 NH2 N COOH


kynurenic acid
kynurenine (NMDA antagonist,
NH2 O
HN O OH O OH anticonvulsant &
5-hydroxy-
tryptohan neuroprotective)
aromatic amino
acid decarboxylase kyrnuenine mono-oxygenase (MO)
HO O NH
2
C COOH
OHC
NH2
HO2C OH NH2
serotonin OH 3-Hydroxykynurenine
2-amino-3-(3-oxoprop-
5-hydroxy- HN NH2 O 1-enyl)-fumaric acid
tryptamine kynureinase (KYNU)
(5-HT) N-acetyl
transferase non-enzymatic COOH
HO cyclization
NH2
O
H OH
3-Hydroxyanthranilic acid
N CH3
OH (toxic metabolite)
quinolinate
O OH
N-acetyl- HN N
5-HT COOH
5-hydroxyindole- O 2-Amino-3-carboxymuconic
OHC NH2
O-methyltransferase HOOC semialdehyde Picolinic acid
H3C O
Aminocarboxymuconate semialdehyde
O decarboxyiase
H (ACMSD) N COOH
N CH3 COOH
OH

HN O N COOH Quinolinate phosphoribosyl transferase (QPRT) NAD+


N
Quinolinic acid
melatonin niacin
(toxic metabolite)

Metabolic Pattern of Tryptophan, Serotonin, Melatonin, Kynurenic and Quinolinic Acid. Evidence suggest that some proinflammatory cytokines can not only induce
Figure 41.15
sickness behavior, but also enhance activity of the ubiquitous indoleamine 2, 3 dioxygenase (IDO), leading to deficient tryptophan (TRP) levels, thus a reduction in 5HT and
melatonin production, and a shift to the production of kynurenine (KYN) and other neurotoxic tryptophan-derived metabolites. (Maldonado 2013; modified from Darlington et al., 2010, and Stone,
Forrest, & Darlington, 2012)
Emotional
Chemical Stimuli 5-HT NA
Physical – +
(+)
(–) (+) +/–
Hypothalamus Hypothalamus

CRH
CRH NTS
(+)
(–) (+)? Pituitary P
Pituitary CNS
(–) Inflammation

ACTH

IL-1 ACTH – + IL-1 X
(+)
Adrenal TNF-α
gland
+
(+) Adrenal cortex –
Peripheral – Cytokines
GCs
(–) Inflammation
CS +/–
IL-6

Effects on heart. Immune system


(+)
blood vessels,
gonads and other Liver Inflammation, Neurotransmitters & the HPA Axis. The brain
Figure 41.17
organ systems regulates the immune system, and the immune system modulates brain
activity through various interactions. The hypothalamic–pituitary–adrenal
Neuroendocrine Circuits and the Effects of Inflammation.
Figure 41.16 (HPA) axis is the main route of immunoregulation, next to (para)sympa-
Various stressors can activate the HPA axis. The hypothalamus is thetic innervation of organs of the immune system. Immune cells secrete
stimulated to secrete CRH, which leads to ACTH secretion into the cytokines, which not only regulate the immune response but also act to
peripheral circulation. ACTH in turn triggers adrenal GC release and block the actions of other humoral substances at different levels of the HPA
production. The CRH system is inhibited by GCs in a negative feed- axis (black bars). (Schiepers, Wichers, & Maes, 2005)
back loop. TNF-κ and IL-1 are produced from inflammatory sites and
are potent activators of the HPA axis. IL-6 acts synergistically with GCs
to stimulate the hepatic secretion of acute phase proteins. Although
GCs are widely known for their anti-inflammatory actions, “(–),” more
recently also proinflammatory effects have repeatedly been reported,
OX I DAT I V E S T R E S S H Y P OT H E S I S
“(+)?” Abbreviations: HPA = hypothalamic-pituitary-adrenal; CRH =
corticotrophin-releasing hormone; ACTH = adrenocorticotrophic hor- The oxidative stress hypothesis proposes that decreased
mone; GCs = glucocorticoids; IL = interleukin; TNF = tumor necrosis
factor; (+)  =  enhancing; (–)  =  suppressing. (Dinkel, Ogle, & Sapolsky, 2002)
brain oxidative metabolism leads to abnormalities of vari-
ous neurotransmitter systems, causing cerebral dysfunction
and the behavioral symptoms of delirium. Thus, severe ill-
ness processes, combined with both decreased oxygen sup-
hippocampal–adrenal circuit may contribute to the ampli- ply and/or increased oxygen demand, may lead to the same
fication of deliriogenic factors. There is evidence that, rela- common end problem; namely, decreased oxygen availabil-
tively early during the metabolic stress leading to delirium, ity to cerebral tissue. This has been the subject of an exten-
the hippocampus begins to malfunction. This may explain sive review described elsewhere Figure 41.18 (Maldonado,
some of the memory dysfunction and errors in informa- 2013). In summary, hypoxia triggers diverse reconfigura-
tion processing, leading to the confabulations that are tions of widespread neuronal network at all levels of the
commonly seen in delirious patients. The loss of normal nervous system (i.e., molecular, cellular, synaptic, neuronal,
inhibition of adrenal steroidogenesis results in the continu- network): synaptic transmission is depressed through pre-
ous secretion of peak amounts of corticosteroids, leading to synaptic mechanisms and excitatory/inhibitory alterations
further mitochondrial dysfunction and apoptosis and fur- involving K+, Na+, and Ca 2+ channels. Cerebral ischemia
ther exacerbation of the catecholamine disturbances. A key leads to a rapid depletion of energy stores, triggering a com-
abnormality related to cortisol excess in delirium seems to plex cascade of cellular events, including cellular depolar-
be abnormal “shut-off” of the HPA axis tested by the DST. ization and Ca2+ influx, resulting in excitotoxic cell death
In experimental models, the hippocampal formation is of (see Figure 41.19).
prime importance for normal HPA axis shut-off (Olsson, Inadequate oxidative metabolism may be one of the causes
1999). In this brain area, a close interaction between of the problems observed in delirium; namely, inability to
neurotransmitters—notably ACH, 5HT, and NE—and maintain ionic gradients causing “spreading depression”;
glucocorticoid receptors is relevant for the development of abnormal neurotransmitter synthesis, metabolism and release;
delirium in elderly patients with stroke and neurodegenera- and a failure to effectively eliminate neurotoxic byprod-
tive brain diseases. ucts. Decreased oxygenation causes a failure in oxidative

848 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
A Basic Pathoetiological Model of Delirium. This figure illustrates the theorized intersections between the oxidative stress hypothesis (OSH) and the neurotransmitter hypothesis (NTH) of the
Figure 41.18
delirium, demonstrating potential common biochemical outcomes, which may explain the complex cognitive and behavioral changes characteristic of delirium. The OSH proposes that a number of physiolog-
ical processes (e.g., hypoxia, severe illness, infectious processes) may give rise to increased oxygen consumption and/or decreased oxygen availability, with associated increased energy expenditure and reduced
cerebral oxidative metabolism, leading to cerebral dysfunction and associated cognitive and behavioral symptoms of delirium. The NTH was proposed after clinical observations that delirium occurred after
the use of substances (e.g., medications, toxins) that alter neurotransmitter function or availability. The OSH intersects with the NTH as decreased oxygenation causes a failure in oxidative metabolism, lead-
ing to a failure of the ATP-ase pump system, which leads to an inability to maintain adequate ionic gradients, which in turn leads to significant electrolyte alterations (e.g., influx of Na+ and Ca2+; efflux of
K+) and subsequent alterations (e.g., excess release or decreased availability) of several neurotransmitters (e.g., GLU, DA, Ach). (Maldonado 2008b, 2013)
Decreased cerebral perfusion
Depletion in energy stores

Cerebral Ischemia

Apoptosis Excitotoxicity Inflammation (Programmed Cell Death)


Cell shrinkage Cytokine (IL-1, TNF-α) production
Chromatin aggregation Chemokines
Preservation of cell membrane EAAs (Glutamate, Aspartate), Adhesion molecules (selections, integrins, ICAM-1)
integrity Catecholamine Proteases
Preservation of mitochondria BBB breakdown
Lack of inflammation
Activation of ionophore-linked channels Activation of Metabotropic
(NMDA, AMPA) receptors

Influx of intracellular Ca2+

Protein phosphorylation Modification


Proteolysis
NO, oxygen free radical production
Altered gene expression

Neuronal Cell Death


Schematic diagram representing events leading to ischemic brain injury.
Harukuni & Bhardwaj, Neurol Clin 2006.

Mechanisms of Brain Injury after Global Cerebral Ischemia. During cerebral ischemia, excess glutamate exits into the extracellular com-
Figure 41.19
partment due to cellular depolarization, coupled with its impaired uptake, which results in increases in intracellular Ca2+. The cascade of events
responsible for glutamate excitotoxicity includes three distinct processes: (1) induction, whereby extracellular glutamate efflux is transduced by
receptors on the neuronal membrane to cause intracellular Ca2+ overload, which leads to lethal intracellular derangements; (2) amplification of the
derangement, with an increase in intensity and involvement of other neurons; and (3) expression of cell death triggered by cytotoxic cascades. Excess
release of Ca2+ and its intracellular influx is thought to be the primary trigger for a variety of complex, deleterious intracellular processes that result
from activation of catabolic enzymes such as phospholipases (which lead to cell membrane breakdown, arachidonic acid, and free radical formation)
and endonucleases (which lead to fragmentation of genomic DNA and energy failure due to mitochondrial dysfunction). (Harukuni & Bhardwaj, 2006)

metabolism, which leads to a failure of the Adenosine triphos- Hypoxia also leads to a reduced synthesis and release of
phate (ATP)-ase pump system. When the pump fails, the ionic acetylcholine, especially in the basal forebrain cholinergic
gradients cannot be maintained, leading to significant influxes centers. Indeed, cholinergic neurotransmission is particularly
of sodium (Na+) followed by calcium (Ca2+), while potas- sensitive to metabolic insults, such as diminished availability
sium (K+) moves out of the cell. Some have theorized that of glucose and oxygen. ACh synthesis requires acetyl coen-
it is the excess inward flux of Ca2+ that precipitates the most zyme A, which is a key intermediate linking the glycolytic
significant neurobehavioral disturbances observed in deliri- pathway and the citric acid cycle. Thus, reduction in cerebral
ous patients. The influx of Ca2+ during hypoxic conditions oxygen and glucose supply and deficiencies in enzyme cofac-
is associated with the dramatic release of several neurotrans- tors such as thiamine may induce delirium by impairing ACh
mitters, particularly glutamate (GLU) and dopamine (DA). production.
Glutamate further potentiates its own release as GLU stimu- There are definite data correlating poor oxygenation and
lates the influx of Ca2+, and it accumulates in the extracellular cerebral dysfunction. Studies have demonstrated a strong
space as its reuptake and metabolism in glial cells is impeded correlation between intraoperative oxygen saturation and
by the ATP-ase pump failure. In addition, at least two factors postoperative mental function (Rosenberg & Kehlet, 1993).
facilitate dramatic increases in DA:  first, the conversion of Studies have demonstrated that delirium can be induced
DA to norepinephrine (NE), which is oxygen dependent, is in healthy control subjects by dropping PaO2 to 35 mmHg
significantly decreased; second, the catechol-o-methyl trans- (Gibson  & Peterson, 1981). A  study of ICU patients dem-
ferase (COMT) enzymes, required for degradation of DA, onstrated that three measures of oxygenation (i.e., hemo-
get inhibited by toxic metabolites under hypoxic conditions, globin level, hematocrit, pulse oximetry; see Figure 41.4)
leading to even greater accumulation of DA (Graham, 1984). were worse in the patients who later developed delirium,
At the same time, serotonin (5HT) levels fall moderately in and that clinical factors associated with greater oxidative
the cortex, increase in the striatum, and remain stable in the stress (e.g., sepsis, pneumonia) occurred more frequently
brainstem (Broderick & Gibson, 1989). among those diagnosed with delirium (Seaman et al., 2006).

850 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
A  recent study demonstrated that intraoperative cerebral Several studies have demonstrated a relationship between
oxygen desaturation was a significant risk factor for post- melatonin secretion and delirium. A study of elderly patients
operative delirium in CABG patients (Figure 41.5) (Slater after major abdominal surgery had their plasma melatonin
et al., 2009). levels measured preoperatively and every 2 hours during
the postoperative period, and demonstrated that those who
did not develop delirium did not have significant changes
DI U R N A L DY S R E GU L AT ION OR M E L ATON I N
in plasma melatonin levels compared with preoperative val-
DY S R E GU L AT ION H Y P OT H E S I S
ues (Shigeta, Yasui, et al., 2001). Furthermore, patients who
The diurnal dysregulation or melatonin dysregulation experienced postoperative delirium could be broken down
hypothesis suggests that the 24-hour internal “clock” (cir- into two groups:  those who became delirious postopera-
cadian pattern) is maintained by environmental factors, tively, with an otherwise uncomplicated recovery; and those
primarily light exposure, which affect melatonin secretion who developed postoperative complications such as pneumo-
(BaHammam, 2006). Sleep deprivation may lead to the devel- nia. Patients who went on to become delirious, but without
opment of memory deficits (Walker  & Stickgold, 2004). In other complications, demonstrated reduced plasma melato-
fact, “chronic partial sleep deprivation” (i.e., sleeping limited nin levels. Those delirious but with other complications had
to 4 hours per night, for 5 consecutive nights) has been shown elevated plasma melatonin levels, similar to that seen in early
to lead into cumulative impairment in attention, critical sepsis (Mundigler et al., 2002). Similarly, a study of subjects
thinking, reaction time, and recall (Dinges, 2006). In turn, undergoing thoracic esophagectomy demonstrated a signifi-
sleep deprivation (even just 36 consecutive hours) may lead to cant correlation between ICU psychosis (i.e., delirium) and
symptoms of emotional imbalance (i.e., short temper, mood an irregular melatonin circadian rhythm (i.e., abnormally low
swings, and excessive emotional response). Cumulative sleep serum levels of melatonin), measured every 6 hours during the
debt can cause delirium in itself, and can aggravate or perpet- first 4 postoperative days (P =.0001) (Miyazaki et al., 2003).
uate delirium (Ito, Harada, Hayashida, Ishino, & Nakayama, Of note, some have observed a relationship between the
2006; Pandharipande  & Ely, 2006). Similarly, others have motoric delirium subtype and melatonin levels. A  study of
found sleep deprivation to consistently precede onset of delir- hospitalized elderly medical patients evaluated them daily
ium in cardiac surgical patients (Sveinsson, 1975), and that using the Delerium Rating Scale (DRS) and conducted uri-
ICU patients with sleep deprivation were significantly more nary measures of their 6-sulphatoxymelatonin (6-SMT), the
likely to develop delirium than patients without sleep depriva- chief metabolite of melatonin and a reliable proxy for serum
tion (Helton, Gordon, & Nunnery, 1980). melatonin (Matthews, Guerin,  & Wang, 1991). The mean
Melatonin plays important roles in multiple bodily func- and standard deviation values of 6-SMT for each of the three
tions (Reiter, 1991; Brzezinski, 1997; Verster, 2009). It has a motoric subtypes were calculated, and it was found that dur-
chronobiotic effect (affecting aspects of biological time struc- ing periods of hyperactive delirium, subjects had decreased
ture), sleep–wake cycle regulatory effects, and helps reset urinary 6-SMT levels compared with when they had recov-
circadian rhythm disturbances. It has extensive antioxidant ered. Patients with hypoactive delirium had raised 6-SMT
activity (with a particular role in the protection of nuclear levels compared with when they had recovered, while those
and mitochondrial DNA), extensive anti-inflammatory activ- with the mixed form of delirium had no difference in urinary
ity, and some anti-nociceptive and analgesic effects; and data metabolite concentrations (Balan et al., 2003).
suggest that melatonin receptors appear to be important in
mechanisms of learning and memory. Melatonin also inhibits
N EU RON A L AG I NG H Y P OT H E S I S
the aggregation of the amyloid beta protein into neurotoxic
micro-aggregates responsible for the neurofibrillary tangles Finally, the neuronal aging hypothesis suggests that elderly
characteristic of Alzheimer’s disease, and it prevents the patients are at increased risk for developing delirium, based
hyperphosphorylation of the tau protein. All of these factors on the fact that aging is associated with age-related cerebral
may have potential implications regarding the development of changes in stress-regulating neurotransmitter and intracel-
delirium in the medically ill and postoperative patient. lular signal transduction systems, and observations suggest-
In addition, current evidence suggests that acute and ing that chronic neurodegeneration is accompanied by an
chronic sleep deprivation is associated with decreased propor- inflammatory response characterized by a chronic, but selec-
tions of natural killer cells, lower antibody titers following tive, activation of CNS microglial cells that are “primed” to
influenza-virus immunization, reduced lymphokine-activated produce exaggerated inflammatory responses to immuno-
killer activity, and reduced IL-2 production. Conversely, cyto- logical challenges (Figure 41.20) (Cunningham et al., 2005,
kines may play a role in normal sleep regulation by increasing 2009). This inflammation is assumed to contribute to disease
non-REM sleep and decreasing REM sleep, and, during inflam- progression through the production of inflammatory medi-
matory events, an increase in cytokine levels may intensify their ators. Data suggest that the aging process is associated with
effects on sleep regulation. Moreover, sleep deprivation may a twofold to fourfold increase in baseline levels of circulat-
alter endocrine and metabolic functions, altering the normal ing inflammatory mediators, including cytokines and acute
pattern of cortisol release and contributing to impairment of phase proteins (Rosczyk, Sparkman,  & Johnson, 2008).
the regulation of glucocorticoid secretion via negative feedback This may help explain why even minor surgical trauma was
mechanisms, insulin resistance, and glucose intolerance. associated with increased IL-1b hippocampal expression in

41. D E L I R I U M • 851
(a) Periphery: acute insult Vulnerable brain: aging/dementia

Systemic Endothelial cells of


Cerebral vasculature Primed
Inflammatory Tissue Macrophages Microglia
Event
Infection, Injury
Surgery IL-1β
TNF-α
PAMPS INDIRECT IL-1β
IL-6
IL-1β TNF-α
IFN α/β
PGE2 PGE2
PGE2
TNF-α NO
IFN α/β DIRECT
GCs

Neuronal Neuronal
(b) Systemic inflammatory events Neuronal
Dysfunction Damage Death
IL-1RI
70
TNFp55R
60
GR EP1-4
50
Functional decline

Reversibly impaired
40 neuronal function Reversible Irreversible
30

20 disease+SIEs
disease w/o SIEs
10
normal SIEs
LONG-TERM
0 DELIRIUM COGNITIVE
12

13

14

y
15

16

y
17

18

y
19

DECLINE
er

er

er
SI

SI

SI
ov

ov

ov
c

c
re

re

re

Disease progression (weeks)

Effect of Inflammation on the Aging Brain. (a) Systemic inflammatory events trigger the release of inflammatory mediators by tissue mac-
Figure 41.20
rophages and brain vascular endothelial cells. These mediators may affect neuronal function directly, or via the activation of microglial cells that have
become primed by neurodegenerative disease or aging. Inflammatory mediators may cause reversible disruption of neuronal function, leading to various
neurobehavioral syndromes including delirium. These mediators may also induce acute neuronal synaptic or dendritic damage that may be reversible
and contribute to delirium, or may be irreversible and contribute to long-term cognitive decline. (b) Successive systemic inflammatory insults induce
acute dysfunction, which is progressively less reversible each time, and also contribute to the progression of permanent disability. Abbreviations: IL-1RI,
interleukin 1 receptor type I; TNFp55, TNFp55 receptor; GCs, glucocorticoids; GR, glucocorticoid receptor; NO, nitric oxide; EP1–4, prostaglandin
receptors 1–4; PAMPs: pathogen-associated molecular patterns; IFNα/β, interferon α/β; SIEs, systemic inflammatory events.

animal studies. Aging is also associated with decreased vol- published meta-analysis on the matter found a positive asso-
ume of ACh-producing cells and decreased cerebral oxida- ciation between POCD and the APOE sigma4 allele.
tive metabolism. Thus, the decline in cognitive functioning In the end, any or all of the above theories lead to the same
associated with the normal aging process is aggravated by common result: changes in neurotransmitter concentration or
the presence of even mild hypoxia, which further inhibits receptor sensitivity (due to any of the mechanisms discussed
ACh synthesis and its release. As previously discussed, the by the previous hypotheses) mediate the different symptoms
probability of experiencing delirium rises with age. Of note, and clinical presentations of delirium. In general, the most
some have suggested that prostaglandins are responsible for commonly described neurotransmitter changes associated
the LPS-induced inhibition of sickness behaviors, either with delirium are excess release of norepinephrine (NE),
independent of cytokine activity or as complement (i.e., sen- dopamine (DA), and/or glutamate (GLU), and increased
sitizing certain targets to the effect of cytokines) (Teeling Ca+ channel activity (Ca+ Ch); reduced availability of ace-
et al., 2007). tylcholine ( Ach) and/or melatonin (MEL); and either
There has been some controversy about the contribution of a decreased and increased activity in serotonin (5HT),
apolipoprotein E and the development of postoperative cogni- histamine (H1&2), and/or gamma-amino butyric acid
tive dysfunction (POCD). To date, there have been nine pub- (GABA), probably depending on the etiology or motoric
lished papers on the topic; four of them found that APOE-E4 presentation (see Table 41.7).
genotype was not associated with POCD (Abildstrom et al.,
2004; Tagarakis et al., 2007; van Munster, Korevaar, de Rooij,
Levi,  & Zwinderman, 2007; Bryson et  al., 2011), while the C L I N IC A L PR E S E N TAT ION OF   DE L I R I U M
other five suggested a positive relationship between POCD
and APOE genotype (Adamis et  al., 2007; Ely et  al., 2007; Delirium is an acute or subacute organic mental syndrome
Leung et al., 2007; Zhang et al., 2008; van Munster, Korevaar, characterized by disturbance in attention (often evidenced
Zwinderman, Leeflang, & de Rooij, 2009). (For a summary by a reduced ability to direct, focus, sustain, and shift atten-
of all published studies see Table 41.6.) Of note, the only tion) and awareness (with impaired orientation to the

852 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Table 41.6 ASSOCIATION BET W EEN APOLIPOPROTEIN E GENOTYPE AND POSTOPER ATIVE COGNITIVE
DYSFUNCTION

STUDY(N = 10) POPULATION FINDINGS P-VALUE

Abildstrom et al., 2004 Patients aged 40 yrs. • One week after surgery, the incidence of POCD was 11.7% P = 0.41
Prospective, n = 967 and older undergoing in patients with the epsilon 4 allele and 9.9% in patients
non-cardiac surgery without the epsilon4 allele.
• Conclusion: unable to show a significant association
between apolipoprotein E genotype and POCD.

Adamis et al., 2007 ≥70 y/o hospitalized • Recovery was significantly (P <.05) associated with lack of p < 0.05
Prospective, n = 164 medically ill patients APOE 4 allele and higher initial interferon (IFN)-gamma.
• A model incorporating gender, APOE epsilon4 status and
IGF-I levels predicted recovery or not from delirium in
76.5% of cases, with a sensitivity 0.77 and specificity 0.75.
• It further found a positive relationship between delirium
with APOE genotype, IFN-gamma, and IGF-I, but not
with IL-6, IL-1, TNF-alpha, and leukemia inhibitory
factor.

Leung et al., 2007 nested cohort study; • The presence of one copy of the e4 allele was associated P = 0.005
Prospective, n = 190 patients aged ≥ 65 yr, with an increased risk of early postoperative delirium
scheduled to undergo (28.3% vs. 11.1%; P = 0.005). Even after adjusting for
major noncardiac surgery covariates, patients with one copy of the e4 allele were still
requiring anesthesia. more likely to have an increased risk of early postoperative
delirium (OR, 3.64; 95% CI, 1.51–8.77) compared with
those without the e4 allele.

Tagarakis et al., 2007, Elderly adults undergoing • Study confirmed the high incidence of cognitive decline NS
n = 137 cardiac bypass surgery; and delirium after coronary surgery, but it does not
excluded dementia support the role of the APOE-epsilon 4 allele in the
occurrence of delirium.

Van Munster et al., 2007 Acutely admitted patients The OR for carriers of an APOE epsilon 4-allele compared P = 0.12
n = 415 to the Department of with patients without an APOE epsilon 4-allele for
Medicine of 65 yrs developing delirium was 1.17 (95% CI: 0.49–2.78) in the
and over cognitively intact patients and 0.42 (95% CI: 0.14–1.30)
in the cognitively impaired patients. No relation existed
between the total number of APOE epsilon 4-alleles and the
different delirium subtypes (P = 0.12).

Ely et al., 2007 Mechanically ventilated Using multivariable regression analysis to adjust for age, p = .005
n = 53 ICU patients admission diagnosis of sepsis or acute respiratory distress
syndrome or pneumonia, severity of illness, and duration
of coma, the presence of APOE4 allele was the strongest
predictor of delirium duration (OR, 7.32; 95% CI,
1.82–29.51, p =.005).

Zhang et al., 2008 Elderly patients (>60 yo) The presence of the e4 allele and low level of education were P = 0.01
n = 196 scheduled for major both associated with an increased risk of EA (36.9% vs.
abdominal surgery 15.8%, P = 0.005; 30% vs. 14.3%, P = 0.01). After adjustment
requiring general for covariates, the patients with the copy of the e4 allele were
anesthesia shown to have a greater likeliness of an increased risk of EA
(OR: 4.32; 95% CI: 1.75–10.05).

Van Munster et al., 2009 Medical department and The OR for delirium adjusted for age, cognitive, and P = 0.04
n = 656 + orthopedic/traumatology functional impairment of sigma4 carriers compared with
Meta-analysis department of university non-sigma4 carriers was 1.7 (95% CI: 1.1–2.6). Four studies
hospital from 2003 to 2007. were added to the meta-analysis, which included 1,099
patients in total. The OR for delirium in the meta-analysis
was 1.6 (95% CI: 0.9–2.7) of sigma4 carriers compared with
non-sigma4 carriers.
This study and meta-analysis suggest an association between
delirium and the APOE sigma4 allele.
Bryson et al., 2011 Patients ≥60 yrs. of age Delirium predicted POCD at discharge (OR 2.86; 95% CIs p = 0.625
n = 88 undergoing open aortic 0.99 to 8.27) but not at three months. Apolipoprotein E-ε4
repair. genotype was not associated with either delirium or POCD
following adjustment for covariates.
Table 41.7 THEOR IZED NEUROCHEMICAL MECHANISMS ASSOCIATED WITH CONDITIONS LEADING TO DELIR IUM

HPA NMDA CHANGES


DELIR IUMSOURCE ACH DA GLU GABA 5HT NE TR P PHE HIS CYTOK A XIS ACTIVITY IN R BF EEG MEL INFLA M CORT

Anoxia/hypoxia         , ╬ ╬  ╬    

Aging          ╬ ╬  ╬    

TBI          ╬      ╬ 

CVA          ╬   ╬   ╬ 

Hepatic failure          ╬ ╬  ╬    
(encephalopathy)

Sleep deprivation   ╬        ╬    ╬ ╬ 

Trauma, Sx,             ╬    
& Post-op

ETOH and           ╬      
CNSdepressants
withdrawal

Infection/sepsis           ╬ ╬ ╬    

Dehydration and       ? ?   ╬   ╬  ╬ 
electrolyte imbalance
Medical illness    ╬         ╬ ╬  ╬ 
Legend & Abbreviations:  = likely to be increased or activated;  = likely to be decreased or slowed;  = no significant changes; (╬) = probably a contributor, exact mechanism is unclear; (−) = likely not to
be a contributing factor; CVA = cerebro-vascular accident; Sx = surgery; ETOH = alcohol; CNS-Dep = central nervous system depressant agent; ACH = acetylcholine; DA = dopamine; GLU = glutamate;
GABA = gamma-amino butyric acid; 5HT = 5-hydroxytryptamine or serotonin; NE = norepinephrine; Trp = tryptophan; Phe = phenylalanine; His = histamine; Cytok = cytokines; HPA axis = hypot
halamic-pituitary-adrenocortical axis; NMDA = N-methyl-D-aspartic acid; RBF = regional blood flow; EEG = electroencephalograph; Mel = melatonin; Inflam = inflammation; Cort = cortisol.
Source: Adapted from Maldonado, J. R. (2013). Neuropathogenesis of delirium: Review of current etiologic theories and common pathways. American Journal of Geriatric Psychiatry, 21, 1190–1222.
environment; Criterion A; APA, 1994). An additional dis- course is an important observation and may even aid in the
turbance is present in cognition (e.g., memory deficit, disori- diagnosis of delirium.
entation), language, visuospatial ability, or perception (e.g., As described above, global disturbance of cognition is
hallucinations or delusions; Criterion C; APA, 1994). Often, a cardinal feature of delirium. All “main cognitive func-
the syndrome is accompanied by a disordered sleep–wake tions” (e.g., perception, thinking, memory) are impaired or
cycle, (e.g., sleep–wake cycle reversal, sundowning) and alter- abnormal to some extent. In addition, delirious patients have
ations in psychomotor activity (e.g., decreased or increased difficulty sustaining attention and responding to stimuli
depending on the type of delirium; APA, 1994). By defini- selectively. Usually, patients are easily distracted by their sur-
tion, these changes develop over a relatively short period of rounding environment. Delirium may be better understood
time (i.e., hours to days) and represent a change from baseline as a disorder of awareness, not arousal. Yet, arousal problems
attention and awareness, and have a tendency to fluctuate in may be associated with the various presentations of delirium.
severity during the course of a day (e.g., waxing and waning; For example, in the case of hypoactive delirium, the activating
Criterion B; APA, 1994). Similarly, the presentation cannot system may be hypoactive, in which case the patient would
be better explained by a preexisting psychiatric disorder (e.g., appear apathetic, somnolent, and quietly confused. In other
schizophrenia or dementia) and are deemed to be caused as patients, the brainstem’s activating system may be hyperac-
a result of a primary medical disorder and/or its treatment, tive, in which case the patient is agitated and hypervigilant,
or mediated by psychoactive substances (e.g., acute use or exhibiting psychomotor hyperactivity and agitation, leading
intoxication, or substance withdrawal, as in delirium tremens; to the hyperactive type of delirium.
Criteria D & E; APA, 1994). Often, the sleep–wake cycle is disturbed. This is not only
Some have suggested there are three core domains of symptomatic of delirium, but in itself may either cause or
delirium: cognitive deficits (characterized by attention and vig- exacerbate the condition due to sleep deprivation. More com-
ilance disturbances); circadian rhythm dysregulation (charac- monly, patients exhibit sleep–awake cycle reversal: night sleep
terized by fragmentation of the sleep–wake cycle); and higher is usually shortened and fragmented; while daytime wakeful-
cortical dysfunction (characterized by semantic and compre- ness is often reduced, presenting in the form of drowsiness
hension deficits, executive dysfunction, and thought process and the tendency to nap or being difficult to arouse during
disturbances) (Meagher et al., 2007). In fact, sleep–wake cycle the day.
disturbance, inattention, and language/thought process/com- Problems with orientation are likely to be associated
prehension abnormalities were the most frequent, consistent, with the patients’ inability to adequately process incom-
and differentiating symptoms of delirium in medical/surgi- ing information and their impaired memory formation. In
cal populations (Trzepacz et al., 2001; Meagher et al., 2007; fact, delirious patients seem to experience a reduced ability
Franco, Trzepacz, Mejia, & Ochoa, 2009). to discriminate and integrate perceptions. This may include
The clinical features of delirium include a prodromal a reduced ability to relate incoming stimuli meaningfully to
phase, usually marked by restlessness, anxiety, irritability; previously acquired knowledge. Thus, often patients have a
and the experiencing of nightmares, transient hallucina- difficult time telling apart perceptions, images, dreams, illu-
tions (often hypnagogic), and sleep disturbances, which usu- sions, and hallucinations. This may be the reason why patients
ally develop over a period of several hours to days before the commonly report their experiences as happening “as if dream-
patient exhibits the full-blown syndrome. Often patients ing while still awake.” Often patients experience emotional
recover from a comatose or medication-induced state (e.g., dysregulation or lability.
midazolam or propofol infusion) in a confused and agitated Delirious patients’ memory is often impaired in most
state (i.e., emergence delirium). aspects (i.e., registration, retention, and recall). They may
Even though delirium was originally defined as a transient exhibit various degrees of both anterograde and retrograde
syndrome, “chronic” forms may be seen in patients experienc- amnesia, though their remote memory is often spared.
ing protracted medical problems (e.g., chronic infections), Delirious patients often exhibit a disorganized and frag-
those with concomitant psychiatric disorders (e.g., bipolar mented thinking pattern. The disturbances range from an
disease, psychosis), those with baseline cognitive impairment inability to direct thoughts at will, to a complete inability to
(e.g., delirium superimposed on dementia, traumatic brain think coherently. This includes an inability to reason, solve
injury [TBI]), and patients with combined delirium and problems, anticipate the consequences of actions, and grasp
CNS-depressant withdrawal (e.g., delirium tremens, opioid the meaning of abstract words. This disorganization of think-
or benzodiazepine withdrawal) (see “Impact of Delirium” ing, coupled with memory disturbance and misinterpreta-
section). More commonly, patients exhibit a rapid and fluctu- tion of environmental stimuli, often leads to the development
ating course with symptoms varying rapidly over time, giving of delusional thinking (usually of paranoid nature) and
rise to the classic waxing and waning picture. Patients often confabulation.
experience the appearance or exacerbation of behavioral dis-
turbances in the evening and at night, leading to the classic
DE L I R I U M S U B T Y PE S
“sundowning” syndrome (Cardinali, Brusco, Liberczuk,  &
Furio, 2002). Whether this represents a function of biologi- Historically, the first delirium subtypes have been based on
cal circadian rhythm or results from decreased sensory input behavioral and motoric characteristics (Liptzin  & Levkoff,
is unclear. The appearance of lucid intervals in the clinical 1992). Factor analysis has confirmed the existence of at least

41. D E L I R I U M • 855
two different clusters of symptoms:  hyperalert/hyperactive experienced the worst prognosis (e.g., prolonged hospi-
features (e.g., agitation, hyperreactivity, aggressiveness, hallu- tal stay, higher morbidity and mortality) (Kiely, Jones,
cinations, delusions); and the hypoalert/hypoactive features Bergmann, & Marcantonio, 2007). Frequently these patients
(e.g., decreased reactivity, motor and speech retardation, facial present with depressive-like symptoms (e.g., unawareness of
inexpressiveness) (Camus et al., 2000). the environment, lethargy, apathy, decreased level of alert-
Subsequent studies have suggested that there are at least ness, psychomotor retardation, decreased speech produc-
three types of delirium, based on their clinical (motoric) man- tion, and episodes of unresponsiveness or staring). Patients
ifestations:  hyperactive, hypoactive, and mixed (Meagher, with hypoactive delirium often endorsed depressive symp-
O’Hanlon, O'Mahony, Casey, & Trzepacz, 2000; Meagher & toms, such as low mood (60%), worthlessness (68%), and
Trzepacz, 2000). It is important to recognize that delirium frequent thoughts of death (52%) (Farrell  & Ganzini,
seems to manifest differently in different populations under 1995). Studies have demonstrated that a large percentage of
study (see Figure 41.21). For example, studies in the general these patients are inappropriately diagnosed and treated as
adult medical population revealed that, among referrals to depressed. Hypoactive delirium is unrecognized (and there-
a consultation psychiatry service, the most common type fore untreated or mismanaged) in 66%–84% of patients. In
appeared to the mixed type (46%), followed by the hyper- fact, studies have found that patients with hypoactive delir-
active (30%) and the hypoactive (24%) among those meet- ium demonstrated a seven-fold risk of under-recognition by
ing International Classification of Diseases–10 (ICD-10) nurses and that older, visually impaired, demented patients
(Box 41.3) criteria for delirium (Meagher et al., 2000). Yet, the with hypoactive delirium were 20 times less likely to be rec-
numbers seem quite different when we look at the critically ognized by nurses as having delirium (Inouye, Foreman,
ill populations. Among medical ICU (MICU) patients, the Mion, Katz,  & Cooney, 2001). Our own experience at
most common type continues to be the mixed type (54.9%), Stanford University Hospital, in parallel to that of others
but there is a significant reversal on the remaining types, with (Farrell  & Ganzini, 1995; Kishi et  al., 2007), found that
a much larger percentage of hypoactive (43.5%) than hyperac- 42% of the time when the psychiatry consultation service
tive (1.6%) types (Peterson et al., 2006). Meanwhile, among was called to treat a patient for “depression,” the patient’s
surgical ICU (SICU) patients, there is a significant reversal correct diagnosis was hypoactive delirium (Maldonado,
on the types of delirium found, with the most common form Dhami,  & Wise, 2003). The same study found that nearly
of delirium being the hypoactive type (64%), followed by the 80% of these patients had been inappropriately prescribed
mixed type (27%), and only a minority of subjects exhibiting antidepressant medications.
the purely hyperactive (9%) type (Pandharipande, Cotton To most physicians, the clearest and most recognizable
et al., 2007). Finally, a study of elderly patients in the medicine form is the hyperactive type. Most clinicians agree that a
wards found that the significant majority of subjects experi- confused, disoriented patient who does not have a preexist-
enced the hypoactive subtype (65%), compared to the hyper- ing psychiatric diagnosis, who suddenly becomes agitated,
active (25%) or mixed (10%) subtypes (Khurana et al., 2011). combative, and/or assaultive, is probably suffering from the
What is most significant about these findings is not hyperactive or “agitated type” of delirium. We use the term
only that hypoactive delirium is often missed or misdiag- “mixed type” to describe the classic “waxing and waning”
nosed (Liptzin  & Levkoff, 1992), but that these patients pattern, commonly seen in medically ill patients who appear

Delirium Subtypes

Hyperactive (3 or more)2
• Hypervigilance
• Restlessness
• Fast/loud speech Hypoactive (4 or more)2
General Poputation1
• Anger/irritability ☐ Unawareness
• Combativeness ☐ Lethargy
• Impatience 46% ☐ Decreased alertness
• Uncooperative ☐ Staring
• Laughing Hyperactive24%
☐ Sparse/slow speech
• Swearing/singing 30% ☐ Apathy
• Euphoria ☐ Decreased motor
• Wandering activity
• Easy startling
• Distractibility Hypoactive Hyperactive Mixed
• Nightmares
• Persistent thoughts
34% 25% 10%
2%
54%
44% 64% 65%

2% MICU3 SICU4 Hospitalized Elderly5

Figure 41.21 Motoric subtype of delirium depending on population under study.

856 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Box 41.3 ICD-10 DIAGNOSTIC CRITERIA FOR DELIRIUM

For a definite diagnosis, delirium symptoms, mild or severe, should be present in each one of the following areas:
1. Impairment of consciousness and attention, on a continuum from clouding to coma (e.g., reduced ability to direct, focus, sustain, and
shift attention);
2. Global disturbance of cognition (e.g., perceptual distortions, illusions and hallucinations—most often visual; impairment of abstract
thinking and comprehension, with or without transient delusions, but typically with some degree of incoherence; impairment of
immediate recall and of recent memory but with relatively intact remote memory; disorientation for time as well as, in more severe
cases, for place and person);
3. Psychomotor disturbances (e.g., hypo- or hyperactivity and unpredictable shifts from one to the other; increased reaction time;
increased or decreased flow of speech; enhanced startle reaction);
4. Disturbance of the sleep-wake cycle (e.g., insomnia or, in severe cases, total sleep loss or reversal of the sleep-wake cycle; daytime
drowsiness; nocturnal worsening of symptoms; disturbing dreams or nightmares, which may continue as hallucinations after
awakening);
5. Emotional disturbance (e.g. depression, anxiety or fear, irritability, euphoria, apathy, or wondering perplexity).
The onset is usually rapid, the course diurnally fluctuating, and the total duration of the condition less than six months. The above
clinical picture is so characteristic that a fairly confident diagnosis of delirium can be made even if the underlying cause is not clearly
established. In addition to a history of an underlying physical or brain disease, evidence of cerebral dysfunction (e.g., an abnormal elec-
troencephalogram, usually but not invariably showing a slowing of the background activity) may be required if the diagnosis is in doubt.

agitated and combative at times, with alternating episodes of P = 0.002); conversely, patients with no delirium were more
somnolence and hypoactivity. likely to be discharged home and less likely to need convales-
In addition to motoric subtypes, researchers have recently cence or long-term care than those with SSD (see Table 41.8)
described a new delirium presentation form:  subsyndromal (Ouimet et al., 2007).
delirium (SSD). Different from the other three described
above, this is not differentiated by motor type but by an
incomplete presentation of diagnostic criteria, along with DI AG N O S I N G DE L I R I U M
cognitive impairment. In one study, patients with SSD
(defined as meeting ≥ 2 DSM-delirium criteria, but did not Several studies have demonstrated that hospital staff in gen-
meet Confusion Assessment Method [CAM] diagnostic eral, and physicians in particular, are not good at identifying
criteria for delirium) had outcomes similar to or worse than delirium. In fact, studies have demonstrated that delirium is
those with mild CAM-defined delirium (e.g., nursing home misdiagnosed, detected late, or missed in over 50% of cases
placement or death at 6 months: 27% vs. 0%) (Marcantonio, across the various healthcare settings (Kean & Ryan, 2008).
Ta, Duthie,  & Resnick, 2002). Others have demonstrated Despite its high prevalence in critically ill patients, delirium
that when the appropriate screening or diagnostic test is used remains unrecognized in as many as 66%–84% of patients
to allow differentiation between no delirium, subsyndromal experiencing this complication (Francis et al., 1990; Inouye,
delirium, and clinical delirium (e.g., severity scale), patients 1994; Rolfson, McElhaney, et al., 1999; Ely, Margolin et al.,
with SSD in the general medical wards experienced longer 2001; Inouye et al., 2001; Pisani, Redlich et al., 2003; Pisani,
acute care hospital stays, increased post-discharge mortal- Inouye et al., 2003; Pandharipande et al., 2007; Steis & Fick,
ity, more symptoms of delirium, and a lower cognitive and 2008; Swigart, Kishi, Thurber, Kathol,  & Meller, 2008;
functional level at follow-up than patients with no SSD, even Steis, Shaughnessy,  & Gordon, 2012). Unfortunately, the
after adjusting for illness severity, baseline cognitive status, associated mental status changes are often misattributed
and severity of baseline functional status (Cole et al., 2003). to dementia, depression, or just an expected occurrence of
In addition, patients with SSD had the same set of risk fac- medical illness. Several studies have demonstrated a high
tors that predict the likelihood of developing DSM-defined rate (45%–66%) of missed diagnosis (i.e., either missed com-
delirium (e.g., older age, baseline dementia, and greater physi- pletely or misattributed to another disorder) of delirium by
ological severity of illness). A recent ICU study similarly con- general medicine and surgical services (Farrell  & Ganzini,
cluded that patients with SSD (which defined SSD as patients 1995; Armstrong, Cozza,  & Watanabe, 1997; Kishi et  al.,
who have <4 points on the Intensive Care Delirium Screening 2007). A recent study of an adult, acute hospital population
Checklist [ICDSC]) (Bergeron, Dubois, Dumont, Dial,  & found that only 63.6% of patients with delirium were recog-
Skrobik, 2001) experienced greater mortality (10.6% vs. 2.4%; nized by nursing staff to be confused or delirious, whereas
P < 0.001), longer ICU length of stay (5.2 days vs. 2.5 days; P < 43.6% had confusion documented in the medical case notes
0.001), and longer overall hospital stay (40.9 days vs. 31.7 days; (Ryan et al., 2013). The authors concluded that “even if some

41. D E L I R I U M • 857
Table 41.8 SUBSY NDROMAL DELIR IUM AND CLINICAL OUTCOMES

NO DELIR IUM (ND) SUBSY NDROM AL (SD) CLINICAL(CD) P VALUE*

ICU Mortality 2.4% 10.6% 15.9% P < 0.001

ICU LOS 2.5 d 5.2 d 10.8 d P < 0.001

Hospital LOS 31.7 d 40.9 d 36.4 d ND vs. SD, P = 0.002


ND vs. CD, P < 0.001
SD vs. CD, P = 0.137
Severity of illness 12.9 16.7 18.6 ND vs. SD, P < 0.001
(APACHE II) ND vs. CD, P < 0.001
SD vs. CD, P < 0.016
*Pair-wise comparison
Source: Ouimet et al., 2007

recognized, but undocumented, delirium cases occurred, or delirium are a “normal occurrence” in certain medical
these findings suggest that delirium is not a high diagnostic settings, such as the ICU) (Wong, Chiu, & Chu, 2005;
and therapeutic priority, despite its treatability and relevance Wong, Holroyd-Leduc, et al., 2010; Adams, 1988; Inouye,
to outcomes, especially for poorer prognosis in the elderly” 1991; Ely, Siegel, & Inouye 2001; Breitbart et al., 2002;
(Ryan et al., 2013). Rabinowitz, 2002; Carnes, Howell, et al., 2003; Cole
Factors contributing to the poor detection rate of delirium et al., 2003; Pisani et al., 2003; Dantzer, Capuron, et al.,
can be divided into three large components, which have sig- 2008; Steis & Fick, 2008; Davis & MacLullich, 2009;
nificant interplay (Table 41.9): Khan et al., 2009; Collins et al., 2010; Boot, 2012; Steis
et al., 2012); and
1. Patient factors (e.g., older subjects, patients experiencing
3. Systems factors (e.g., lack of consensus over the optimal
comorbid dementia, fluctuating course of presentation,
assessment of delirium, location of care [worse in surgical
presence of hypoactive features) (Liptzin & Levkoff,
than in medical settings], busy clinical settings [espe-
1992; Inouye et al., 1993; Farrell & Ganzini, 1995;
cially low nurse-to-patient ratio], inadequate application
Inouye, 1999; Inouye et al., 2001; Pisani et al., 2003;
of “sedation holidays” in sedated-ventilated patients, and
Peterson et al., 2006; Kiely et al., 2007; Lowery et al.,
the rapid transfer of patients from one unit to another,
2007; McAlpine et al., 2008; Meagher et al., 2008;
which may decrease the proper documentation and
Givens, Jones, & Inouye, 2009; Leonard et al., 2009;
diagnosis) (Inouye, 1991, 1998; Inouye, Schlesinger, &
Collins, Blanchard, Tookman, & Sampson, 2010;
Lydon, 1999; Shinn & Maldonado, 2000; Inouye
Khurana et al., 2011; Srinonprasert et al., 2011; Morandi,
et al., 2001; Pisani, Redlich, McNicoll, Ely, & Inouye,
McCurley, et al., 2012; Steis & Fick, 2012; Ryan
2003;(Pisani, Redlich et al. 2003) Siddiqi & House,
et al., 2013);
2006; Bickel et al., 2008; Davis & MacLullich, 2009;
2. Clinician/practitioner factors (e.g., lack of knowledge and Collins et al., 2010; NICE, 2010; Wong et al., 2010;
training, lack of confidence, lack of suspicion, lack of time Boot, 2012; Brummel, Morandi, et al., 2012; Neto,
of the clinical staff, expectation that altered mental status Nassar, et al., 2012; Steis et al., 2012).

Table 41.9 FACTORS CONTR IBUTING TO THE POOR DETECTION R ATE OF DELIR IUM

PATIENT FACTORS CLINICI AN/PR ACTITIONER FACTOR S SYSTEMS FACTORS

• Older subjects • Lack of knowledge and training • Lack of consensus over the optimal assessment
• Patients experiencing • Lack of confidence of delirium
comorbid dementia • Lack of suspicion • Location of care (worse in surgical rather than
• Fluctuating course of • Lack of time of the clinical staff medical settings)
presentation • Expectation that altered mental status or • Busy clinical settings (especially low
• Presence of hypoactive delirium are a “normal occurrence” in certain nurse-to-patient ratio)
features medical settings, such as the ICU • Inadequate application of sedation holidays in
sedated-ventilated patients
• The rapid transfer of patients from one unit
to another, which may decrease the proper
documentation and diagnosis

858 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
In order to accurately diagnose delirium, clinicians need Box 41.4 OBJECTIVE MEASUR ES FOR THE DIAGNO-
to be vigilant and have a high level of suspicion, particularly SIS OF DELIRIUM
in patient populations at higher risk, such as the elderly and
those with the risk factors included in the mnemonic “end • DSM-IV-TR (“Gold Standard”; APA, 1994)
acute brain failure” (see Table 41.2).
To date, the diagnostic gold standard for delirium remains • Delirium Rating Scale (DRS) (Trzepacz, Baker, et al., 1988)
the Diagnostic and Statistical Manual for Mental Disorders • Confusion Rating Scale (CRS) (Williams, Ward, et al., 1988)
(DSM). The previously published version was the Fourth
Edition-Text Revised, published in 1994 (DSM-IV-TR; APA, • Confusion Assessment Method (CAM) (Inouye et al., 1990)
1994). A  new DSM-5 (5th Edition) has changed the criteria • Delirium Symptom Interview (DSI) (Albert et al., 1992)
slightly. The new DSM criteria change the diagnostic focus
away from a “disturbance of consciousness” (DSM-IV) to a • Delirium Assessment Scale (DAS) (O’Keeffe, 1994)
“disturbance in attention”—Criterion A. Criterion B relates • Cognitive Test for Delirium (CTD) (Hart et al., 1996)
to the acute or subacute onset of symptoms (to differenti-
ate from more chronic cognitive deficits such as dementia; • NEECHAM Confusion Scale (Neelon & Champagne, 1996)
formerly Criterion C). New Criterion C refers to additional • Confusional State Evaluation (CSE) (Robertsson, Karlsson,
disturbances in cognition not better explained by another et al., 1997)
neurocognitive disorder (akin to former Criterion B in
DSM-IV). New to DSM-5 is Criterion D, which stipulates • Memorial Delirium Assessment Scale (MDAS) (Breitbart
that A and B “do not occur in the context of severely reduced et al., 1997)
level of arousal or coma.” And, finally, Criterion E is identi- • Delirium Severity Scale (DSS) (Bettin, Maletta, et al., 1998)
cal to Criterion D in DSM-IV, which stipulates that there
is physical evidence that can give an organic etiology for the • Delirium Index (DI) (McCusker, Cole, et al., 1998)
syndrome. Notice that DSM-5 seems to have got rid of the
• Delirium Rating Scale–Revised-98 (DRS-R-98) (Trzepacz
presence of psychotic symptomatology (formerly Criterion B
et al., 2001)
in DSM-IV) as part of the diagnostic criteria.
The International Statistical Classification of Diseases and • Intensive Care Delirium Screening Checklist (ICDSC)
Related Health Problems (ICD-10; see Box 41.3) provides its (Bergeron et al., 2001)
own diagnostic criteria (WHO, 1992, 2010). Yet, both cri-
• Confusion Assessment Method for the Intensive Care unit
teria have at their core the same three cardinal features: dis-
(CAM-ICU) (Ely, Margolin, et al., 2001)
turbance of consciousness (i.e., reduced clarity of awareness
of the environment, with reduced ability to focus, sustain, • Delirium Detection Score (DDS) (Otter, Martin, et al., 2005)
or shift attention); global disturbance of cognition (e.g.,
• Nursing Delirium Screening Scale (Nu-DESC) (Gaudreau et
problem-solving impairment or memory impairment) and/
al., 2005a)
or perceptual disturbance (e.g., illusions or hallucinations);
and an acute or subacute onset with a tendency to fluctuate • Delirium Detection Tool–Provisional (DDT-pro) (Kean,
(WHO, 2010; APA, 2013). These criteria are commonly used Trzepacz, et al., 2010)
by medical personnel for the definite diagnosis of delirium.
• Diagnostic and Statistical Manual of Mental Disorders, 5th
There are numerous clinically available instruments (see
Edition (DSM-5) (“New Gold Standard”) (APA, 2013)
Box 41.4) developed to help medical providers screen for the
presence of delirium. These instruments were designed to
help nonpsychiatrists (e.g., nurses, internists, and research
assistants) diagnose and follow the progression of delirium concerns regarding the generalizability of the CAM-ICU
(Trzepacz, Baker,  & Greenhouse, 1988; Marcantonio et  al., findings (Young & Arseven, 2010; Boot, 2012):
1994). All of these scales have been derived from, and vali-
dated against, expert psychiatric assessments using DSM or On closer scrutiny of the methodology of CAM-ICU
ICD diagnostic criteria. Some of these tools were designed studies, it is important to highlight some consider-
and are better used for the purpose of screening for delirium ations when interpreting the data:  (1)  three of the
(e.g., CAM, based on DSM-III-R, original sample size n = studies examined were conducted by the tool’s author
56 [Inouye et al., 1990], CAM-ICU, based on CAM, origi- who may have had a vested interest, and given that
nal sample size n = 38 [Ely  et  al.,  2001]). The CAM and only 39% of the study population were intubated, it
CAM-ICU offer a binary result (i.e., delirium or no delirium), is difficult to conclude CAM-ICU’s validity in intu-
and there is no severity scale. Unfortunately, these tools have bated patients (van Eijk, van Marum, et  al., 2009);
a high false-positive rate (as high as 10%); therefore, the team (2)  the study population sizes were relatively small,
that developed the instrument recommends that all patients thus CAM-ICU adaptability into different health-
identified as delirious by screening instruments “have fur- care organizations remains unclear (Page, Navarange,
ther evaluation to confirm the diagnosis” (Inouye et  al., et  al., 2009); (3)  examination of the exclusion crite-
1990; Liptzin & Levkoff, 1992). In fact, some have expressed ria (e.g., history of psychosis, learning disabilities,

41. D E L I R I U M • 859
depression, dementia) of these studies raises ques- of mental illness (e.g., psychosis, learning disabilities,
tions to CAM-ICU transferability to these popula- depression, dementia) were excluded, despite reports
tions, especially as these conditions can have similar indicating delirium being of a medical cause (Boot,
characteristics but very different treatment pathways; 2012, p.187))
(4)  studies have demonstrated although CAM-ICU
has high validity in the non-intubated ICU patients, In fact, a prospective study was used to assess the diagnostic
it seems to fail to detect milder, subtler symptoms of validity of the CAM administered at the bedside by nurses
hypoactive delirium; (5)  individuals with a history in daily practice during a 5-month period to all patients
(n = 258) consecutively admitted to an acute geriatric ward
of a university hospital (Lemiengre, Nelis, et al., 2006). The
Box 41.5 PRIMITIVE R EFLEXES CAM as administered by bedside nurses had a 66.7% sensi-
tivity and a 90.7% specificity. The study suggested that bed-
Primitive reflexes are clinical features that indicate brain dysfunc- side nurses had difficulties with the identification of elderly
tion but that cannot be precisely localized or lateralized. When patients with delirium and that additional education about
present, these signs suggest cortical disease, especially frontal delirium is warranted, with special attention to guided train-
cortex, resulting in disinhibition of usually extinguished or sup- ing of bedside nurses in the use of an assessment strategy such
pressed primitive reflexes. Their clinical significance is uncertain as the CAM for the recognition of delirium symptoms.
and is difficult to correlate with psychiatric illnesses and other A subsequent study compared the diagnostic accuracy of
behavior disorders, including delirium. “delirium experts who diagnosed delirium in 75/181 sub-
• Glabellar Reflex: With the examiner’s fingers outside jects vs. the CAM-ICU as performed in ‘routine practice’ ”
of patient’s visual field, tap the glabellar region at a rate (i.e., nonresearch setting, conducted by a nonresearch nurse),
of one tap per second. A pathological response is either where delirium was diagnosed in 35 out of 181 cases. This
absence of blink, no habituation, or a shower of blinks. yielded a sensitivity of 47% (95% CI = 35%–58%), specificity
Normal response = blinking to the first few taps with rapid of 98% (95% CI  =  93%–100%), positive predictive value of
habituation. 95% (95% CI = 80%–99%), and negative predictive value of
72% (95% CI  =  64%–79%). This suggested that the “speci-
• Rooting Reflex: Tested by stroking the corner of the patient’s ficity of the CAM-ICU as performed in routine, daily prac-
lips and drawing away. Pursing of the lips and movement of tice appears to be high but sensitivity low. The low sensitivity
the lips or head toward the stroking is a positive response. hampers early detection of delirium by the CAM-ICU” (Van
• Snout Reflex: Elicited by tapping the patient’s upper lip with Eijk et al., 2011).
finger or percussion hammer causing the lips to purse and the Moreover, a recent systematic review and meta-analysis
mouth to pout. of studies on delirium in critically ill patients (i.e., inten-
sive care units, surgical wards, emergency rooms) published
• Suck Reflex: Tested by placing your knuckles between the between 1966 and 2011 (including 16 studies, n = 1,523; and
patient’s lips. A positive response would be puckering of covering five delirium screening tools) demonstrated that the
the lips. CAM-ICU was the most specific bedside tool for the assess-
• Grasp Reflex: Elicited by stroking the patient’s palm toward ment of delirium in critically ill patients (Neto et al., 2012).
fingers or crosswise while the patient is distracted, causing the Considering only the two most commonly used tools, the
patient’s hand to grasps the examiner’s fingers. pooled sensitivities and specificities of the CAM-ICU for
detection of delirium were 75.5% and 95.8%, compared to the
• Palmomental Reflex: Test by scratching the base of the ICDSC at 80.1% and 74.6%, respectively. All but one study
patient’s thumb (noxious stimulus of thenar eminence). was performed in a research setting, and that study suggested
A positive response occurs when the ipsilateral lower lip and that, with routine use of the CAM-ICU, half of the patients
jaw move slightly downward, and does not extinguish with with delirium were not detected. The authors concluded that
repeated stimulation. “the low sensitivity of the CAM-ICU in routine, daily prac-
• Babinski Sign: Downward (flexor response) movement of the tice may limit its use as a screening test” (Neto et al., 2012).
great toe in response to plantar stimulation. Other scales have been designed to diagnose and assess
the severity of a delirium episode. These include the Delirium
• Adventitious Motor Overflow: Seen as the examiner tests one Rating Scale (DRS), based on DSM-III, original sample size
hand for sequential finger movements, and the fingers of n = 20 (Trzepacz et al., 1988); the DRS-98, based on DSM-IV,
the other hand wiggle or tap. Also if there are choreiform original sample size n = 68 (Trzepacz, Mittal, et  al., 2001);
movements. the Memorial Delirium Assessment Scale (MDAS), based
• Double Simultaneous Stimulation Discrimination: Tested with
on DSM-IV, original sample size n = 68 (Breitbart et  al.,
the patient’s eyes closed. The examiner simultaneously brushes
1997); and the Intensive Care Delirium Screening Checklist
a finger against one of the patient’s cheeks and another finger
(ICDSC), based on DSM-IV, original sample size n = 93
against one of the patient’s hands, asking the patient where he
(Bergeron et  al., 2001), among others. (Please see Box 41.4
has been touched.
for a comprehensive list of delirium diagnostic tools.) A sig-
nificant advantage of diagnostic tools that measure delirium

860 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
severity is that they provide clinicians a means to measure the ICU using a validated tool. After their review of avail-
the severity of the episode and determine whether the con- able delirium assessment tools, the 2013 PAD guideline
dition seems to be worsening or improving (i.e., consecutive group concluded that the Confusion Assessment Method
measures). A severity score is also useful in research studies, as for the ICU (CAM-ICU) and the Intensive Care Delirium
particular interventions may not be able to fully prevent delir- Screening Checklist (ICDSC) are the ICU delirium screen-
ium but may lessen the severity of the episode. Finally, sever- ing tools with the strongest validity and reliability (Davidson,
ity scales may provide the ability to diagnose subsyndromal Harvey, Bemis-Dougherty, Smith, & Hopkins, 2013).
delirium (i.e., patients presenting with mental status changes The most critical part of the assessment, given the char-
that do not rise to the level of full DSM or ICD diagnostic acteristic waxing and waning of this syndrome, is to add to
criteria). Clinical experience and accumulating research data the clinical examination all available clinical information,
seem to indicate that the presence of subsyndromal symp- including interview of the family members and caregiv-
toms may have similarly negative effects in the patients who ers, nursing and medical staff, and a thorough review of the
are experiencing them. A study of ICU patients demonstrated chart for behaviors exhibited during the preceding 24 hours.
that patients experiencing subsyndromal delirium experi- Another potential clue of the presence of delirium may come
enced significantly greater ICU mortality (10.6% vs. 2.4%), from a thorough neurological examination. In our clinical
longer ICU stays (5.2 days vs. 2.5 days), and longer total hospi- experience, patients with delirium tend to exhibit a reemer-
tal stays (40.9 days vs. 31.7 days) compared to patients experi- gence of primitive signs (see Box 41.5). This appears to be
encing similar medical illness but without delirium (Ouimet more consistent in cases of hypoactive delirium. The relation-
et al., 2007). As expected, these subsyndromal patients expe- ship between poor cognitive status and primitive reflexes has
rienced less ICU mortality (10.6% vs. 15.9%), shorter ICU been described in patients suffering from HIV-related cogni-
stays (5.2  days vs. 10.8  days), but longer total hospital stays tive disorders (Tremont-Lukats, Teixeira, et al., 1999) and in
(40.9  days vs. 36.4  days) compared to patients experiencing cases of dementia (Paulson, 1977). One study described the
similar medical illness but with full clinical delirium (see presence of primitive reflexes in post-cardiotomy patients
Table 41.8). suffering from postoperative neuropsychiatric complica-
In general, no delirium screening or assessment tool is tions (Liu & Hsieh, 1993). A second study found that 55% of
expected to be perfect. Some have criticized the fact that even patients exhibiting at least two primitive reflexes and 80% of
the two psychometrically valid scales most commonly used in those exhibiting at least three primitive reflexes in the postop-
the ICU (i.e., CAM-ICU, ICDSC) have detection ranges of erative period developed delirium (Nicolson, Chabon, et al.,
delirium from 10% to >80%, in similar populations. It may 2011). Further studies are needed to determine whether an
very well be that vigilance and close clinical monitoring are assessment for the presence of primitive reflexes may add to
more useful than any commonly used tool. In fact, some the diagnostic accuracy for delirium or at least assist in the
authors have reported that the CAM-ICU scoring may be characterization of delirium type, and whether such assess-
affected by sedation, thus limiting its diagnostic accuracy ment has any prognostic value.
(Kress, 2010). A study of “real-life” conditions (as opposed to a Some have advocated the use of the electroencephalogram
research setting) suggested that “the CAM-ICU in daily prac- (EEG) as a way to identify and diagnose delirium. Engel and
tice showed not quite as good test characteristics as presented Romano were the first to describe the relationship between
in the original validation studies” (van Eijk et  al., 2011). In delirium and the diffuse slowing and progressive disorganiza-
fact, these authors found that “after stratification according to tion of rhythm seen in the EEG (Engel, Webb, & Ferris, 1945)
type of delirium, sensitivity of the CAM-ICU was lowest in (Engel, Webb et al. 1945). The most common EEG findings in
the hypoactive subgroup (31%; 95% CI, 17%–48%); highest delirium include slowing of peak and average frequencies, and
in the hyperactive delirious patients (100%; 95% CI, 56%– decreased alpha activity but increased theta and delta waves.
100%); and intermediate in the mixed-type patients (53%; Studies suggest that EEG changes correlate with the degree
95% CI, 35%–74%); exhibiting particularly poor test char- of cognitive deficit, but there does not appear to be a rela-
acteristics in neurocritical care patients (sensitivity 17%; 95% tionship between EEG patterns and delirium motoric type
CI, 1%–64%) (van Eijk et al., 2011). A recent study suggested (Romano & Engel, 1944; Engel & Romano, 1959; Koponen,
that clinical assessments by critical care physicians may iden- Partanen, et al., 1989; Koponen, Hurri, et al., 1989; Trzepacz,
tify delirium more rapidly and more accurately than screening Sclabassi, & Van Thiel, 1989; Trzepacz, Leavitt, et al., 1992;
tool assessments (Bigatello et al., 2013). Engel  & Romano, 2004; Plaschke, Hill, et  al., 2007). The
It is unclear what the new DSM-5 diagnostic criteria will clinical usefulness of EEG in the diagnosis of delirium may
mean to the diagnostic accuracy of existing tools. It is likely be constrained by its limited specificity (given that there are
that some of them may require revisions, yet  all will need many conditions and medications that may affect the EEG)
to be validated taking the new criteria into consideration. and the practicality of conducting the test (particularly in the
Nonetheless, it may be important for the clinician to be aware case of agitated and combative patients). Still, the EEG can
that the 2013 American College of Critical Care Medicine be useful in differentiating delirium from other psychiatric
(ACCM)/Society of Critical Care Medicine (SCCM) clini- and neurological conditions such as catatonic states, seizure
cal practice guidelines for pain, agitation, and delirium activity (e.g., nonconvulsive status), medication side effects
(PAD), based on available evidence, strongly recommend that (e.g., posterior reversible encephalopathy syndrome due to
critically ill patients be routinely monitored for delirium in the use of calcineurin inhibitors), or the manifestations of

41. D E L I R I U M • 861
the behavioral and psychological symptoms associated with M A N AG E M E N T OF   DE L I R I U M
dementia (BPSD). BPSD affects between 50% and 90% of
patients suffering from dementia and is usually characterized
A DE QUAT E M E DIC A L M A N AG E M E N T
by three main syndromes:  agitation/aggression, psychosis,
and an affective component. At times, apathy (as in the case of The management of delirium has five main components (see
hypoactive delirium) or disinhibition (as in the case of hyper- Box 41.6):
active delirium) can be seen, particularly when there is frontal
lobe involvement (Finkel, Costa e Silva, et al., 1996). 1. Recognition of patients at risk (see Tables 41.1
Finally, a 24-hour accelerometer-based activity monitor and 41.2);
has been used to categorize the motoric behavior of patients
with delirium. The continuous wavelet transform (CWT) 2. Implementation of prevention techniques (with pharma-
provided by the instrument can then be used to character- cological and nonpharmacological approaches), especially
ize a delirium as hyperactive, hypoactive, or mixed (Godfrey, in populations identified to be at high risk;
Conway, et  al., 2009, 2010; Meagher, 2009). The technique 3. Enhanced surveillance and screening;
has yet to find a clinical use, however.

Box 41.6 ALGORITHM FOR THE MANAGEMENT OF DELIRIUM

I. Recognition of patients at risk


A. A particular patient’s odds of developing delirium are associated with the interaction between the following conditions:
i. Knowledge of a patient’s characteristics.
ii. Predisposing and precipitating medical risk factors (Table 41.4).
iii. Modifiable and non-modifiable risk factors for that particular patient or patient population (Table 41.5).
iv. Specific medical conditions and surgical procedures the patient is exposed to.
B. Obtaining the patient’s baseline level of cognitive functioning using information from accessory sources (e.g., Informant
Questionnaire on Cognitive Decline in the Elderly [IQCODE]).
II. Implementation of prevention techniques
A. A key focus should be placed on prevention strategies, particularly in “at risk” populations.
i. Avoid all pharmacological agents with high deliriogenic potential or anticholinergic load, if possible.
ii. Promote a non-pharmacological sleep protocol.
iii. Early mobilization.
B. For patients in the ICU, especially those on ventilation or IV sedation, consider:
i. Sedating to a prescribed or target sedation level (e.g., RASS).
ii. Consider using the sedative agent with lowest deliriogenic potential:
• Dexmedetomidine use is associated with the lowest incidence of delirium.
• Propofol use is a good second choice; followed by midazolam.
iii. Reassess pain levels daily and titrate opioid agents to the lowest effective level required to maintain adequate analgesia.
• Hydromorphone is preferred as baseline agent of choice for pain management.
• Limit the use of fentanyl to rapid initiation of analgesia and as rescue agent.
• Avoid the use of opioid agents for sedation or management of agitation or delirium.
iv. Provide daily sedation holidays:
• Interrupt sedative infusions daily until the patient is awake
• Restart sedation, if needed, at the lowest effective dose
• Reassess target sedation level (e.g., RASS).
III. Enhanced surveillance, screening and early detection
A. Most important aspects of surveillance:
i. Knowledge about the condition and presenting symptoms.
ii. A high level of suspicion.
B. Be vigilant for the development of delirium in high-risk groups:
i. Use a standardized surveillance tools (e.g., DRS-R-98; MDAS; CAM)
ii. Use psychiatric consultants (i.e., DSM-5/ICD-10 criteria)
iii. Be particularly aware of the presence of hypoactive delirium and its different manifestations.
C. Use psychiatric consultants to help with assessment and design of the treatment plan, if available.
D. Train medical personnel at all levels.
IV. Treatment of all forms of delirium
A. Identify and treat underlying medical causes.
(continued)

862 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
B. Treat or correct underlying medical problems and potential reversible factors.
C. Conduct an inventory of all pharmacological agents that have been administered to the patient.
i. Discontinue any medication or agent known to cause delirium or to have high anticholinergic potential, if possible, or insti-
tute a suitable alternative.
D. Implement early mobilization techniques, to include ALL of the following components:
i. Daily awakening protocols (sedation holiday), as described above.
ii. Remove IV lines, bladder catheters, physical restraints, and any other immobilizing apparatuses as early as possible.
iii. Initiate aggressive PT and OT as soon as it is medically safe to do.
a. In bedridden patients, therapy may be limited to daily passive range of motion.
b. Once medically stable, get the patient up and moving as early as possible.
iv. Provide patients with any required sensory aids (i.e., eyeglasses, hearing aids).
v. Promote as normal a circadian light rhythm as possible.
a. Better if this can be achieved by environmental manipulations, such as light control (i.e., lights on and curtains open
during the day; lights off at night) and noise control (i.e., provide ear plugs, turn off TVs, minimize night staff chatter).
b. Provide as much natural light as possible during the daytime.
vi. Provide adequate intellectual and environmental stimulation as early as possible.
E. Avoid using GABA-ergic agents to control agitation, if possible.
• Exception: cases of CNS-depressant withdrawal (i.e., alcohol, benzodiazepines, barbiturates); or when more appropriate
agents have failed and sedations are needed to prevent patient’s harm.
F. Adequately assess and treat pain.
i. Avoid the use of opioid agents for behavioral control of agitation.
ii. Rotate opioid agents from morphine to hydromorphone or fentanyl.
G. For the treatment of delirium (all types), consider using:
i. Acetylcholinesterase inhibitor (e.g., rivastigmine) for patients with a history of recurrent delirium or delirium superimposed
on known cognitive deficits. Physostigmine, for known causes of anticholinergic delirium.
ii. Melatonin (e.g., 3 mg q HS) or melatonin agonists (e.g., ramelteon 8 mg q HS) to promote a more natural sleep. If that is
ineffective, consider trazodone (e.g., 25–100 mg q HS) or mirtazapine (e.g., 3.75–7.5 mg q HS).
iii. Serotonin antagonist (e.g., ondansetron).
H. In case of hyperactive delirium, consider the use of the following agents:
i. Dopamine antagonist agents (to address DA excess) (e.g., haloperidol, risperidone, quetiapine, aripiprazole).
a. Moderate-dose haloperidol (e.g., <20 mg/24 hr in divided doses), is still considered the treatment of choice, if the
patient’s cardiac condition allows it.
b. Before using haloperidol:
• Obtain 12-lead ECG; measure QTc.
• Check electrolytes; correct K+ and Mg+, if needed.
• Carefully review the patient’s medication list and identify any other agents with the ability to prolong QTc.
• If possible, avoid other medications known to increase QTc and/or inhibitors of CPY3A4.
c. When the use of haloperidol is contraindicated or not desirable, atypical antipsychotics should be considered:
• There is better evidence for risperidone and quetiapine.
• ii.Limited data for: olanzapine, aripiprazole, perospirone.
• Avoid: clozapine, ziprasidone.
d. Discontinue dopamine antagonist agents’ use if QTc increases to >25% of baseline or >500 msec.
ii. Alpha-2 agonist agents (to address the NE excess) (e.g., dexmedetomidine, clonidine, guanfacine)
a. Consider changing primary sedative agents from GABA-ergic agents (e.g., propofol or midazolam) to an alpha-2 agent
(e.g., dexmedetomidine), starting at 0.4 mcg/kg/hr, then titrate dose every 20 minutes to targeted RASS goal.
iii. Anticonvulsant and other agents with glutamate antagonism or Ca+ Ch modulation (e.g., VPA, gabapentin, amantadine,
memantine).
I. Consider the use of NMDA-receptor blocking agents, to minimize glutamate-induced neuronal injury (e.g., amantadine,
memantine), particularly in cases of TBI and CVA.
J. G. In case of hypoactive delirium:
i. Evidence suggests that DA antagonists may still have a place, given the excess DA theory.
a. If haloperidol is used, recommended doses are in the very low range (i.e., 0.25–1.0 mg/24 hr). This is usually given as a
single nighttime dose, just before sundown.
b. If an atypical is preferred, consider an agent with low sedation (i.e., risperidone, aripiprazole).
ii. In cases of extreme psychomotor retardation or catatonic features, in the absence of agitation or psychosis, consider the use
of psychostimulant agents (e.g., methylphenidate, dextroamphetamine, modafinil) or conventional dopamine agonists (e.g.,
bromocriptine, amantadine, memantine).

(Modified from Maldonado, 2008a, 2009, 2011)


4. Treatment or correction of underlying medical problems have found psychiatric consultations to be superior to a mul-
and potentially reversible factors (e.g., correction of tidisciplinary geriatric team approach in treating individuals
electrolyte imbalances, infection, end organ failure, with depression and delirium (Cole, Fenton, et al., 1991).
sleep deprivation, pain, metabolic and endocrinological The recognition of patients at risk begins with knowl-
disturbances, substance or medication intoxication or edge of your patient’s characteristics; an assessment of the
withdrawal); and predisposing and precipitating medical risk factors to which
your patient is, or may be, exposed (Table 41.2); and being
5. Treatment of all forms of delirium (with pharmacological
acquainted with the modifiable and nonmodifiable risk fac-
and nonpharmacological approaches).
tors for that particular patient or patient population (Table
We advocate an integrative approach that incorporates 41.3). Finally, certain medical conditions and surgical proce-
input and collaboration among all pertinent medical teams, dures are more likely to be associated with the development
including psychosomatic medicine, critical care, internal of delirium than others. It is likely that a particular patient’s
medicine, surgery, neurology, nursing, respiratory therapy, odds of developing delirium are associated with the interac-
physical therapy, and occupational therapy teams, which we tions among these conditions. These have been discussed in
like to call the “CNS pharmacotherapy algorithm” (see Figure detail elsewhere (Maldonado, 2008a, 2014).
41.22) (Maldonado, 2008a, 2011, 2014). The idea is to opti- Prevention techniques have been found to be rather effec-
mize patient comfort (i.e., adequate sedation), minimize pain tive, especially when targeted to patients at high risk for devel-
(i.e., adequate analgesia), prevent or treat delirium, restore an oping delirium. There is a whole host of pharmacological and
adequate sleep–wake pattern, begin physical mobilization as nonpharmacological techniques available. It is important
early as possible, improve cognitive recovery, and return to that providers carefully consider the potential side effects ver-
baseline functional level as soon as medically possible. Others sus the benefits associated with effective delirium prevention.

Is Pt @ targeted RASS?
Default RASS: -1 to +1
NO
YES Reassess Daily Goals

YES
Are there known Continue Daily
intracranial processes? Awakening Protocol

Neurology Consult
&
Management RASS
Algorithm Score?

RASS= +2 to +3 RASS= ≥+4


RASS= -2 to -5

YES
Re-assess Metabolic problems? Address & Correct This constitutes a
Neuro Medico/Psychiatric
Status Emergency
NO

YES Initiate Pain


Hold/DC ALL Pain? Management
CNS depressant Rxs Algorithm
Dose opioids to pain level; not NO
behavior-reassess 1hr
Initiate Substance
Likely YES Assess pertinent Physiological causes:
Abuse & With drawal
Substance Abuse
Management Algorithm • Metabolic
Or Withdrawal?
• Pain
Improved MSE? • Anxiety
NO • Delirium
YES
Assess & Initiate • CNS Stimulant Abuse
YES Anxiety Management • CNS Depressant Withdrawal
Likely Anxiety
Algorithm
Hypoactive Disorder?
Delirium
NO
If possible harm to patient/staff, initiate
Psych Consult & Initiation Psych Consult & Initiation
Behavioral Rescue Protocol/
Hypoactive Delirium Hyperactive Delirium Call urgent Psychiatry Consult
Algorithm Algorithm

Figure 41.22 CNS Pharmacotherapy Model. (Adapted from Maldonado, 2008a, 2011, 2014)

864 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
The early recognition and prompt treatment of delirium is 2. Treatment or correction of underlying medical problems
of utmost importance, not only because of issues of safety and potentially reversible factors (e.g., correction of electro-
for patient and staff, but also because of the serious negative lyte imbalances, infection, end organ failure, sleep depriva-
consequences of no treatment or delayed treatment. In fact, tion, pain, metabolic and endocrinological disturbances,
a recent study demonstrated that patients whose delirium substance or medication intoxication or withdrawal); and
treatment was delayed were more frequently mechanically
3. Correction of the neurochemical derangement (triggered
ventilated (50.0% vs. 22.3%; p = 0.012), had more nosoco-
by the underlying cause) which leads to the behavioral
mial infections (including pneumonia) (p < 0.05), and had
manifestations of delirium (see Table 41.7).
a higher mortality rate (p < 0.001) (see Figure  41.23) than
patients whose treatment was promptly started (Heymann
et  al., 2010). These data suggest that a delay in initiating
delirium therapy may be associated with increased morbidity NON PH A R M AC OL O G IC A L
and mortality. M A N AG E M E N T S T R AT E G I E S
The most important aspects of adequate surveillance Nineteen studies have been published on nonpharmacological
and early detection include knowledge about the condition prevention strategy in the non-ICU environment (see Table
and presenting symptoms (of all motor forms) and main- 41.10). The original study suggesting that nonpharmacological
taining a high level of suspicion, especially in populations approaches may be of use was a nonrandomized clinical trial,
at risk. There are several surveillance tools and techniques where hospitalized elderly patients were assessed for manifes-
(discussed earlier in this chapter) that can be used to effi- tations of delirium in response to the correction of environ-
ciently screen subjects, especially those at greatest risk. mental factors commonly associated with increased risk for
Adequate training of medical personnel at all levels is para- delirium (Inouye et al., 1999). The multicomponent protocol
mount. Surveillance should be effectively implemented by consisted of simple techniques applied by the hospital staff,
all practitioners. including reorientation, appropriate cognitive stimulation
Once diagnosed, adequate management of delirium three times a day, the implementation of a nonpharmacologi-
includes the following steps: cal sleep protocol to help normalize patient’s sleep–wake cycle,
early mobilization after surgery or extubation, timely removal
1. Management of the behavioral and psychiatric manifesta- of catheters and restraints, correction of sensory deficiencies
tions and symptoms to prevent the patient from self-harm (i.e., by eyeglasses and hearing aids), and early correction of
or harming of others (e.g., use of tranquilizing agents to dehydration and electrolyte abnormalities. As a result of these
manage agitation); environmental manipulations, they observed an astonishing
40% reduction in the risk of delirium (see Figure 41.24). This
landmark work was preceded by two negative studies and fol-
1.0 lowed by seven positive and five negative ones. Among the
Delayed therapy
Immediate therapy
positive studies, there were three positive randomized clinical
trials, and these are condensed in Table 41.11. They were all
0.8 conducted among elderly orthopedic patients, and all found
that the implementation of nonpharmacological multicom-
ponent interventions led to a reduction in the occurrence of
Cumulative survival

0.6
delirium and fewer medical complications in the intervention
groups (Marcantonio, Flacker, 2001; Vidan, Serra, Moreno,
0.4 Riquelme, & Ortiz, 2005; Lundstrom et al., 2007). Given the
findings reported by others, a multicomponent approach has
been developed that targets identified, treatable contributing
0.2 factors and addresses them early on: the Hospital Elder Life
Program (Inouye  & Charpentier, 1996; Inouye, Bogardus,
et al., 1999). This program involves the implementation, by an
0.0
interdisciplinary team, of targeted interventions for known
0 20 40 60 80 risk factors (i.e., cognitive impairment, sleep deprivation,
Time since first occurence of delirium (days) immobility, dehydration, vision or hearing impairment) for
Survival Estimates for Delirious ICU Patients with
Figure 41.23
cognitive decline in the elderly.
Immediate or Delayed Delirium Treatment. Kaplan–Meier survival More recently, a meta-analysis of published studies has
estimates for 184 ICU patients with delirium treated within 24 hours of found that, although the multicomponent interventions
diagnosis (immediate therapy group) and 20 patients in whom delirium appeared to be effective in preventing delirium among post-
treatment started > 24 hours after diagnosis (delayed therapy group); operative hip replacement patients, the intervention made
the delayed therapy group showed a lower probability of surviving until
ICU discharge than the immediate therapy group (vertical markers
no difference in a number of important secondary measures,
show censorship points). There was a significantly higher risk of death in including discharge location or post-discharge dependency,
the delayed therapy group versus the immediate therapy group (log rank length of hospital stay (p = 0.12), or mortality rate (p = 0.77)
test, P < 0.001; Breslow test, P = 0.013). (Heymann et al., 2010) between intervention and control groups (Holroyd-Leduc,

41. D E L I R I U M • 865
Table 41.10 NONPHAR MACOLOGICAL PR EVENTION STR ATEGY STUDIES IN THE NON-ICU EN VIRONMENT

DELIR IUM INCIDENCE % P VALUE


DELIR IUM
STUDY(N = 19) POPULATION INTERVENTION DEFINITION CONTROL INTERVENTION

Schindler et al., 1989 CABG NP—peri-op Psych intervention vs. DSM-III 0 (0/17) 12.5 (2/16) ns
RCT, n = 33 usual care

Wanich et al., 1992 Gen IM elderly NP—nursing intervention for elderly DSM-III 22 (22/100) 19 (26/135) p = 0.61
NRCT, n = 235 patients hospitalized patients vs. usual care

Inouye et al., 1999 Gen IM elderly NP—multi-component intervention CAM 15 (64/426) 9.9 (42/426) p = 0.02
NRCT, n = 852 patients vs. usual care

Millisen et al., 2001 Traumatic hip Fx Sx NP—multi-component vs. usual care CAM 23.3 (14/60) 20 (12/60) p = 0.82
NRCT, n = 120 repair

Marcantonio et al., 2001 Elderly pts. after Hip NP—multi-component intervention CAM 50 (32/64) 32 (20/62) p = 0.04
RCT, n = 126 Fx Sx vs. usual care

Tabet et al., 2005 Gen IM elderly NP—Staff education vs. usual care Single assessment 19.5 (25/128) 9.8 (12/122) p = 0.034
NRCT, n = 250 patients psychiatrist

Wong et al., 2005 Traumatic hip NP—multi-component vs. usual care CAM 35.7 (10/28) 12.7 (9/71) p = 0.012
Pre- and post-eval Fx Sx repair

Vidan et al., 2005 Elderly pts. after Hip NP—multi-component intervention CAM 45.2 (70/155) 61.7 (100/164) p = 0.003
RCT, n = 319 Fx Sx vs. usual care For ≥1 major
complications

Lundstrom et al., Elderly pts. after NP—multi-component intervention Organic Brain 75.3 (73/97) 54.9 (56/102) p = 0.003
RCT, n = 199 Hip Fx Sx vs. usual care Syndrome
Scale (OBSs)

(Caplan & Harper, 2007) Geriatric ward NP—usual care vs. CAM 38.1 (8/21) 6.3 (1/16) p = 0.032
Pre–post eval, n = 37 volunteer-mediated intervention
(Inouye style)

Taguchi et al., 2007 Esophageal CA Normalization of their natural NEECHAM scale 40 (2/5) 16 (1/6) P = 0.014
RCT, n = 11 subjects circadian rhythm by light therapy

Benedict, Hazelett, Acute Care for Elders NP—Delirium prevention protocol Modified [3.24] [3.76] p = 0.368
et al., 2009 (ACE) units vs. usual care NEECHAM scale
NRCT, n = 65

Schweickert et al., 2009 MICU Early exercise and mobilization (PT & CAM-ICU Return to Return to pp = 0.0202
RCT, n = 104 OT) at daily sedation interruption vs. independence independence by
sedation interruption by D/H: 35% D/H: 59%
Delirium Delirium duration:
duration: 2 days
4 days
Holroyd-Leduc Traumatic hip NP—multi-component delirium CAM Pre-implemen- Post-implemen- p = 0.84
et al., 2010 Fx Sx repair strategies tation tation incidence
NRCT, n = 134 incidence 31
33 (20/64)
(23/70)

Holroyd-Leduc Comprehensive geriatric assessment CAM (2) The multicomponent


et al., 2010 and multi-component Organic Brain interventions for
Meta-analysis, 3 studies interventions targeted at precipitants Syndrome the management of
(n = 489) of delirium Scale (1) delirium had:
• No effect on mortality
(summary RR 1.08,
95% CI 0.81–1.44;
p for heterogeneity =
0.77).
• No effect on length
of stay (summary
weighted mean
difference 3.25 days,
95% CI -2.85 to
9.34 days; p for
heterogeneity =0.12)
• There was no impact
on post-discharge
dependency,
function, or the need
for institutional care.

Björkelund, Hommel, Elderly hip Fx Sx NP—multi-component delirium OBS Scale 34 (45/132) 22 (29/131) p = 0.096
et al., 2010 repair strategies
NRCT, n = 263

Colombo et al., 2012; All patients admitted NP—patients underwent both CAM-ICU 35.5 (60/170) 22 (31/144) p = 0.020
n = 314; to mixed (med/surg) a reorientation strategy and
ICU over a year environmental, acoustic, and visual
stimulation. Study compared pts.
in observation arm (control) vs.
intervention arm

Gagnon et al., 2012, Palliative care NP—multi-component administered Confusion rating 43.9 (370/842) 49.1 (330/674) P = 0.045
randomized delirium patients, in 2 cancer to patient and family education vs. scale (CRS)
prevention trial; n = 1516 centers usual care
Martinez et al., 2012 Older adults in gen Randomized to receive a CAM 13.3 (19/143) 5.6 (8kal/14 4) P = 0.027
n = 287 medicine ward multi-component management
protocol, delivered by family
members (144 patients) or standard
management (143 patients).
0.25 had no effect on the primary outcome, overall delirium inci-
Usual care dence (p = 0.66); and, similarly, it had no effect on secondary
Cumulative Incidence
0.20 outcomes of delirium severity, total delirious days, or dura-
tion of the first delirium episode (Gagnon, Allard, Gagnon,
Merette, & Tardif, 2012).
of Delirium
0.15
Finally, a study examined the administration of nonpro-
0.10 Intervention fessional, family-provided multicomponent intervention
among high-risk elderly patients admitted to a medicine ward
0.05 Median length
(Martinez, Tobar, Beddings, Vallejo,  & Fuentes, 2012). The
of stay intervention consisted of educating family members about
0.00 the clinical features and prognostic implications of delirium,
1 3 5 7 9 11 plus the usual content of the traditional multicomponent
Day intervention, but fully administered by family members. They
found a significant reduction in the occurrence of delirium in
A MultiComponent Intervention to Prevent Delirium in
Figure 41.24
Hospitalized Older Patients. (Inouye et al., 1999) the intervention group (p = 0.027). It is unclear why the mul-
ticomponent protocol yielded no results when implemented
by the nursing staff in a “real world” scenario, but a possible
explanation is that the average medicosurgical unit may not
Khandwala, & Sink, 2010). A separate study on the same pop-
have the same nurse-to-patient ratio as research units, making
ulation and by the same researcher, but this one conducted
the implementation of the protocol challenging. Therefore, an
in a nonresearch environment, found that the intervention
approach that combines the best of all available protocols and
had no effect on the overall delirium rate (p = 0.84) and
that takes into consideration available resources and limita-
no significant difference on secondary measures (i.e., mean
tions of each clinical setting may work best.
length of hospital stay), falls, or discharge to long-term care
Of significance, a study of mechanically ventilated, critically
facilities (Holroyd-Leduc et  al., 2010). Finally, a large study
ill patients used a combination of early physical and occupa-
(n = 1,516) followed patients from admission to death at
tional therapy during periods of daily interruption of sedation
seven palliative care centers and compared usual care versus
to demonstrate a significantly greater return to independent
a nonpharmacological delirium-preventive intervention. The
functional status (p = 0.02), shorter duration of delirium
study found that the multicomponent preventive intervention
(p = 0.02), and more ventilator-free days (p = 0.05) (Schweickert

Table 41.11 SELECTED NONPHAR MACOLOGICAL RCT-PR EVENTION INTERVENTION STUDIES

STUDY AUTHORS;
TYPE; SIZE STUDY DETAILS FINDINGS

Marcantonio Studied the effectiveness of proactive While there were not statistical differences between intervention
et al., 2001 geriatric consultation compared to usual and control groups regarding baseline measures and characteristics,
RCT, n = 126 care in reducing delirium in patients 65 the study found a reduction in the occurrence of delirium in the
and older admitted emergently for surgical intervention group (32%) compared to usual care (50%) (P = 0.04).
repair of hip fracture. There was an even greater reduction in cases of “severe delirium,”
occurring in 12% of intervention patients and 29% of usual-care
patients. Despite this reduction in delirium, length of stay did not
significantly differ between intervention and usual care.

Vidan et al., 2005 Studied patients were randomly assigned The study showed that intervention subjects experienced shorter
RCT, n = 319 to daily multidisciplinary geriatric hospital stay (median length 16 vs. 18 days) (P = 0.06); lower
intervention or usual care during in-hospital mortality (0.6% vs. 5.8%, P = 0.03) and lower rate of major
hospitalization in the acute phase of hip medical complications (45.2% vs. 61.7%, P = 0.003). Overall, after
fracture. adjustment for confounding variables, the geriatric intervention
was associated with a 45% lower probability of death or major
complications. More patients in the geriatric intervention group
achieved a partial recovery at 3 months (57% vs. 44%, P = 0.03),
but there were no difference between groups at 6 and 12 months
assessments.
Lundström Studied randomly assigned elderly patients Patients in the intervention group experience less postoperative
et al., 2007 after femoral neck fracture repair to delirium (54.9% vs. 75.3%, p = 0.003); those who did experience
RCT, n = 199 postoperative care in a specialized geriatric delirium did so for a shorter duration compared with controls (5.0 ±
ward (i.e., staff education focusing on the 7.1 days vs. 10.2 ± 13.3 days, p = 0.009). Patients in the intervention
assessment, prevention, and treatment of group suffered from fewer complications (e.g., decubitus ulcers,
delirium and associated complications) or a urinary tract infections, nutritional complications, sleeping problems,
conventional orthopedic ward. and falls) and shorter hospital stays (28.0 ± 17.9 days vs. 38.0 ±
40.6 days, p = 0.028).

868 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
et al., 2009). This is the first randomized, nonpharmacological time patients require mechanical ventilation, shorten the
intervention study shown to reduce ICU delirium. Others have length of the patient’s stay in the intensive care unit, and
found similar results (i.e., lower total amount benzodiazepine decrease the overall hospital length of stay (Strom  & Toft,
use, higher level of functional mobility, and reduction in delir- 2011), thus providing further evidence that the use of seda-
ium rates) by combining sedation reduction and early mobi- tive drugs should be reduced. Patients should be mobilized
lization with physical rehabilitation (Needham & Korupolu, as early as possible, and individual patient sedative needs
2010). It is likely that the minimization of offending agents should be evaluated on a daily basis to minimize the untow-
(e.g., sedatives), along with early mobility, may have something ard effects of these agents and optimize the care of each indi-
to do with the decreased rate of delirium among this cohort. vidual patient.
Previously, I have suggested an approach that incorporates Given the high rate of underdiagnosed and missed cases,
nonpharmacological and pharmacological approaches to the the routine use of assessment scales or diagnostic interviews
prevention and treatment of delirium, which is summarized in by properly trained personnel will facilitate timely treatment
Box 41.6 but expanded elsewhere (Maldonado, 2008a, 2009, and prevention of complications of delirium. Early involve-
2011). Given the significant adverse consequences of delirium, ment of a medical-psychiatric team is valuable for both pre-
the primary goal should be prevention, particularly in popu- vention and early intervention. An active search for possible
lations at high risk. Key among them is minimizing the det- etiologies of delirium must first rule out the common causes
rimental effects of prolonged and uninterrupted sedation. of the syndrome (see Table 41.2). This must include a review
Therefore, we recommend that sedation be targeted to a pre- of all medications and identification, and possible discon-
determined goal using sedation scales such as the Richmond tinuation of agents with high deliriogenic potential (see
Agitation Sedation Scale (RASS; see Sessler, et al., 2002). As Table 41.4). Appropriate diagnostic tests and assays should
previously noted, daily awakening and sedation holidays are be ordered and reviewed in a timely fashion and all abnormal
highly recommended. These will allow patients to awaken; findings addressed accordingly.
then, if still needed, sedation should be restarted at the lowest Conduct an inventory of all pharmacological agents that
needed level—again targeting a predetermined sedation level. have been administered to the patient. Any medication or
Certain sedative agents are preferred to others, given pharma- agent known to cause delirium (see Table 41.4) or to have high
cological and pharmacodynamic qualities; and some appear to anticholinergic potential (see Table 41.5) should be discontin-
have less deliriogenic potential than others (Maldonado, van ued if possible; if not, a suitable less anticholinergic alternative
der Starre, Wysong, & Block, 2004; Pandharipande, Pun et should be substituted.
al., 2007; Maldonado, 2008a, 2009, 2011; Riker et al., 2009). Immobilizing lines and devices (e.g., chest tubes, IV lines,
Therefore, dexmedetomidine may be the preferred choice bladder catheters) should be removed as early as possible.
when clinically indicated and tolerated, followed by propofol, Similarly, physical restraints should be avoided and elimi-
and last by midazolam. Similarly, some analgesic agents may nated as soon as it is safe to do. More importantly, early mobi-
be better at preventing or controlling delirium. For example, lization (e.g., daily interruption of sedation, minimization
despite its faster onset of action, fentanyl’s clearance is much of restraint use, early physical and occupational therapy) has
slower than hydromorphone; thus it tends to accumulate, and proven to be a key factor in minimizing delirium and improv-
this may be problematic (see Figure 41.25). Therefore, we pre- ing the odds of returning to independent functioning upon
fer the use of hydromorphone as the baseline agent of choice discharge home (Schweickert et al., 2009).
for pain management and sedation supplementation and limit Early correction of sensory deficits should be under-
the use of fentanyl for rapid initiation of analgesia and as a res- taken. That is, eyeglasses and hearing aids should be replaced
cue agent. or fitted (if the patient was not using them before the hos-
More recently, a randomized study suggested that not pitalization) as soon as possible. This will allow patients to
using sedation, compared to the standard treatment with familiarize themselves with the environment and reorient
sedation and a daily wake-up trial, may further reduce the themselves early on. It will also minimize the occurrence
of misperceptions or misinterpretation of environmental
300
yl cues and stimuli. Environmental isolation should be mini-
an
Fe
nt mized if possible. Family members and loved ones should be
Minutes to 50% decrement

240
in plasma concentration

encouraged to visit and provide a familiar and friendly envi-


180 Meperid
ine ronment, as well as provide appropriate orientation and stim-
ulation to patients, especially those with baseline cognitive
120 deficits. Dehydration and electrolyte abnormalities should
e
don be corrected as quickly and safely as possible. Malnutrition
tha
Me
60 Alfentanil
Sufentanil Morphine should be corrected, unless the patient is end-stage and their
Remifentanil
Hydromorphone
delirium is a pre-terminal phenomenon. In that situation,
0
0 120 240 360 480 600
the decision to use a feeding tube, for example, would be a
Infusion Duration
personal decision that different patients and families would
make differently.
Minutes to 50% decrement in plasma concentration of com-
Figure 41.25 For intubated, sedated ICU patients, the approach
monly used opioid agents. must include judicious management of sedative and pain

41. D E L I R I U M • 869
management agents in order to allow for early extubation, PH A R M AC OL O G IC A L M A N AG E M E N T
minimize time in bed, and limit the amount of potentially S T R AT E G I E S
deliriogenic agents a patient is exposed to. Daily cessation of
It cannot be overstated that the “definitive treatment” of
sedative protocols combined with daily spontaneous breath-
delirium is the accurate identification and treatment of its
ing trials and early mobilization (through physical and occu-
underlying cause(s). Nevertheless, pharmacological interven-
pational therapy) have resulted in a significant decrease in the
tion with various psychoactive agents is often needed to help
duration of mechanical ventilation, the length of stay in inten-
manage agitated patients and for the correction of the neu-
sive care, and the number of days patients experienced acute
rotransmitter derangements associated with delirium symp-
brain dysfunction, compared with control groups (Kress,
toms. Of note, no pharmacological agent has received Food
Pohlman, O'Connor, & Hall, 2000; Kress et al., 2003; Girard
and Drug Administration (FDA) approval for the treatment
et  al., 2008; Schweickert et  al., 2009; Needham, Korupolu,
of delirium.
et al., 2010).
In an attempt to “first, do no harm,” we should first
Finally, environmental manipulations to normalize the
identify all agents that may contribute to, exacerbate, or
sleep–wake cycle may be considered, from reducing noise lev-
perpetuate delirium. Chief among these are anticholinergic
els and decreasing nighttime tests and procedures to increas-
substances and GABA-ergic agents (Box 41.2). As described
ing the amount of natural light during daytime hours. Early
above, all such agents (e.g., benzodiazepines, propofol) may
correction of sleep disturbance, preferably by nonpharmaco-
cause or aggravate delirium and its behavioral manifestations
logical means, should be attempted.
(Marcantonio et  al., 1994; Ely et  al., 2001; Pandharipande,
Some have advocated the use of bright light therapy
Shintani, et  al., 2006). The use of benzodiazepines in the
(similar to that used for the treatment of seasonal affec-
management of delirium should be limited to (a)  patients
tive disorder) as an alternative to medication agents but
experiencing delirium related to the withdrawal from a
despite the logic behind this approach, very little evidence
CNS-depressant agent (e.g., alcohol, barbiturates, benzodiaz-
for it exists. A  small pilot study demonstrated that bright
epines), or (b) when other more appropriate agents have failed
light therapy (5000 lux applied for 2 hours in the morning)
and the level of agitation and need for behavioral control out-
led to a significant difference (P = 0.014) in the incidence
weighs the potentially detrimental effects of benzodiazepines.
of delirium (Taguchi, Yano, & Kido, 2007). Yet there were
Following the same principle, we should avoid the use of opi-
serious methodological problems with this study, including
oid agents for behavioral control of agitated patients and use
a very small sample, lack of blinding, and late initiation of
these agents only for pain management, as opioids have been
the intervention (after extubation rather than immediately
implicated in the development of delirium in many patient
after surgery). A  second, larger study (n = 228) conducted
populations (Morrison et al., 2003; Centeno, Sanz, & Bruera,
in a geriatric monitoring unit compared bright light therapy
2004; Morita et  al., 2005; Fong et  al., 2006; Vaurio et  al.,
(2000–3000 lux; 6–10 pm daily) against placebo. The study
2006; Gaudreau et al., 2007, Vella-Brincat & Macleod, 2007;
found that among those randomized to bright light therapy,
Wang et al., 2007). A summary of a comprehensive delirium
there were significant improvements in modified Barthel
management algorithm can be found in Figure 41.26.
Index (MBI) scores, especially for the hyperactive and
Pharmacological management strategies have mostly been
mixed delirium subtypes (P < 0.05); significant improve-
limited to the use of antipsychotic agents, acetyl cholinester-
ments on the DRS sleep–wake disturbance subscore, for
ase inhibitors, and sedative-hypnotic agents. In this section,
all delirium subtypes; and improvements in the mean total
I  will address possible pharmacological interventions based
sleep time (from 6.4 hours to 7.7) (P < 0.05) and length of
on the neurotransmitter being targeted. In general, deliri-
first sleep bout (SB; 6.0 compared with 5.3 hours) (P < 0.05),
ous patients have an excess of dopamine, norepinephrine,
with decreased mean number of SBs and awakenings when
glutamate, and hyperactivity of calcium channels (Ca+ Ch);
compared to those in the placebo group (Chong, Tan, Tay,
reduced levels of acetylcholine and melatonin; and variable
Wong, & Ancoli-Israel, 2013).
levels of serotonin, histamine, and gamma-aminobutyric
The National Institute for Health and Clinical Excellence
acid, depending on the type and cause of delirium. We will
(NICE) provided a set of guidelines containing 13 specific
tailor the discussion regarding pharmacological prophylaxis
recommendations for the prevention of delirium in elderly
and treatment based on these assumptions. Studies will be
at-risk patients (O’Mahony, Murthy et  al., 2011). They are
discussed in order of publication. In the only published set of
mostly based on the correction of modifiable factors that may
guidelines for sedation management of critically ill adults, the
precipitate delirium, and the implementation of the multi-
Society of Critical Care Medicine recommended haloperidol
component intervention package. The full version of these
as “the preferred agent for the treatment of delirium in criti-
recommendations can be found at their website: http://guid-
cally ill patients” (Jacobi, Fraser, et  al., 2002). Accordingly,
ance.nice.org.uk/CG103/Guidance/pdf/English. Similarly,
haloperidol is used by about 78%–80% of intensivists as the
the Society of Critical Care Medicine has developed pain,
standard pharmacological agent to manage delirium in the
agitation, and delirium guidelines, promoted mobility to
ICU, while novel agents, the second-generation antipsychot-
improve care of critically ill patients, and has developed
ics, are gaining popularity for delirium management, are now
tools to facilitate and rapidly implement the translation of
being used by 35%–40% of ICU physicians (Patel, Gambrell,
guideline care recommendations into practice (Davidson
et al., 2009; MacSweeney, Barber, et al., 2010). A more recent
et al., 2013).

870 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
COMPREHENSIVE DELIRIUM ALGORITHM

Delirium Suspected

Non-Rx Delirium Treatment - For ALL Deliria Types:


! Correct malnutrition, dehydration & electrolyte abnormalities
Consider intracranial, (“–“) ! Remove non-essential immobilizing devices & lines
CAM/
metabolic, noxious & ! Review Rx lits and remove potential offending agents
CAM-ICU
psychiatric causes
! Correct sensory deficits
(“+“) ! Implement non-Rx sleep protocol to promote circadian rhythm
(“–“) ! Provide adequate intellectual stimulation– avoid TV
DRS-R-98 ! Address pain- but avoid the use of opioids to control behavior
! Avoid all GABAergic agents & anticholinergic substances
(“+“) ! Continue daily awakening protocol on intubated/sedated patients
Neurology Consult Psychiatry Consult Initiate Pain
& Management & Management Management
Algorithm Algorithm Algorithm Delirium
Management
Protocol Hypoactive Delirium

Initiate Substance
Intoxication or Assess for Substance
Withdrawal Intoxication or
Management (“+“) Withdrawal
Delirium Hypoactive Delirium Algorithm:
Algorithm ! Avoid GABA-ergic & anticholinergic agents
Motoric Type
(“–“) AND
! Add Melatonin 6 mg PO q 2000; or ramelteon 8 mg q 2000
Initiate Anxiety ! Consider ultra-small dose haloperidol (if QTc <500 msec)
Assess for Anxiety (“–“)
Management ! If SGA, use low dose risperidone or aripiprazole (if QTc <500 msec)
(“+“) (GAD-7) ! Consider use of rivastigmine
Algorithm
! In cases of extreme psychomotor retardation consider the use
(“–“) of psychostimulant agents
Monitor
Initiate Depression Hypoactive Delirium ! Obtain daily EKG- D/C antipsychotics if QTc > 500 msec
Assess for
Management ! Use of other Rx that may prolong QTc
(“+“) Depression (PHQ)
Algorithm
(“–“)

Initiate Psychosis Hyperactive Delirium Algorithm:


Management Assess for Psychosis ! Avoid GABA-ergic & anticholinergic agents
Algorithm (“+“)
AND
! Add Melatonin 6 mg PO q 2000; or ramelteon 8 mg q 2000
! Consider the use of alpha-2 agonist agents for sedation/agitation
! Consider moderate dose haloperidol (if QTc <500 msec)
! If SGA, use low dose risperidone or quetiapine (if QTc <500 msec)
! Consider use of rivastigmine, If Hx of cognitive deficit or dementia, or in protracted cases
! In cases of extreme psychomotor retardation consider the use of psychostimulant agents
Monitor
! Obtain daily EKG-D/C antipsychotics if QTc > 500 msec
! Use of other Rx that may prolong QTc

Figure 41.26 Comprehensive Delirium Algorithm. Source: (Maldonado, 2008a, 2011)

“post-publication perspective” of practice guidelines added no the treatment group (10.5% vs. 32.5%; p <.05). Similarly, the
new information about pharmacological agents for the man- intensity and duration of postoperative delirium were more
agement of delirium in the ICU, but emphasized the benefits severe and lasted longer in the control group.
of early detection and patient mobilization. It is important The first double-blind, randomized, placebo control stud-
to point out that soon after the publication of these perspec- ies (DBRPCT) compared haloperidol to placebo, also in
tives a number of significant studies were published, all con- at-risk elderly orthopedic patients (Kalisvaart, de Jonghe,
clusively demonstrating the delirium-prophylactic benefits of et al., 2005). Subjects received oral haloperidol 0.5 mg/d up to
antipsychotic agents in at-risk populations, which will be dis- 72 hours preoperatively and then until the third postoperative
cussed in the next section. day. The study found that prophylactic haloperidol use did not
Of note, no pharmacological agent has received FDA alter the incidence of postoperative delirium when compared
approval for the prevention of delirium. with placebo (15.1 vs.16.5%; p = ns). Nevertheless, it did have
a positive effect on the severity (the mean highest DRS-R-98
score ±SD was 14.4 ± 3.4 vs. 18.4 ± 4.3; P <.001) and dura-
Dopamine Excess
tion (5.4 vs. 11.8 days; P <.001) of delirium and shortened the
Dopamine Antagonists for Delirium Prevention length of hospital stay (mean number of days in the hospital
There are very few control studies on the efficacy of antipsy- was 17.1±11.1 vs. 22.6 ± 16.7; P <.001).
chotic agents for the prevention of delirium, yet most studies Many additional studies have followed, all with positive
suggest a positive effect (see Table 41.12). The first random- findings. A  study of patients undergoing cardiac surgery
ized controlled trial compared haloperidol to placebo in with cardiopulmonary bypass (CPB), randomly assigned to
at-risk, elderly orthopedic patients (Kaneko, Jianhui, et  al., receive either 1 mg of risperidone or placebo sublingually
1999). The incidence of delirium was significantly lower in as soon as they regained consciousness postoperatively,

41. D E L I R I U M • 871
Table 41.12 ANTIPSYCHOTIC AGENTS FOR PR EVENTION OF DELIR IUM

DELIR IUM INCIDENCE%


STUDY(N = 9) POPULATION INTERVENTION DELIR IUM DEFINITION CONTROL/INTERVENTION P-VALUE

Kaneko et al., 1999; RPCT GI surgery Prophylaxis haloperidol vs. PBO IV DSM-III-R 32.5 10.5 p < 0.05
postoperatively for 5 days

Kalisvaart et al., 2005 Elderly hip-replacement Sx PBO vs. haloperidol 1.5 mg/d started DSM-IV 16.5 15.1 p = 0.91
DBRPCT, n = 430 preop, continued for up to 3 days postop. CAM (36/216) (32/212)
DRS-R98

Prakanrattana & Cardiac Sx under CPB PBO vs. sublingual risperidone CAM 31.7 11.1 p = 0.009
Prapaitrakool, 2007 immediately p-Sx (20/63) (7/63)
DBRPCT, n = 126

Girard et al., 2010* Med/Surg ICU in PBO vs. haloperidol vs. ziprasidone: days CAM-ICU 12.5 14.0 15.0 p = 0.66
DBRPCT, n = 101 mechanical ventilation alive without delirium or coma; [1.2–17.2] [6.0–18.0] [9.1–18.0]
conducted in 6 tertiary medical centers. days days days

Larsen et al., 2010 Elderly elective total PBO vs. 5 mg of orally disintegrating DSM-III-R 40.2 14.3 p < 0.001
DBRPCT, n = 495 joint-replacement olanzapine 1-dose pre- and 1-dose (82/204) (28/196)
post-surgery

Wang et al., 2012 Elderly, non-cardiac Sx PBO vs. HAL (0.5 mg bolus, followed by CAM 23.2 15.3 p = 0.031
DBRPCT, n = 457 cont infusion 0.1 mg/hr x 12hrs)

van den Boogaard et al., ICU patients at “high risk” PBO vs. HAL (1 mg/8 hr.) within 24 hrs. CAM-ICU 75 65 P = 0.01
2013, 2013)DBRPCT, n = of admission to ICU
476

Hirota & Kishi, 2013; Various clinical settings Meta-analysis of 6 studies (3 HAL, Various tools Sensitivity analysis showed that SGA P < 0.00001
Meta-analysis (RCTs), 6 1 olanzapine, 2 risperidone) using were superior to PBO (NNT = 4; p <
studies, n = 1689 antipsychotic agent for delirium 0.0001), whereas HAL failed to show
prophylaxis superiority to PBO
Teslyar et al., 2013; Postoperative elderly Medication administered included Various tools The pooled relative risk of the five P < 0.01
Meta-analysis (RCTs); 5 patients haloperidol (3), risperidone (1), and studies resulted in a 50% reduction
studies, n = 1491 olanzapine (1). in the relative risk of delirium
among those receiving antipsychotic
medication compared with placebo.
* Underpowered.
Abbreviations: RPCT=randomized placebo control trial; GI=gastrointestinal; PBO=placebo; DBRPCT=double blind randomized, placebo controlled trial; DSM=Diagnostic and Statistical Manual of Mental Disorders;
CAM= Confusion Assessment Method; DRS=Delirium Rating Scale; Sx=surgery; CPB=cardio-pulmonary by-pass machine; ICU=intensive care unit. APA=antipsychotic agent; OLA=olanzapine; HAL=haloperidol;
RIS=risperidone
demonstrated a significantly lower incidence of postop- were reported, including torsades de point (TdP) (van den
erative delirium in the risperidone group (11.1% vs. 31.7%; Boogaard et al., 2013).
P = 0.009) (Prakanrattana & Prapaitrakool, 2007). A study Two meta-analyses of studies using dopamine antagonist
examined the effects of olanzapine pretreatment (5 mg agents for delirium prophylaxis found that pooled relative risk
Zydis® formulation administered just preoperatively, and of published studies suggested a 50% reduction in the relative
again 5 mg administered immediately after surgery upon risk of delirium among those receiving antipsychotic medica-
awakening) to placebo in elderly subjects undergoing ortho- tion compared with placebo (p <0.01). The studies suggest
pedic joint-replacement surgery. The incidence of delirium that perioperative use of prophylactic dopamine antagonist
in the intervention group was significantly lower in the agents (both typical and second-generation antipsychotics),
study group compared to placebo (15% vs. 41%; P < 0.0001) compared to placebo, may effectively reduce the overall risk
(Larsen, 2007). of postoperative delirium, thereby potentially reducing mor-
A large (n = 457; ages ≥65) postoperative delirium pre- tality, disease burden, length of hospital stay, and associated
vention study compared the efficacy and safety of low-dose healthcare costs (Hirota & Kishi, 2013; Teslyar, Stock, et al.,
IV-haloperidol (i.e., 0.5 mg intravenous bolus injection, fol- 2013). A  third meta-analysis, including all postoperative
lowed by continuous infusion at a rate of 0.1 mg/hr for 12 hrs prevention strategies, found that both typical (3 RCTs; n =
vs. placebo) vs. placebo in critically ill patients after noncar- 965; RR = 0.71; 95% CI = 0.54–0.93) and atypical antipsy-
diac surgery. The study found the incidence of delirium dur- chotics (3 RCTs; n = 627; RR = 0.36; 95% CI = 0.26–0.50)
ing the first 7 days after surgery was 15.3% in the haloperidol decreased delirium occurrence, compared to placebos (Zhang
group vs. 23.2% in the control group (p = .031). Other mea- et al., 2013).
sures also appeared significantly improved:  the mean time
to onset of delirium and the mean number of delirium-free Dopamine Antagonists for Delirium Treatment
days were significantly longer (6.2  days [95% CI 5.9–6.4] The literature has long recognized that intravenous neuro-
vs. 5.7 days [95% CI 5.4–6.0]; p = .021; and 6.8 ± 0.5 days leptic agents are the recommended emergency treatment for
vs. 6.7 ± 0.8 days; p = .027, respectively); the median length agitated and mixed type delirium (Adams, Fernandez,  &
of ICU stay was significantly shorter (21.3 hrs [95% CI, Andersson, 1986; Fernandez, Holmes, Adams, & Kavanaugh,
20.3–22.2] vs. 23.0 hrs [95% CI, 20.9–25.1]; p = .024) in the 1988; Sanders, Murray,  & Cassem, 1991; Ziehm, 1991;
haloperidol group than in the control group. There was no Riker, Fraser, & Cox, 1994; Inouye et al., 1999). Intravenous
significant difference with regard to all-cause 28-day mortal- administration of haloperidol has always been thought supe-
ity between the two groups (0.9% [2/229] vs. 2.6% [6/228]; rior to oral administration because the IV route has more
p = .175). No drug-related side effects were documented reliable absorption, even in cases of systemic organ failure.
(Wang et al., 2012). Intravenous haloperidol use has the added advantage of
Finally, a study assessed the efficacy of haloperidol pro- requiring no patient cooperation, thus facilitating its use even
phylaxis against delirium in “at risk” ICU patients (n = in uncooperative and agitated patients. Studies suggest that
177) (van den Boogaard, Schoonhoven, van Achterberg, van the IV use of high-potency neuroleptic agents is associated
der Hoeven,  & Pickkers, 2013). The results of the interven- with minimal effects on blood pressure, respiration, and heart
tion were compared to a historical control group and a con- rate (Ayd, 1978; Tesar, Murray,  & Cassem, 1985; Adams,
temporary group that did not receive haloperidol prophylaxis, 1988; Sanders et  al., 1991; Riker et  al., 1994; Stern, 1994;
mainly due to noncompliance to the protocol, mostly during Segatore & Adams, 2001). And, some have suggested that hal-
the implementation phase. The protocol consisted of low-dose operidol has a lower incidence of extrapyramidal symptoms
haloperidol infusion (i.e., IV-haloperidol 1 mg/8 hr (or a when administered IV versus orally (7.2% vs. 22.6%; p <0.01)
lower dose, for patients ≥80 years), started as soon as patients (see Figure  41.27) (Sanders, Minnema,  & Murray, 1989;
were identified to be at increased risk, within the 24-hour Maldonado & Kang, 2003).
period after ICU admission. All patients experiencing delir- Intravenous haloperidol has been identified as the agent of
ium received therapeutic doses of haloperidol according to choice for the management of delirious critically ill patients by a
the unit’s customary protocol. Patient characteristics were
comparable between the prevention and the control groups;
with the exception of prophylactic haloperidol, as per study 100%
protocol. The intervention resulted in a lower delirium inci-
80%
dence (65% vs. 75%, p =.01), more delirium-free days (median,
20  days [interquartile range 8 to  27] vs. median 13  days 60% PO
[3  to  27], p =.003), fewer ICU readmissions (11% vs. 18%, 40%
IV
p =.03), and less frequent unplanned removal of tubes/lines 23%
7%
(12% vs., 19%, p =.02) in the intervention group compared to 20%
the control group. Beneficial effects of haloperidol appeared 0%
most pronounced in the patients with the highest risk for EPS
delirium. Haloperidol was stopped in 12 patients because of
Haloperidol Route of Administration and Incidence of EPS.
Figure 41.27
QTc-time prolongation (n = 9), renal failure (n = 1) or sus- Rate of extrapyramidal symptoms and their association to mode of
pected neurological side-effects (n = 2). No other side effects administration for haloperidol. (Maldonado, 2003)

41. D E L I R I U M • 873
number of national organizations, including Britain’s National over the last few years for the management of psychiatric and
Institute for Health and Clinical Excellence (NICE, 2010) the behavioral symptoms (e.g., agitation, psychosis, delirium) in
American Psychiatric Association (APA, 1999, 2004)  and medically ill patients. Large studies, particularly head-to-head
the Society of Critical Care Medicine (Shapiro et  al., 1995; comparisons between SGAs and more conventional agents
Jacobi et  al., 2002). Since then, a “best evidence topic in car- (i.e., haloperidol), are lacking; therefore, most clinical prac-
diac surgery” was written according to a structured proto- tice is derived from case reports and open-label (OL) studies.
col, addressing the issue of haloperidol safety for critically ill Following are summaries of the only seven RCTs available; all
patients (Khasati, Thompson, & Dunning, 2004). Their search others are summarized in Table 41.13.
included 294 papers and concluded that haloperidol should be The first study was a DBRCT that compared haloperidol
considered the first-line drug for agitated patients following (average. dose 1.4 mg/d) vs. chlorpromazine (average dose
cardiac surgery. Of note, in September of 2007, the FDA issued 36.0 mg/d) versus lorazepam (avg. dose 4.6 mg/d) among
a “black-box” warning for the off-label clinical practice of using HIV/AIDS patients (Breitbart et al., 1996). Results showed
IV-haloperidol (FDA, 2007). It is important to remember that that treatment with either haloperidol or chlorpromazine in
haloperidol has never been approved by the FDA for IV use. relatively low doses resulted in significant improvement in the
Other antipsychotic agents have been studied for the symptoms of delirium, while no improvement was found in
treatment of delirium. Most of these include atypical or the lorazepam group. Treatment with either neuroleptic was
second-generation antipsychotic (SGA) agents. Because of associated with an extremely low prevalence of extrapyrami-
the stigma and potential side effects associated with typical dal side effects; all patients receiving lorazepam, however,
antipsychotics, SGAs have being been used at increasing rates developed treatment-limiting adverse effects.

Table 41.13 DOPAMINE ANTAGONIST—ANTIPSYCHOTIC AGENTS AS TR EATMENT OF DELIR IUM

DELIR IUM
DEFINITION/
MEASUR E OF
STUDY(N = 32) POPULATION INTERVENTION R ESPONSE R ESULTS

Breitbart et al., AIDS, medical Haloperidol vs. DSM-IIIR/DRS Tx either HAL or CPM resulted in
1996; DBRCT, patients chlorpromazine vs. significant improvement in the symptoms
n = 30 lorazepam of delirium, while no improvement was
found in the LOR group. Tx neuroleptic was
associated with an extremely low prevalence
of EPS, while all patients receiving LOR
developed treatment-limiting adverse effects.

Sipahimalani & Med/Surg patients Haloperidol vs. DRS Improvement was similar in both groups
Masand, 1998; OL, olanzapine (mean DRS +SD H = 11.1 ± 7.1; O = 10.3 ±
n = 22 4.8; P = 0.760), with extrapyramidal
symptoms found only in haloperidol
patients. No side effects in olanzapine
group.

Schwartz et al., Med/Surg patients Quetiapine vs. DRS Effectiveness of 350% in reducing DRS
2000; Single-blind; haloperidol— scores. When compared with haloperidol,
n = 11 retrospective chart there was no difference in onset of symptom
review resolution, duration of treatment, and
overall clinical improvement.

Kim et al., 2001; Med/Surg patients Olanzapine PO, DSM-IV/DRS 50% decrease in Delirium Rating Scale
OL, n = 20 variable dose scores (from pre of 20.0 ± 3.6, to post
of 9.3 ± 4.6; P < 0.01). No side effects,
including EPS.

Breitbart et al., Hospitalized Olanzapine PO, DSM-IV/MDAS Olanzapine was effective in treating 76%
2002; OL, n = 79 cancer patients variable dose of delirium patients as evidenced by the
MDAS; caused excessive sedation in 30%
of patients.

Horikawa et al., Med/Surg patients Risperidone PO DSM-IV/DRS At a low dose of 1.7 mg/d, on average,
2003; OL, n = 10 risperidone was effective in 80% of
patients, and the effect appeared within
a few days. Most commonly cited adverse
effects included sleepiness (30%) and mild
drug-induced Parkinsonism (10%).
(continued)

874 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Table 41.13 (CONTINUED)

DELIR IUM
DEFINITION/
MEASUR E OF
STUDY(N = 32) POPULATION INTERVENTION R ESPONSE R ESULTS

Sasaki et al., 2003; Med/Surg patients Quetiapine PO, DSM-IV/DRS 100% of patients on quetiapine achieved
OL, n = 12 flexible doses resolution of delirium (mean on day
4.8 ± 3.5 days); no EPS reported.

Kim et al., 2003; Elderly medical Quetiapine PO, DSM-IV/DRS 100% of patients on quetiapine achieved
OL, n = 12 inpatients flexible doses resolution of delirium by day 10 (mean
on day 5.9 ± 2.2 days); no EPS reported.
Delirium Rating Scale scores along
with scores of the Mini-Mental State
Examination and Clock Drawing Test
continued to improve throughout the
3-month study period.

Liu et al., 2004; Med/Surg patients Risperidone DSM-IV/various Patients treated with haloperidol were
retrospective record with hyperactive (average dose younger than patients treated with risperidone
review, n = 77 delirium 1.17±0.76 mg/d) vs. (P < 0.05). The mean hyperactive syndrome
haloperidol (avg. scale score was higher in the haloperidol
dose 4.25±2.62 than that of the risperidone group. The
mg/d) authors found no significant difference in the
efficacy or frequency of response rate between
haloperidol and risperidone (100% vs. 95%;
p = ns). Patients on risperidone experienced
less EPS (7% vs. 69%).

Mittal et al., 2004; Patients admitted to Risperidone, 0.5 mg DSM-IV/DRS Rapid resolution of delirium while
OL, n = 10 Med-Surg unit PO BID, flexible receiving low dose
PRNs risperidone (mean dose 0.75 mg/d); no EPS
reported.

Parellada, Baeza, Prospective, Risperidone PO DSM-IV/ Risperidone was administered at the time
et al., 2004; OL multicenter, DRS of diagnosis, and treatment was maintained
observational PANSS-P according to clinical response. Found
7-day study MMSE a significant decrease in DRS scores in
90.6% of treated patients and significantly
improved all symptoms measured by the
scales from baseline to day 7 (p < 0.0); only
3% side effects.

Pae, Lee, et al., 2004 Med/Surg patients Quetiapine PO DRS-R9/ DRS-R-98 and clinical global scores were
OL, n = 22 CGI8 significantly reduced by 57.3% and 55.1%,
respectively. Quetiapine was effective
and safe.

Han et al., 2004; Med/Surg patients Haloperidol vs. CAM/ Both groups showed significant improvement
DBRCT, n = 28 risperidone, 7 days DRS in baseline DRS and MDAS scores with
medication trial MDAS either haloperidol (75%) or risperidone (42%)
(P < 0.05). There was no significant difference
in improvement of DRS (p = 0.35) or MDAS
(p = 0.51) scores, comparing haloperidol with
risperidone patients.

Hu et al., 2004; Med/Surg elderly Haloperidol vs. DRS/ All groups showed a decrease in DRS scores
RPCT, n = 175 patients olanzapine vs. CGI by 7th day, compared with baseline (p <
placebo, 7 days 0.01). Decrease in DRS scores of treated
medication trial patients at day 7 (OLA 72.2%; HAL 70.4%)
differed significantly from DRS scores of
PBO patients (29.7%; (p < 0.01), but not
from each other (p > 0.05).
(continued)
Table 41.13 (CONTINUED)
DELIR IUM
DEFINITION/
MEASUR E OF
STUDY(N = 32) POPULATION INTERVENTION R ESPONSE R ESULTS

Skrobik et al., 2004; Critically ill Med/ Haloperidol (avg. DIS ICU Delirium Index Screening Checklist
OL-prospective RCT, Surg patients 6.5 mg/d) vs. Scores were reduced in both groups,
n = 73 olanzapine (avg. 4.5 compared with baseline (p < 0.05), but there
mg/d) was no significant difference in DIS scores
between active Tx groups (p = 0.9). EPS
were found in 13% of haloperidol patients,
but 0% in olanzapine group.

Toda et al., 2005; n Elderly inpatient Risperidone, OL, DSM-IV/ Resolution reported in 7 subjects (mean
= 10 general medicine 0.5 mg oral sol.; DRS dose 0.92 ± 0.47 mg/d); 1 non-responder;
flexible titration, side effects requiring Tx discontinuation
PRN reported in 2.

Lee et al., 2005; Med/Surg patients Amisulpride vs. DRS-R98/ After treatment, DRS-R-98 scores were
RCT, n = 40 quetiapine CGI significantly decreased from the baseline
in both treatment groups (P < 0.001)
without group difference. Both atypical
antipsychotics were generally well tolerated.

Straker et al., 2006 Medically ill patients Aripiprazole PO DSM-IV/ 50% of patients had improved significantly
OL, n = 14 was used in a flexible DRS-R98 by day 5, as indicated by a 50% reduction
dosing range, from CGI in DRS-R-98 scores. 86% of patients had a
5 mg/day to 15 mg/ 50% reduction in their DRS-R- 98 scores
day, titrated as by end of treatment. Mean CGI Severity
clinically indicated scores at the beginning of treatment were
5.2, with a mean CGI Improvement score
after treatment of 2.1, indicating much
improvement.

Takeuchi et al., 2007; Med/Surg patients Perospirone, OL DSM-IV/ Perospirone was effective in 86.8% of
OL, n = 38 DRS-R98 patients, within several days (5.1 ±
4.9 days). The initial dose was 6.5 ± 3.7 mg/
day and maximum dose of perospirone was
10.0 ± 5.3 mg/day. There were no serious
adverse effects.

Maneeton, Medically ill patients Quetiapine, flexible CAM/DRS, CGI 88% subjects responded. Means (SDs) dose
Maneeton, & dosing and duration (SD) of quetiapine treatment
Srisurapanont, 2007; were 45.7 (28.7) mg/day and 6.5 (2.0) days,
OL, n = 17 respectively. The DRS and CGI-S scores of
days 2–7 were significantly lower than those
of day 0 (p < 0. 001) for all comparisons).
Only two subjects were shown to have mild
tremor.

Reade et al., 2009; Med/Surg ICU Agitated delirium ICDSC DEX significantly shortened median
OL-RCT were randomized to Time time to extubation from 42.5 to 19.9 hours
receive an infusion (P = 0.016); it significantly decreased
of either haloperidol ICU length of stay, from 6.5 to 1.5 days
0.5 to 2 mg/hour or (P = 0.004); and of patients requiring
dexmedetomidine ongoing sedation, it reduced the time
0.2 to 0.7 μg/kg/hr propofol was required in half (79.5% vs.
41.2%; P = 0.05).

Devlin et al., 2010; MICU PBO vs. quetiapine ICDSC Tx w QUE was associated with: a shorter
DBRPCT, n = 36 (50 mg BID) time to first resolution of delirium
(multienter-3) (p =.001), a reduced duration of delirium
(p =.006), less agitation (p =.02), greater
chance to be discharged home vs.
long-term-care facility (p =.06), and lower
requirement of as-needed haloperidol
(p =.05).
(continued)
Table 41.13 (CONTINUED)

DELIR IUM
DEFINITION/
MEASUR E OF
STUDY(N = 32) POPULATION INTERVENTION R ESPONSE R ESULTS

Girard et al., 2010; Mechanically PBO vs. HAL vs. CAM-ICU Patients in the haloperidol group spent
DBRPCT, n = 101 ventilated Med/Surg Ziprasidone a similar number days alive without
ICU patients delirium or coma (14.0 [6.0 –18.0] days) as
did those on ziprasidone (15.0 [9.1–18.0]
days) and PBO groups (12.5 [1.2–17.2] days;
p = 0.66).

Kim et al., 2010; Elderly, Med/Surg Risperidone vs. DSM-IV/DRS-r-98 Significant within-group improvements
SB-RCT, n = 32 patients olanzapine in the DRS-R-98 scores over time were
observed at every time point in both
treatment groups. The response rates
did not differ significantly between the
two groups (risperidone group: 64.7%,
olanzapine group: 73.3%); and there were
no difference in the safety profiles and side
effects between groups.
Tahir et al., 2010; Med/Surg patients Quetiapine vs. DSM-IV/DRS-r-98, Quetiapine has the potential to more
DBRCT, n = 42 placebo CGI quickly reduce the severity of non-cognitive
aspects of delirium. The study was
underpowered for treatment comparisons.

Grover et al., 2011; Med/Surg patients Haloperidol DSM-IV/DRS-r-98 Patient in all three groups experience a
Prospective, single (0.25 to 10 mg) significant reduction in DRS-R98 severity
blind, n = 64 vs. olanzapine scores and a significant improvement in
(1.25 to 20 mg) vs. MMSE scores over the period of 6 days;
risperidone (0.25 to with no difference between the treatment
4), flexible dosing groups. Rate of side effects was also similar.

Boettger et al., 2011; Med/Surg patients at Aripiprazole, DSM-IV/MDAS The mean dosage of aripiprazole required
OL, n = 21 Cancer Center flexible dosing was 18.3 mg (range of 5–30) daily at
T3. Patients treated for delirium with
aripiprazole experienced significant
improvement and resolution of delirium,
with MDAS scores declining from a mean
of 18.0 at baseline (T1) to mean of 10.8 at
T2 and a mean of 8.3 at T3. There was a
100% resolution of hypoactive delirium,
compared to patients with hyperactive
delirium (58.3%).

Hakim et al., 2012; Patients aged Randomized using a ICDSC Seven (13.7%) patients in the risperidone
PCRCT 65 yrs. or older computer-generated group experienced delirium vs. 17
who experienced list to receive (34%) in the placebo group (p = 0.031).
subsyndromal placebo (n = 50) or Competing-risks regression analysis showed
delirium after 0.5 mg risperidone that failure to treat subsyndromal delirium
on-pump cardiac (n = 51) every 12 hrs with risperidone was an independent
surgery by mouth risk factor for delirium (p = 0.002). Two
(3.9%) patients in the risperidone group
experienced extrapyramidal manifestations
versus one (2%) in the placebo group
(p = 1.0).

Kishi et al., 2012; Adult delirious Risperidone given DRS-R98 Entry DRS-R-98 score = 19.8 ± 6.8;
OL, n = 29 cancer patients orally once per day 7-d follow-up score = 14.3 ± 7.8, results
(mean dosage, 1.4 ± demonstrate DRS-R98 scores were
1.3 mg/day improved in 79.3% of patients during the
study period (p < 0.001); 38% achieved
remission (i.e., DRS-R-98 ≤10).
(continued)
Table 41.13 (CONTINUED)

DELIR IUM
DEFINITION/
MEASUR E OF
STUDY(N = 32) POPULATION INTERVENTION R ESPONSE R ESULTS

Tagarakis et al., Consecutive patients Ondansetron IV (8 Self-developed rating Noted a statistically significant
2012; n = 80 who developed mg) vs. haloperidol scale: 0–4 improvement in the test score rating after
post-op delirium IV (5 mg); patients the administration of both ondansetron
after on-pump heart were evaluated (from 3.1 to 1.2, percentage improvement
surgery before and 10 min 61.29%, p < 0.01) and haloperidol (from
after the injection 3.1 to 1.3, ± percentage improvement
58.064%, p < 0.01).

Yoon et al., 2013; Patients with Assigned to receive Korean version of Haloperidol, risperidone, olanzapine, and
Observational study; delirium at a tertiary either haloperidol the Delirium Rating quetiapine were equally efficacious and safe
n = 80 level hospital (n = 23), risperidone Scale–Revised-98 in the treatment of delirium. The treatment
(n = 21), olanzapine (DRS-K) response rate was lower in patients over
(n = 18), or 75 years old than in patients under 75 years
quetiapine (n = 18) old, especially for olanzapine.
Maneeton et al., Medically ill patients 25–100 mg/day of DRS-R-98 and total Over the trial period, means (standard
2013; DBRCT, n with delirium quetiapine sleep time deviation) of the DRS-R-98 severity scores
= 52 (n = 24) or were not significantly different between
0.5–2.0 mg/day of the quetiapine and haloperidol groups
haloperidol (n = 28) (–22.9 [6.9] versus –21.7 [6.7]; P = 0.59).
Concluding that low-dose quetiapine and
haloperidol may be equally effective and
safe for controlling delirium symptoms.
Abbreviations: RPCT=randomized placebo control trial; DBRPCT=double blind randomized, placebo controlled trial; OL=open label; trial; DSM=Diagnostic
and Statistical Manual of Mental Disorders; CAM= Confusion Assessment Method; DRS=Delirium Rating Scale; MDAS=Memorial Delirium Assessment Scale;
GI=gastrointestinal; PBO=placebo; Sx=surgery; CPB=cardio-pulmonary by-pass machine; ICU=intensive care unit. APA=antipsychotic agent; OLA=olanzapine;
HAL=haloperidol; RIS=risperidone; AIDS=autoimmune deficiency syndrome; CPM=chlorpromazine; EPS= extra-pyramidal syndrome; LOR=lorazepam; PO=
by mouth.

The second included a head-to-head comparison between quetiapine was associated with a shorter time to first resolu-
haloperidol and an SGA (risperidone), demonstrating that tion of delirium (1.0 vs. 4.5  days; p =.001), a reduced dura-
delirium scores decreased significantly over the study period tion of delirium (36 vs. 120 hrs.; p =.006), less agitation based
with no significant differences between groups (Han & Kim, on the Sedation-Agitation Scale score (6 vs. 36 hrs.; p =.02),
2004). An RPCT 7-day treatment trial compared haloperidol greater chance to be discharged home versus long-term-care
to olanzapine versus placebo among hospitalized elderly sub- facility (89% vs. 56%; p =.06), and lower requirement of PRN
jects, again showing no difference in the rate of improvement haloperidol (3 vs. 4 days; p =.05).
of delirium with either active agent, but significantly greater The most recent published blinded RCT compared que-
for each than placebo (olanzapine 72.2%; haloperidol 70.4%; tiapine (flexible dosing regimen of 25 to 175 mg per day,
placebo 29.7%; p <0.01, against PBO) (Hu, Deng,  & Yang, in divided doses) versus placebo among medico-surgical
2004). Another study compared enteral haloperidol (avg. dose patients (Tahir et al., 2010). Results indicated the quetiapine
6.5 mg/d) versus olanzapine (avg. dose 4.5 mg/d) in the criti- group improved more rapidly, 82.7% faster (standard error
cal care setting (Skrobik, Bergeron, Dumont,  & Gottfried, [S.E.] 37.1%, P =.026) than the placebo group, in terms of
2004). Results show that measures of delirium decreased over the DRS-R-98 severity score; and 57.7% faster (S.E. 29.2%,
time in both groups, as did the administered dose of as-needed P =.048) than the placebo group, in terms of the noncogni-
(PRN) benzodiazepines, with the degree of clinical improve- tive subscale.
ment being similar in both treatment arms. A Cochrane Database review compared haloperidol with
An RCT comparing the antipsychotics amisulpride versus risperidone, olanzapine, and placebo in the management of
quetiapine showed that after intervention, DRS-R-98 scores delirium and concluded that there was no significant difference
were significantly decreased from the baseline in both treat- between low-dose haloperidol (<3.0 mg per day) and the atypi-
ment groups (P <0.001) without group difference, and that cal antipsychotics olanzapine and risperidone (OR 0.63, p =
both SGAs were generally well tolerated (Lee, Won, et  al., 0.25); that low-dose haloperidol did not have a higher incidence
2005). A  study on the use of SGA involved the compari- of adverse effects than the atypical antipsychotics; and that
son of quetiapine (50 mg twice daily [BID]) versus placebo low-dose haloperidol was effective in decreasing the intensity
in medical ICU patients in three academic centers (Devlin, and duration of delirium in postoperative patients, compared
Roberts, et  al., 2010). The study found that treatment with with placebo (Lonergan, Britton, Luxenberg, & Wyller, 2007).

878 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
A subsequent meta-analysis on atypical antipsychotic 3. Evidence does not indicate major differences in response
agents for the treatment of delirium found that risperidone rates between clinical subtypes of delirium; and
was the most thoroughly studied SGA for the management
4. There are no significant differences in efficacy for halo-
of delirium (Ozbolt, Paniagua, & Kaiser, 2008). Most studies
peridol versus atypical agents (Meagher et al., 2013).
found that risperidone was approximately 80%–85% effec-
tive in treating the behavioral disturbances of delirium at
doses of 0.5–4.0 mg per day. On the other hand, olanzapine Regarding side-effect incidence, a meta-analysis of all
was approximately 70%–76% effective in treating the behav- randomized controlled trials in which SGAs have been com-
ioral manifestations of delirium at doses of 2.5–11.6 mg per pared with conventional drugs concluded that of the SGAs,
day. There were very few studies conducted using quetiapine; only clozapine was associated with significantly fewer extra-
although available data suggest that it also appears to be a safe pyramidal symptoms (EPS (p = 0.008) and higher efficacy
and effective alternative to high-potency antipsychotics. In than low-potency conventional drugs. A  reduced frequency
the limited number of trials comparing SGAs to haloperidol, of EPS was noted with olanzapine, which was of border-
haloperidol consistently produced a higher rate (an additional line significance (p = 0.07) (Leucht, Wahlbeck, Hamann, &
10%–13%) of extrapyramidal side effects. Kissling, 2003). Another study analyzed the data on neu-
Finally, there have been two case reports on the use of roleptics acquired in the German multicenter drug surveil-
aripiprazole as an effective treatment of delirium (Alao  & lance program database (Projekt “Arzneimittelsicherheit in
Moskowitz, 2006; Straker, Shapiro, & Muskin, 2006); and a der Psychiatrie”) program from 1993 to 2000 and found that
retrospective case-matched study of aripiprazole vs. haloperi- severe adverse drug reactions (ADRs) occurred in 1.1% of the
dol that found an identical delirium resolution rate (76.2%) patients (Bender, Grohmann, et al., 2004). In contrast to the
for both agents (Boettger, Friedlander, Breitbart,  & Passik, results from controlled trials, atypical neuroleptics caused
2011). In addition, there are two single case reports on the more severe ADRs than did typical neuroleptics. Overall,
use of ziprasidone for the management of delirium (Leso & SGAs were found to be superior in EPS and urological ADRs,
Schwartz, 2002; Young & Lujan, 2004) and one OL study (n but when clozapine is excluded, typicals and SGAs have similar
= 38) on the use of perospirone, a recently developed atypical occurrence rates of severe ADRs. Other problems to consider
antipsychotic with potent serotonin 5-HT2 and dopamine when choosing an alternative agent include the fact that SGAs
D2 antagonist activity (Takeuchi et al., 2007). may be associated with weight gain, dyslipidemia, high blood
A systematic literature review of 28 delirium treatment pressure, and ultimately with cardiovascular disease, diabetes,
studies with antipsychotic agents concluded that: and metabolic syndrome (Henderson, 2008). Therefore, when
considering the use of SGAs, clinicians must consider these
factors and weigh potential risks and benefits before prescrib-
1. Around 75% of delirious patients who receive short-term
ing these agents to a critically ill patient. Finally, there have
treatment with low-dose antipsychotics experience clini-
been numerous reports attributing the onset of delirium to the
cal response;
use of SGAs (e.g., clozapine, olanzapine), probably due to their
2. This response rate appears quite consistent across differ- anticholinergic potential (Bender, Grohmann, et  al., 2004).
ent patient groups and treatment settings; But these data are limited to small case reports (see Box 41.7).

Box 41.7 CASE R EPORTS LINKING SGA TO DELIRIUM GENER ATION

• Case reports implicating risperidone use in causing delirium:


• Kato et al. (2005), Psychosomatics
• Morikawa and Kishimoto (2002), Canadian Journal of Psychology
• Tavcar and Dernovsek (1998), Canadian Journal of Psychology
• Chen Chen and Cardasis (1996), American Journal of Psychology
• Case reports implicating quetiapine use in causing the delirium:
• Sim, Brunet, and Conacher (2000), Canadian Journal of Psychology - Balit et al. (2003), Annals of Emergency Medicine
• Miodownik et al. (2008), Clinical Neuropharmacology
• Case reports implicating olanzapine use in causing the delirium:
• Robinson, Burk, & Raman (2003), Journal of Postgraduate Medicine
• Steil (2003) Der Nervenartzt
• Samuels and Fang (2004), Journal of Clinical Psychology - Morita et al. (2004) Journal of Pain & Symptom Management
• Prommer (2005) Journal of Pain & Symptom Management
• Tuglu, Erdogan, & Ebay (2005) Journal of Korean Medicine
• Weizberg et al. (2006) Clinical Toxicology
• Lim, Travino, & Tampi (2006), Annals of Pharmacotherapy

41. D E L I R I U M • 879
Please note that antipsychotic agents should only be used are to discontinue antipsychotic use if QTc increases to
short-term for treating agitation in the cognitively impaired >25% of baseline or >500 msec. Of note, studies suggest
individual. Serious concerns have been raised about the that haloperidol may have the lowest ratio of cardiac death
stroke and mortality risk of atypical antipsychotics when among all dopamine-antagonist agents, both typical and
administered long-term for the management of agitation atypical (Hatta, Takahashi, et  al., 2001; Harrigan, Miceli,
in patients with dementia. The dementia antipsychotic et al., 2004).
withdrawal trial (DART-AD) found that at 12 months, the Our recommended schedule in the case of hyperactive states
cumulative probability of survival was similar in the active is to use antipsychotic agents (e.g., haloperidol, risperidone, que-
group and the placebo group (70% versus 77%; p  =  ns) tiapine) at one of the following time schedules: “TID” (thrice
(Ballard, Hanney, et  al., 2009). Yet, there was a reduction daily) schedule (0600 hrs, 1000, 2000) or “QID” (four times a
in survival of the patients who continued to receive antipsy- day) schedule (0600, 1000, 1600, 2200). The idea is to enhance
chotics (after 12 months) compared with those who received behavioral control during the daytime and help promote better
placebo (24-month survival 46% vs. 71%; 36-month sur- nighttime rest without causing oversedation. Therefore, day-
vival 30% vs. 59%). time doses are significantly lower than the nighttime dose (e.g.,
Regarding concerns for QTc prolongation, data from the a 20 mg/d haloperidol dose will be distributed as 5 mg IV q
FDA Division of CardioRenal Drug Products Consultation 0600, 1000, and 1600; plus 10 mg at 2200 hrs).
suggested that all antipsychotic agents lengthen the QTc to When treating hypoactive delirium, we recommend doses
varying degrees. They based their conclusions on a study of in the very low range (i.e., haloperidol and risperidone in the
antipsychotic agents in psychotic subjects with normal base- 0.25 to 1 mg/24 hr). Data available suggest that the dopamine
line ECGs who were titrated to the “highest tolerated dose” of excess observed in delirious subjects occurs in all cases, even
one of six antipsychotic agents (see Table 41.14) (FDA, 2000; hypoactive forms. The same data demonstrate that antipsy-
Huffman & Stern, 2003). It may be reasonable to assume the chotic use helps prevent and treat all forms of deliria, includ-
doses used may have been substantially lower than the doses ing the hypoactive type. In these cases, it is usually given as
used for the treatment of hyperactive delirium. We also need a single nighttime dose, just before sundown. Given their
to consider the fact that when these agents are used for the hypoactive state, very sedating agents (e.g., quetiapine, olan-
management of delirium, patients often received parenteral zapine) are usually not recommended. In the experience of
formulations (when available) due to noncompliance, and the author, aripiprazole may prove to be a particularly good
they often experience a number of comorbid medical issues choice for hypoactive cases (usually starting at doses as low
(Huffman  & Stern, 2003). When antipsychotic agents are as 1 mg at bedtime), probably due to its partial dopamine
recommended, it is wise to review the patient’s medication antagonist-agonist properties that may have positive effects
list and identify any other agents with the ability to prolong on attention, concentration, and sleep–wake cycle reversal in
QTc. If possible, avoid other medications known to increase delirium, and its minimal muscarinic and histaminic antag-
QTc and/or inhibitors of CPY3A4. Before and during the use onist activity (thus minimizing adverse cognitive effects).
of continuous antipsychotic management, you should obtain Recent reports seem to support this clinical experience, con-
a 12-lead ECG, and measure QTc and electrolytes. All elec- firming the usefulness of aripiprazole, particularly in hypoac-
trolyte abnormalities should be corrected (especially K+ and tive delirium (Alao & Moskowitz, 2006; Straker, Shapiro, &
Mg+) and the QTc pattern monitored. Recommendations Muskin, 2006; Boettger & Breitbart, 2011). Aripiprazole had
enjoyed the reputation of being the only SGA not to be associ-
ated with QTc-prolongation or TdP, confirmed by both clini-
Table 41.14 EFFECTS OF OR ALLY ADMINISTER ED cal trials data and clinical experience (Gulisano, Cali, et al.,
ANTIPSYCHOTICS ON THE QTC INTERVAL a 2011; Germano, Italiano, et  al., 2014; Li, Luo, et  al., 2014;
MEAN INCR EASE % OF SUBJECTS W ITH Marder et al., 2003). Yet there are at least two case reports of
DRUG IN QTC (MS) > 60 MS INCR EASE IN QTC aripiprazole-induced cardiac arrest due to TdP, when it was
used as a single agent (Nelson & Leung, 2013). There is at least
Thioridazine 35.8 29 one case report of aripiprazole-associated QTc prolongation
Ziprasidone 20.6 21 in a geriatric patient (Hategan  & Bourgeois, 2014). Before
this, all other cases of QTc prolongation had either been asso-
Quetiapine 14.5 11 ciated with concomitant use of another antipsychotic agent or
Risperidone 10.0 4 occurred after an overdose (LoVecchio, Watts,  & Winchell,
2005; Nelson & Leung, 2013).
Olanzapine 6.4 4
Haloperidol 4.7 4
Norepinephrine Excess
Data from Huffman JC, Stern TA. (2003).QTc Prolongation and the use of
antipsychotics: A case discussion. Primary Care Companion to the Journal of α 2-Adrenergic Receptors Agonists versus
Clinical Psychiatry, 5(6), 278–81. Conventional Sedative Agents: Effect on Delirium
a
Data adapted from the U.S. Food and Drug Administration’s Center for Drug Prevention
Evaluation and Research, Psychopharmacological Drugs Advisory Committee There have been several randomized clinical trials looking at
(FDA, 2000), adapted by Huffman & Stern, 2003. anesthetic practice and delirium prevention (see Table 41.15).

880 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Table 41.15 Α 2-ADR ENERGIC R ECEPTORS AGONISTS AND ALTER NATIVE SEDATIVE AGENTS VERSUS
CONVENTIONAL ANESTHETICS: EFFECT ON DELIR IUM PR EVENTION AND TR EATMENT

DELIR IUM DELIR IUM INCIDENCE %


STUDY(N = 11) POPULATION INTERVENTION DEFINITION CONTROL/INTERVENTION P-VALUE

Berggren et al., Femoral neck Epidural vs. halothane DSM-III 38% 50% P < 0.05
1987; RCT, n = 57 Fx repair anesthesia (11/29) (14/28)

Williams-Russo B knee Continuous epidural DSM-III 44% 38% p = 0.69


et al., 1992; replacement Sx bupivacaine + fentanyl (11/25) (10/26)
RCT, n = 60 vs. continuous IV
fentanyl

Aizawa GI surgery Usual care vs. BZD DSM-IV 35% 5% p = 0.023


et al., 2002; administration to (7/20) (1/20)
OL, n = 42 promote sleep p-Sx

Maldonado et al., Cardiac Post-op anesthesia DSM-IV Midazolam Propofol Dexmedeto- p <0.001
2003, 2009; valve Sx w/ MID vs. PROP DRS-R-98 50% 50% midine
RCT, n = 118 vs. DEX 3%
(15/30) (15/30)
(1/30)

Pandharipande Med/Surg ICU DEX vs. lorazepam CAM-ICU 3.0d 7.0d p = 0.01
et al., 2007; in mechanical (2 tertiary care centers);
DBRPCT, n = 106 ventilation days alive w/o delirium
or coma

Reade et al., 2009; Tx-agitated ICU IV haloperidol ICDSC 42.2 hrs 20 hrs p = 0.016
randomized, patients 0.5 to 2 mg/hour vs.
open-label, dexmedetomidine Above numbers represent time to
parallel groups 0.2 to 0.7 μg/kg/hr extubation; prolonged intubation
attributed to delirium and agitation
pilot trial; n = 20

Riker et al., 2009;
Med/Surg ICU Midazolam vs. DEX; CAM-ICU 76.6% 54% p<0.001
DBRPCT, n = 375 in mechanical trial conducted in 68 (93/122) (32/244)
ventilation centers in 5 countries

Hudetz Elective PBO vs. IV ketamine Intensive Care 31% 3% p = 0.01


et al., 2009; CABG or valve (0.5 mg/kg) bolus Delirium (9/29) (1/29)
DBRPCT, n = 58 replacement/ during the induction of Screening
repair w CPB anesthesia Checklist
(ICDSC)

Shehabi Elderly patients Morphine vs. DEX CAM-ICU 15% 8.6% p =.088
et al., 2009; after cardiac
DBRPCT, n = 306 surgery

Rubino Acute PBO vs. clonidine Delirium 1.8 ± 0.8 0.6 ± 0.7 p = 0.001
et al., 2010; type-A aortic IV on delirium Detection
DBRPCT, n = 30 dissection repair neurological outcome Score (DDS)→
and respiratory
function
Jakob, Ruokonen, Adult ICU PROP vs. DEX CAM-ICU 29% 18% p = 0.008
et al., 2012; patients (71/247) (45/251)
RDBCT; n = 498 receiving
mechanical
ventilation
Abbreviations: RCT=randomized control trial; RPCT=randomized placebo control trial; DBRPCT=double blind randomized, placebo controlled trial;
OL=open label; trial; DSM=Diagnostic and Statistical Manual of Mental Disorders; CAM= Confusion Assessment Method; DRS=Delirium Rating Scale;
MDAS=Memorial Delirium Assessment Scale; GI=gastrointestinal; PBO=placebo; Sx=surgery; CPB=cardio-pulmonary by-pass machine; ICU=intensive care
unit. APA=antipsychotic agent; BZDP=benzodiazepine; DEX=dexmedetomidine; OLA=olanzapine; HAL=haloperidol; RIS=risperidone; AIDS=autoimmune
deficiency syndrome; CPM=chlorpromazine; EPS= extra-pyramidal syndrome; LOR=lorazepam; PO= by mouth.
The earliest study examined epidural versus halothane anes- kg/minute), or MID drip (0.5–2 mg/hour), all titrated to achieve
thesia to assess postoperative delirium and found no difference a target sedation level (i.e., a Ramsay Sedation Score of 3 before
between the two approaches (Berggren et al., 1987). A second extubation and 2 after extubation). Upon arrival at the ICU, a
study compared the effects of continuous epidural bupiva- standardized protocol for postoperative care was implemented
caine plus fentanyl versus continuous IV fentanyl anesthesia for all patients. Study results showed there were no significant pre-
on postoperative delirium, and again found no differences operative or intraoperative differences between treatment groups
between these approaches (Williams-Russo et al., 1992). (e.g., age, sex, American Society of Anesthesiology classification,
The majority of patients in the ICU, particularly those who bypass time, clamp time, or lowest temperature achieved) (see
are mechanically ventilated, receive some form of sedation to Table 41.16). The only real difference in management between
reduce their anxiety, encourage sleep, and increase tolerance of groups was the type of postoperative sedation. Final results dem-
the critical care environment, including multiple IV lines, pain onstrated an incidence of delirium of 3% for patients on DEX,
management, endotracheal tubes, and ventilators. Sedative compared to 50% for PRO and 50% for MID (p <.01). The abso-
and analgesic drugs are among the most commonly prescribed lute risk reduction in the incidence of delirium associated with
medications in the ICU (Farina, Levati,  & Tognoni, 1981). using DEX was 47% (95% CI 28%–66%) corresponding to a
As discussed, sedative agents (mostly GABA-ergic) and opi- number-to-treat of 2.1 patients (95% CI 1.5–3.6). Similarly, the
oids may contribute to the development of delirium by one of number of delirious days was also significantly lower in the DEX
six mechanisms: interfering with physiological sleep patterns; group compared to PRO and MID (1% vs. 16% vs. 29%, respec-
interfering with central cholinergic function; increasing com- tively; p <.001).
pensatory upregulation of NMDA and kainite receptors and Since then, two DBRPCT have confirmed the original
Ca2+ channels; disrupting the circadian rhythm of melato- findings and demonstrated the delirium-sparing effects of
nin release; disrupting thalamic gating function; and leading DEX. The first of these compared DEX with lorazepam (LOR)
to CNS-depressant dependence and withdrawal. Therefore, in the ICU on mechanical ventilation. The study found that
sedating alternatives have been thought promising in decreas- DEX-treated patients had more delirium-free days (7 days vs.
ing medication/sedation-induced delirium. 3  days, p =.01), had a lower prevalence of coma (p <0.001),
The first study involving ketamine in delirium preven- and spent more time within sedation goals (p =.04) than
tion was a DBRPCT involving children undergoing dental LOR-treated subjects (Pandharipande et al., 2007). Similarly,
repair in sevoflurane-induced anesthesia (Abu-Shahwan  & a second, much larger trial (i.e., 68 ICUs in five countries)
Chowdary, 2007). The study demonstrated a substantially compared the incidence of delirium among ICU-ventilated
lower incidence of emergence agitation in the ketamine group subjects sedated with either DEX or MID and found a lower
compared with placebo (16.6% vs. 34.2%), but no difference delirium incidence (54% vs. 76.6%, p <.001), shorter intuba-
in time to meet recovery room discharge criteria between the tion time (median time to extubation was 3.7 days DEX vs.
two groups. A second study compared the effects of IV ket- 5.6 days MID; p =.01), and less tachycardia and hypertension
amine versus placebo on postoperative delirium in patients in the DEX-treated group (Riker et al., 2009).
undergoing acute type-A aortic dissection repair, which A study assessing the effects of different classes of sedative
found a significant reduction in the incidence of delirium in agents compared variable doses of morphine or dexmedeto-
patients in the ketamine group (3% vs. 31%, p = 0.01) (Hudetz midine, with open-label propofol titrated to effect (Shehabi
et al., 2009). Postoperative C-reactive protein concentration et al., 2009). This study found that the incidence of delirium
was also lower (p <0.05) in the ketamine-treated patients was comparable between morphine 22 (15.0%) and DEX 13
compared with the placebo-treated patients. (8.6%) (95% CI:  0.256–1.099, p =.088). Yet DEX-managed
In an attempt to demonstrate that different sedative strate- patients spent 3 fewer days (2 vs. 5)  delirious (95%
gies may achieve delirium reduction, the author and his team CI:  1.09–6.67, p =.0317), were more likely to be extubated
(Maldonado, 2003) were the first to report on the use of the novel earlier (95% CI: 1.01–1.60, p =.040), experienced less systolic
sedative agent dexmedetomidine (DEX) as an alternative to the hypotension (23% vs. 38.1%, p =.006), and required less nor-
use of benzodiazepines and related agents (e.g., midazolam [MID], epinephrine (p < 0.001), than morphine-treated patients.
propofol [PRO]) during the postoperative state (Maldonado, Finally, a subsequent study assessed the role of clonidine IV
2003; Maldonado et al., 2003, 2004). Post-cardiotomy patients (0.5 microg/kg bolus, followed by continuous infusion at 1–2
were selected, given the high incidence of delirium in such microg/kg/hr) versus placebo on delirium in post-cardiotomy
patients (around 57%) nationwide (Ebert, Walzer, Huth,  & patients and found no significant difference in the incidence
Herrmann, 2001). We studied patients (n = 118)  undergoing of delirium between groups (33% vs. 40%, p = 0.705) (Rubino
cardiac surgery (i.e., repair or replacement) with cardiopulmo- et al., 2010). Yet, delirium scores were significantly lower in
nary bypass (Maldonado et al., 2009). Intraoperative anesthesia patients treated with clonidine compared to placebo (P =
for the surgical procedures was standardized for all subjects in all 0.001), suggesting that the beneficial effects in delirium pre-
three studied groups. All procedures were performed via median vention previously seen with DEX may extend to all alpha-2
sternotomy in conjunction with CPB and induction of moder- agonists (Rubino et al., 2010).
ate hypothermia. After successful weaning from CPB, patients Despite evidence demonstrating that prophylactic use
were started on one of three randomly assigned, postoperative of DEX in reducing the incidence of delirium and plenty of
sedation regimens: DEX (loading dose: 0.4 μg/kg, followed by a clinical evidence that it is useful in the management of agi-
maintenance drip of 0.2–0.7 μg/kg/hour), PRO drip (25–50 μg/ tation related to hyperactive delirium, there is no study to

882 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Table 41.16 SELECTED POSTOPER ATIVE OUTCOME VAR IABLES FOR CAR DIAC PATIENTS WITH
CAR DIOPULMONARY BYPASS BY INTERVENTION GROUP

DEX MEDE-
TOMIDINE PROPOFOL MIDAZOLA M OVER ALL DEX VS. DEX VS.

(N = 30) (N = 30) (N = 30) P-VALUE PROPOFOL MIDAZOLA M

Delirium

Incidence of delirium (per 1/30 (3%) 15/30 (50%) 15/30 (50%) <0.001 <0.001 <0.001
protocol)

Incidence of delirium 4/40 (10%) 16/36 (44%) 17/40 (44%) <0.001 0.001 0.002

Number of days delirious 2/216 (1%) 45/276 (16%) 75/259 (29%) <0.001 <0.001 <0.001

Average length of delirium days) 2.0 ± 0 3.0 ± 3.1 5.4 ± 6.6 0.82 0.93 0.63

Time Variables

ICU length of stay (days) 1.9 ±.9 3.0 ± 2.0 3.0 ± 3.0 0.11 0,14 0.14

Hospital length of stay (days) 7.1 ± 1.9 8.2 ± 3.8 8.9 ± 4.7 0.39 0.42 0.12

Intubation time (hours) 11.9 ± 4.5 11.1 ± 4.6 12.7 ± 8.5 0.64 0.91 0.34

PRN Medications

Fentanyl (mcg) 320 ± 355 364 ± 320 1088 ± 832 <0.001 0.93 <0.001

Total Morphine equivalents 50.3 ± 38 51.6 ± 36 122.5 ± 84 <0.001 0.99 <0.001


(mg) P‡P

Antiemetic use* 15/30 (50%) 17/30 (57%) 19/30 (63%) 0.58

PRN Medications for the


Management of Delirium**

Lorazepam 1/30 (3%) 7/30 (23%) 6/30 (20%) 0.07 0.06 0.11
Haloperidol 0/30 3/30 (10%) 2/30 (7%) 0.23 0.07 0.15
† Percent of patients who developed delirium.

Sum of average morphine equivalents (fentanyl, oxycodone, and hydrocodone) received in postoperative days 1–3.
* Number of patients who received dolasetron mesylate and/or promethazine HCl in postoperative day 1.
** Average amount over three days. None of these medications were given until a diagnosis of delirium was established.
Source: Maldonado, 2008a.

date confirming its potential in the treatment of delirium. that DEX has an intrinsic delirium-sparing effects, which
A randomized, open-label trial of agitated delirium assigned may be explained by DEX’s pharmacological characteristics.
subjects to receive an infusion of either haloperidol 0.5–2.0 The second theory suggests that the reason patients had sig-
mg/hour or dexmedetomidine 0.2–0.7  μg/kg/hr (Reade nificantly less delirium in the DEX group was not because of
et al., 2009). Results found that DEX significantly shortened its use per se, but because those patients were not exposed to
median time to extubation (from 42.5 to 19.9 hours; P = other sedative agents with much greater delirium potential
0.016), significantly decreased ICU length of stay (from 6.5 to (e.g., GABA-ergic agents, such as propofol and midazolam,
1.5 days; P = 0.004), and, in patients requiring ongoing seda- and lower doses of opioid agents). Third, data suggest that
tion, DEX cut in half the time propofol was needed (79.5% the neuroprotective effect of DEX is mediated by activation
vs. 41.2%; P = 0.05). Finally, a recent meta-analysis including of the alpha2A adrenergic receptor subtype (Ma et al., 2004).
all postoperative prevention strategies found that DEX seda- Fourth, it is worth noticing that DEX has been shown to have
tion was associated with less delirium compared to sedation significant neuroprotective effect, in vivo and in vitro, against
produced by other drugs (2 RCTs with 415 patients, pooled glutamate-induced damage via an increased astrocyte expres-
RR = 0.39; 95% CI = 0.16–0.95) (Zhang et al., 2013). sion of BDNF through an extracellular signal-regulated
The detailed proposed mechanisms for delirium reduction kinase-dependent pathway (Engelhard et al., 2003; Ma et
have been described elsewhere (Maldonado, 2008a). They can al., 2004; Sato, Kimura, Nishikawa, Tobe, & Masaki, 2010;
be summarized in six postulated theories: The first suggests Degos et al., 2013). Fifth, it is also hypothesized that DEX

41. D E L I R I U M • 883
increases the expression of active (autophosphorylated) focal physiological synaptic activity, thus preventing excessive cal-
adhesion kinase, a non-receptor tyrosine kinase playing a piv- cium influx into neurons—believed to be the key early step in
otal role in cellular plasticity and survival (Dahmani, Rouelle, GLU-induced exocytosis); reducing the release of proinflam-
Gressens, & Mantz, 2005). The sixth theory involves DEX’s matory factors from activated microglia; inducing expression
unique mechanism for inducing sleep. Data suggest that DEX of neurotrophic factors, such as glial cell line–derived neuro-
induces a qualitatively similar pattern of c-Fos expression as trophic factor in astroglia; and limiting oxidative injury and den-
seen during normal NREM sleep by acting on endogenous dritic degeneration induced by anticholinesterase neurotoxicity
sleep pathways (i.e., a decrease in the locus ceruleus and tuber- (Giacino & Whyte, 2003; Zaja-Milatovic, Gupta, Aschner, &
omammillary nucleus and an increase in the ventrolateral pre- Milatovic, 2009; Ossola et al., 2011; Kutzing, Luo, & Firestein,
optic nucleus) (Nelson et al., 2003). 2012). Thus, it makes sense to consider their use in various syn-
It is certainly possible that a combination of any or all of dromes associated with excess glutamate and subsequent cogni-
these six mechanisms work together, leading to the results tive decline (e.g., traumatic brain injury, stroke, delirium). In
obtained in the above-mentioned studies. It is this author’s fact, studies have demonstrated that memantine may be effec-
experience that all centrally acting α2-adrenergic recep- tive in reducing the damage induced by acute ischemia/reperfu-
tors agonists are to some extent useful in the management sion (Yigit et al., 2011), while amantadine has been shown to
of hyperactive delirium. In fact, our team uses clonidine to enhance cognitive recovery and minimize delirium after severe
help transition patients off IV-DEX with good success (see TBI in humans (Giacino et al., 2012).
Table 41.17). We also use clonidine, orally and transdermally, A DBPCRT compared the use of gabapentin as an add-on
for prevention of alcohol withdrawal, with excellent results. agent in the treatment of postoperative pain (i.e., spine sur-
Another α2-adrenergic receptor agonist worth considering gery under general anesthesia) to reduce the occurrence of
is guanfacine. This agent is even more α2/ α1 selective and is postoperative delirium (Leung et al., 2006). Subjects received
available in oral dose form. This makes it an excellent alterna- either placebo or gabapentin 900 mg administered orally one
tive to DEX (which requires an IV drip and a monitored bed) to two hours before surgery, and continued once a day for the
and is less hypotensive than clonidine. Our team has been first 3 postoperative days. Results demonstrated that gabapen-
using this agent, as we have clonidine, both as adjunct treat- tin was superior to placebo in reducing delirium occurrence
ment of hyperactive delirium and as a primary agent in the (0% vs. 42%, p = 0.045). Even though the study authors attrib-
prevention and treatment of alcohol withdrawal. uted this to gabapentin’s opioid-sparing effect, they failed to
consider gabapentin’s true modes of action (i.e., modulation
of voltage-sensitive Ca2+ channels, NMDA receptor antag-
Glutamate Excess
onism, activation of spinal alpha-2 receptors, attenuation of
Glutamate Antagonists and Ca+ Channel Na+ dependent action potentials) as potential mediators of
Modulators for Delirium Management its deliriolytic effect.
A number of agents with antiglutamatergic and Ca+ chan- Even though many of us routinely use VPA (either oral
nel blocking qualities are worth considering here:  lamotrig- or IV) in the management of agitated delirious patients who
ine, amantadine, memantine, gabapentin, and valproic acid either are not responsive or cannot tolerate conventional
(VPA) (see Table 41.18). treatment, there is very little literature investigating the effec-
Both amantadine and memantine have been recognized as tiveness of this practice. In fact, there are no RCT or even OL
having neuroprotective effects, probably mediated by their pro- studies on the matter. The only published report on the effec-
tection from glutamate-induced exocytosis (by blocking exces- tiveness of VPA for delirium treatment comes from a report of
sive N-methyl-D-aspartic acid receptors without disrupting six cases (Bourgeois, Koike, Simmons, Telles, & Egglestone,

Table 41.17 CENTR ALLY ACTING Α 2-ADR ENERGIC R ECEPTOR AGONISTS

PRODUCT PROTEIN
DRUG Α 2/Α 1SELECTIVITY DT ½ ET ½ AVAILABILITY BIOAVAILABILITY BINDING

Guanfacine 2,640 2.5 hr 17 hr PO ~100% 70%

Dexmedetomidine 1,600 6m 2 hr IV 70–80% 94%

Medetomidine 1,200

Clonidine 220 11 m 13 hr PO 100% PO 40%


TDS 60% TDS

Methyldopa 12 m 105 m PO/IV 50% <20%


Guanabenz 60 m 6 hr PO 75% 90%
Abbreviations: dT1/2= time to onset; eT1/2= elimination half-life; PO=by mouth; IV=intravenous; TDS= transdermal delivery (patch).

884 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Table 41.18 GLUTAMATE ANTAGONISTS AND CA+ CHANNEL MODULATORS FOR DELIR IUM TR EATMENT

PRODUCT BIO- PROTEIN


DRUG T½ AVAILABILITY AVAILABILITY METABOLISM BINDING MECHANISM ACTION

Lamotrigine 25 hr PO ~100% Hepatic 55% • Stabilizes neuronal membranes


• Inhibits voltage-sensitive Na+
channels and/or Ca+ channels→ ↓
cortical GLU release
• Calcium channel blockers
• Excitatory amino acid antagonists

Amantadine 17 ± 4 hr PO 86–90% None 67% • NMDA receptor antagonist


Renal excretion • ↑ synthesis and release of
dopamine

Memantine 60–80 hr PO 100% Mostly 45% • Noncompetitive NMDA receptor


unchanged renal antagonist
excretion • Blocks the effects of excessive
levels of GLU
• Some Ca+ channel blockade
• 5HT3 antagonist

Gabapentin 5–7 hr PO 60% None <3% • Voltage-gated Ca+ channel


Renal excretion blockade → ↓ cortical GLU release
• NMDA antagonism
• Activation of spinal
alpha2-adrenergic receptors
• Attenuation of Na+ dependent
action potential
VPA 9–16 hr PO/IV 90% Hepatic 90% • GABA transaminase inhibitor
conjugation → ↑ GABA
• Inhibits voltage-sensitive Na+
channels→ ↓ cortical GLU release
• ↓ release of the
epileptogenic amino acid
gamma-hydroxybutyric acid
(GHB)
Abbreviations: T1/2=elimination half-life.

2005). The authors reported that “in all six cases, the use of addition of VPA frequently allowed prompt extubation, since
VPA combined with conventional antidelirium medications agitation and related desaturation were better controlled. No
resulted in improved control of behavioral symptoms without significant side effects were identified in any of the subjects,
significant side effects.” In all cases, standard psychotropic with the exception of nonclinically significant decreases in the
medication regimens commonly used for the management of number of platelets in few subjects. Of note, as in the case of
delirium and/or nonspecific agitation in the general hospital SGAs, there have been case reports on VPA-induced delirium.
setting (i.e., antipsychotics and/or benzodiazepines) had ini- The availability of IV (Depacon) and elixir (Depakene, ideal
tially been tried with limited success, and in all cases the addi- for administration via feeding tube, if needed) forms allow
tion of VPA to the existing regimens resulted in improved administration in non-cooperative subjects, or those intu-
clinical response. Daily doses ranged between 0.5 to 2.5 grams bated and unable to take PO medication. The IV formulation
in divided doses, without significant side effects and with dis- even allows for loading to quickly establish therapeutic serum
cernible therapeutic benefits. levels, then continuing for maintenance or converting to PO
Similarly, the Stanford’s Psychosomatic Medicine when medically indicated. Others have described the useful-
Research Lab presented a larger case series (n = 16) equally ness of VPA in the management of destructive and aggressive
demonstrating the usefulness and safety of VPA in the man- behavior associated with brain injury (Wroblewski, Joseph,
agement of agitated/hyperactive delirium, as an add-on to Kupfer,  & Kalliel, 1997)  and dementia (Haas, Vincent,
ineffective response to conventional therapy (Sher, Lolak, Holt, & Lippmann, 1997; Narayan & Nelson, 1997; Tariot,
Miller, & Maldonado, 2013). In this series, the dose ranged 1999; Porsteinsson, Tariot, et al., 2003; Sival, Duivenvoorden,
between 0.5 to 3.0 grams per day in divided doses. All subjects et al., 2004). As in any patient receiving VPA, close monitor-
in this case series improved in their clinical status with even- ing of liver function tests, bilirubin, platelet count, and amy-
tual resolution of delirium. In the case of intubated patients, lase are advisable, especially in the critically ill patient.

41. D E L I R I U M • 885
Cholinergic Deficit (40% vs. 62%; p <0.001). Also of significance, when sub-
jects developed delirium, the mean duration of delirium was
Acetylcholinesterase Inhibitors for Delirium shorter (mean duration 4  ± 1.71  days vs. 7.86  ± 2.73  days;
Prevention p <0.01), and the use of benzodiazepine and neuroleptic
Despite the logical premise behind the prophylactic use of agents for management of agitation was significantly less
acetylcholinesterase inhibitor agents, studies have not con- in the rivastigmine group (p <0.05). Furthermore, analysis
sistently demonstrated positive results (see Table 41.19). The of behavioral data (as measured by Behavioral Pathology in
first publication suggesting the promise of acetylcholinester- Alzheimer’s Disease (BEHAVE-AD) scale), found that total
ase inhibitors in delirium prevention comes from a study on scores in the rivastigmine group were significantly improved
the use of rivastigmine for delirium prevention among hos- over baseline and at the end of the study (all p <0.001). Further
pitalized elderly patients who had been chronically treated sub-analysis of the BEHAVE-AD individual items indicated
with rivastigmine for dementia (Dautzenberg, Mulder, Olde that rivastigmine provided benefits on all items of the scale,
Rikkert, Wouters,  & Loonen, 2004). A  retrospective chart except for delusions, throughout the study.
review showed that those on rivastigmine had a significantly Unfortunately, the only DBRCT of rivastigmine failed
lower incidence of delirium compared to subjects not treated to demonstrate efficacy in the prevention of postoperative
(45.5% vs. 88.9%; p <0.05). Similarly, a prospective study of delirium. In a study in cardiac surgery patients comparing
elderly subjects suffering from vascular dementia randomly rivastigmine (1.5 mg/d for 3 days preoperatively) against pla-
placed patients on either rivastigmine (6 mg/d) or aspirin (100 cebo found no benefits of the active drug use (Gamberini
mg/d) and followed them at regular intervals for 24 months et al., 2009).
(Moretti, Torre, Antonello, Cattaruzza,  & Cazzato, 2004). Similarly, results have been negative with the use of
The results suggested that those on rivastigmine had a sub- donepezil. The first study was a DBRPCT involving elderly
stantially lower incidence of delirium compared to aspirin patients undergoing elective total joint replacement surgery

Table 41.19 ACETYLCHOLINESTER ASE INHIBITORS IN DELIR IUM PR EVENTION

DELIR IUM INCIDENCE %


DELIR IUM
STUDY(N = 7) POPULATION INTERVENTION DEFINITION CONTROL INTERVENTION P-VALUE

Dautzenberg (P) ≥65y/o Patients who Retrospective 88.9 45.5 P < 0.05
et al., 2004 hospitalized used rivastigmine chart review of (26/29) (4/11)
OL, Retrospective demented chronically with a geriatric service
review, n = 51 patients randomly selected consultations
subgroup of all patients
not treated

Moretti et al., 2004 (P) ≥65y/o–o/p, Cardioaspirin vs. CAM 62 40 P < 0.001


RCT, n = 230 w/ vasc dementia rivastigmine PO qD BEHAVE-AD (71/115) (46/115)
(24-month f/u)

Liptzin et al., 2005 (P) Elderly PBO vs. donepezil DSM-IV 17.1 20.5 p = 0.69
DBRPCT n = 80 elective total joint (14 days pre + 14 days (7/41) (8/39)
replacement post Sx)

Sampson et al., 2007 (P) Elderly PBO vs. donepezil 5 DSI 35.7 9.5 p = 0.08
DBRPCT, n = 33 elective hip mg immediately p-Sx (5/14) (2/19)
replacement + 3 days

Oldenbeuving (P) Delirium p Rivastigmine 3→ DRS ≥ 12 In 16/17 (94%) delirium severity


et al., 2008 CVA 12 mg/d; no PBO improved; mean decrease 14.8→8.5; mean
N = 26 duration: 6.7 days; no side effects

Gamberini (P) Cardiac Sx PBO vs. PO CAM 30 32 p = 0.8


et al., 2009 under CPB rivastigmine 1.51 (17/57) (18/56)
DBRPCT, n = 120 pre-op, until POD #6

van Eijk et al., (Tx) >18 y/o in 2 arms, both receiving CAM-ICU 3d 5d p = 0.06
2010DBRPCT, ICU haloperidol, trial PBO
n = 109 vs. rivastigmine Above refers to delirium duration
in days; trial halted due to mortality
in rivastigmine (n = 12, 22%) was >
than in placebo group (n = 4, 8%;
p = 0.07)
1
Rivastigmine-treated patients who experienced delirium had a shorter duration, lower use of benzodiazepine and neuroleptic for management of agitation, and
improvement in all behavioral aspects measured by the BEHAVE-AD.

886 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
(Liptzin, Laki, Garb, Fingeroth, & Krushell, 2005). Subjects Box 41.8 CASE R EPORTS SUGGESTING A POSITIVE
in the active treatment group received donepezil or placebo EFFECT OF ACETYLCHOLINESTER ASE INHIBI-
for 14 days before surgery and 14 days afterward, and no sig- TORS IN THE TR EATMENT OF DELIRIUM
nificant differences between groups were found (p = 0.69).
A second RCT also failed to demonstrate efficacy of donepezil • Burt, 2000
(5 mg immediately postoperatively and every 24 hrs thereafter
for the first 3 postoperative days) over placebo in preventing • Bruera, Strasser, et al., 2003
postoperative delirium after elective total hip replacement • Dautzenberg, et al., 2003, 2004
surgery in older people without preexisting dementia (9.5%
donepezil vs. 35.7% PBO; p = 0.08) (Sampson et al., 2007). • Fisher et al., 2001
Even though this was a negative study, the trend was in the • Gleason, 2003
right direction (lower incidence of delirium; shorter length of
stay [9.9 vs. 12.1 days; p = 0.09] and shorter delirium duration • Hasse & Rundshagen, 2007
[1.5 vs. 1.8 days; p = 0.83]), and the study had Type-II errors • Hori, Tominaga, et al., 2003
(i.e., it was grossly underpowered).
The results of the acetylcholinesterase trials seem to sug- • Kaufer, Catt, et al., 1998
gest that these agents are either (a)  just not good for delir- • Kobayashi et al., 2004
ium prevention; or (b) they need to be used for much longer
periods of time (as in the case of Dautzenberg and Moretti’s • Logan & Stewart, 2007
study); or (c)  they need to be used at doses much higher • Moretti et al., 2004
than we currently use them in order to achieve acute clinical
efficacy. • Palmer, 2004
A note of warning: more recently a DBRPCT compar-
• Rabinowitz, 2002
ing rivastigmine to placebo was stopped after members of
the data safety and monitoring board unblinded the results • Weizberg et al., 2006
and performed an interim analysis, halting the study due to
• Wengel, Roccaforte, & Burke 1998
concerns about a higher mortality rate in the rivastigmine
group (p = 0.07) (van Eijk et al., 2010). The authors failed • Wengel, Burke, & Roccaforte, 1999
to explain the mechanism by which rivastigmine use may
have led to the higher mortality reported. It is difficult to
interpret their report, as less than 25% of the sample had Physostigmine and Delirium
been studied, and the causes of death for those on the study Acetylcholine deficiency has been postulated as one of the
drug were heterogeneous. It is clear that the study relied on potential causes of delirium; whether this is caused by nor-
a very fast rivastigmine titration (essentially doubling the mal physiological causes (e.g., aging) or due to exogenous fac-
dose every 3 days until reaching the goal of 6 mg BID by day tors (e.g., use of anticholinergic substances). Thus, among the
10). This could have contributed to untoward side effects, tools to treat delirium, particularly when it is presumed to be
which could have potentially worsened underlying cardio- due to the ingestion of substances with anticholinergic poten-
vascular or metabolic processes. Unlike in other delirium tial, we should consider the potential use of physostigmine.
studies, the population in this study was not homoge- Physostigmine, a short-acting acetylcholinesterase inhibi-
neous, as it involved patients with all types of disease in the tor, increases synaptic acetylcholine concentrations and can
ICU. As the authors reported, “we cannot exclude that the overcome the postsynaptic muscarinic receptor-blockade pro-
recorded difference in mortality is due to chance” (van Eijk duced by anticholinergic agents. As a tertiary amine, it can
et al., 2010, p.1835). pass freely into the CNS and reverse both central and periph-
eral anticholinergic effects.
Many reports have demonstrated its utility and safety in
Acetylcholinesterase Inhibitors for Delirium Treatment
cases when delirium has been caused by medication overdose
Addressing the theory that proposes that delirium is caused (whether accidental or intentional) (Stern, 1983); Lipowski,
by a central cholinergic deficiency state, some researchers 1992; Beaver & Gavin, 1998; Richardson, Williams, & Carstairs,
and clinicians have experimented with the use of acetylcho- 2004; Eyer et al., 2011; Hail, Obafemi, & Kleinschmidt, 2013).
linesterase inhibitor agents. Most of the published data con- Physostigmine has been successfully used to treat emergence
sist of small series of case reports associated with the use of delirium in both adults (Brown, Heller,  & Barkin, 2004;
rivastigmine in the treatment of delirium in older persons Haase  & Rundshagen, 2007)  and pediatric patients (Funk,
(Dautzenberg, Mulder, et al., 2004; van den Bliek & Maas, Hollnberger, & Geroldinger, 2008). It has been reported that
2004)(van den Bliek and Maas 2004). There have been at physostigmine attenuates several withdrawal states, especially
least 19 papers, mostly case reports, suggesting that acetyl- alcohol delirium and opiate and nitrous oxide withdrawal
cholinesterase inhibitor agents (e.g., donepezil, galantamine, syndromes; and it may offer a protective mechanism against
physostigmine, rivastigmine) may be effective in the treat- hypoxic damage of the brain (Powers, Decoskey, & Kahrilas,
ment of delirium (see Box 41.8). 1981; Rupreht, Schneck, & Dworacek, 1989).

41. D E L I R I U M • 887
Despite many psychiatrists’ concerns about physostig- Various sedative hypnotic agents are commonly used in
mine administration, its use as a diagnostic test appears to be the medical setting. Nevertheless, benzodiazepines and the
rather safe. In a series of 39 adults treated with varying doses benzodiazepine-receptor agonists (e.g., zolpidem) may cause or
of physostigmine (range 0.5–2.0 mg), 56% of all cases experi- exacerbate delirium. Other sedating agents, such as antidepres-
enced full reversal of delirium. Among the cases known to be sants (e.g., trazodone, mirtazapine) and antipsychotics (e.g., olan-
antimuscarinic in etiology (e.g., overdose with a known sub- zapine) have been used off-label to promote sleep, but these agents
stance) 100% of patients experience full reversal. In this sam- have not been tested in the critically ill population. Clinicians
ple, no patient experience dysrhythmias or signs of cholinergic must also consider factors such as drug–drug interaction and
excess, or required the administration of atropine. One (2.6%) medication half-lives when prescribing. For example, mirtazapine
in 39 patients experienced a brief convulsion, with no known and trazodone may indeed promote night sleep, but their effects
adverse sequelae (Schneir et al., 2003). Similarly, others have may last well into the next day, interfering with cognition, atten-
reported that among 52 consecutive patients with suspected tion, and concentration. Sedative agents with high anticholiner-
anticholinergic delirium, physostigmine controlled agitation gic load, such as antihistaminic agents (e.g., diphenhydramine,
and reversed delirium in 96% and 87% of cases, respectively hydroxyzine) or tricyclic antidepressants (e.g., amitriptyline)
(Burns, Linden Graudins, Brown, & Fletcher, 2000). Again, should be avoided, as they will aggravate delirium even if they are
no significant side effects were reported. immediately effective in promoting sleep. Similarly, benzodiaz-
Given its safety profile and effectiveness, physostigmine epines should also be avoided if at all possible.
should be considered when a delirious patient’s examination An alternative to these is the use of non-benzodiazepine
exhibits signs of a central anticholinergic state (e.g., confu- agents, such as melatonin or melatonin agonists (i.e., ramelt-
sion, sinus tachycardia, markedly dilated and fixed pupils, dry eon). Melatonin has been shown to play an important role
mouth, hypoactive bowels sounds, dry and flushed skin; see Box in the regulation of circadian rhythm and maintenance of a
41.9) and/or when it is known that the patient’s altered mental physiological, well-regulated sleep-wake pattern (Brzezinski,
status is due to the use of known anticholinergic substances 1997). Multiple studies have demonstrated a number of
(e.g., diphenhydramine). An initial physostigmine dose of 1–2 abnormalities linking melatonin and delirium, includ-
mg (0.5 mg in children) given intravenously over 3 to 5 minutes ing (a)  irregular patterns of melatonin circadian rhythm
is the recommended dose. If the response is incomplete, addi- (Miyazaki et  al., 2003; Olofsson, Alling, Lundberg,  &
tional doses of 0.5–1.0 mg every 5 minutes may be given until Malmros, 2004); and (b)  decreased melatonin secretion
delirium resolves or there are signs of cholinergic excess (e.g., (Shigeta et al., 2001; Guo, Kuzumi, Charman, & Vuylsteke,
diaphoresis, salivation, vomiting, diarrhea). A  prolonged PR 2002)  and abnormalities in the 24-hour urinary excretion
interval (>200 ms) or QRS complex (>100 ms and not related of 6-sulphatoxymelatonin (6-SMT); the chief metabolite of
to bundle branch block) interval on ECG are considered the melatonin (Balan, Leibovitz, et al., 2003). (See Table 41.20
only absolute contraindications for physostigmine use. for publications on melatonin abnormalities in delirium.) It
is also theorized that melatonin may have protective effects,
at least in cases of global cerebral ischemia/reperfusion
Melatonin and Sleep Deficit
injury, which may have implications for delirium onset (Sun,
Sleep deprivation is one of the major theories on the develop- Lin, et  al., 2002; Zhang, Guo, et  al., 2002; Cheung, 2003;
ment of delirium, whether caused by medication effect, envi- Cervantes, Morali, et  al., 2008). In fact, melatonin may
ronmental factors, or patient characteristics. Many sedative have a number of beneficial physiological effects which may
and hypnotic agents may worsen sleep and thus are not recom- prove benefical in the management of medically ill, delirious
mended. As described in the prevention section, various seda- patients (see Box 41.10).
tive hypnotic agents are commonly used in the medical setting,
but most of them have significant problems, including caus- Melatonin in Delirium Prevention
ing or worsening delirium, cognitive clouding, and respiratory Several case reports have described the successful use of melato-
depression. nin in preventing postoperative delirium (Hanania & Kitain,
2002). A small DBRPCT demonstrated that the use of mela-
Box 41.9 ANTICHOLINERGIC SIDE EFFECTS tonin (10 mg q HS) was associated with longer and better sleep
quality (p = 0.04) (Bourne et al., 2008). A recent DBRPCT
Cognitive impairment (n  = 222)  compared the effects of melatonin versus placebo
Delirium among elderly patients undergoing hip arthroplasty under
Dilated and fixed pupils, blurred vision spinal anesthesia and found that those receiving melatonin
Dry and flushed skin showed a statistically significant decrease in the percent-
Dry mouth age of postoperative delirium (9.43% vs. 2.65%, p = 0.003)
Fever (Sultan, 2010). Similarly, a recent DBRPCT (n = 145) among
Hallucinations tertiary-care medicine service elderly individuals demon-
Hypoactive bowels sounds, constipation, ileus strated that the administration of low-dose (i.e., 0.5 mg) mela-
Sinus tachycardia tonin was associated with a lower risk of delirium (12.0% vs.
Urinary retention 31.0%, p = 0.014), adjusted for dementia and other comorbidi-
ties (Al-Aama et al., 2011).

888 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
Table 41.20 DELIR IUM-ASSOCIATED MELATONIN ABNOR MALITIES

DELIR IUM
STUDY POPULATION INTERVENTION DEFINITION R ESULTS

Shigeta Men and women None; observational CAM All patients without delirium showed nearly identical
et al., 2001 undergoing preoperative and postoperative melatonin secretion for
N = 29 laparotomy for 24 hours.
digestive disease Conclusions: Abnormal melatonin secretion may be
involved in postoperative sleep disturbances, which
triggered delirium in elderly patients.

Guo et al., Males undergoing None; observational Only three patients regained circadian secretion of cortisol.
2002N = 12 elective CABG with Conclusion: melatonin and cortisol secretion were
CPB disrupted during cardiac surgery with CPB and in the
immediate postoperative period

Miyazaki s/p esophagectomy None; observational CE A significant correlation was seen between ICU psychosis
et al., 2003 for CA and an irregular melatonin circadian rhythm (P =.0001).
N = 41

Balan Patients who None; patients ICD-10 24-hr urinary excretion of 6-sulphatoxymelatonin
et al., 2003 developed were divided DRS (6-SMT)—the chief metabolite of melatonin.
N = 31 delirium during into: hyperactive (7), Among hyperactive patients, the levels of 6-SMT were
hospitalization hypoactive (10), mixed lower during the acute delirium than after recovery
(14) (p<0.001). Among hypoactive patients, 6-SMT levels
were higher during the acute delirium than after
recovery (p < 0.01). No difference in mixed.
Yoshitaka, Egi, 40 postoperative None; observational CAM There was no difference in preoperative melatonin
et al., 2013 patients in intensive concentration. The Δ melatonin concentration at
N = 40 care unit 1-hr post-op was significantly lower in patients with
delirium than in those without delirium (P =.036).
After adjustment of relevant confounders, Δ melatonin
concentration was independently associated with risk of
delirium (OR, 0.50; P =.047).

Melatonin and Melatonin Agonists (n = 222), comparing the effects of melatonin versus placebo
in Delirium Treatment among elderly patients undergoing hip arthroplasty under spi-
Several case reports have described the successful use of mela- nal anesthesia, a number of subjects developed postoperative
tonin in treating severe postoperative delirium unresponsive delirium and received postoperative melatonin for treatment
to antipsychotics or benzodiazepines (Hanania  & Kitain, (n = 62). It was found that 58% of subjects experienced delir-
2002). As part of the Sultan (2010) DBRPCT noted above ium resolution after treatment administration (Sultan, 2010).

Box 41.10 MELATONIN PHYSIOLOGICAL EFFECTS

• Melatonin plays important roles in multiple bodily functions, which may have implications for the development of delirium in the
medically ill:
• Chronobiotic effect (affecting aspects of biological time structure)
• Sleep–wake cycle regulatory effects
• Helps reset circadian rhythm disturbances
• Extensive antioxidant activity (with a particular role in the protection of nuclear and mitochondrial DNA)
• Extensive antiinflammatory activity
• Anti-nociceptive and analgesic effects
• Melatonin receptors appear to be important in mechanisms of learning and memory
• Inhibits the aggregation of the amyloid beta protein into neurotoxic microaggregates responsible for the neurofibrillary tangles
characteristics of Alzheimer’s disease, and it prevents the hyperphosphorylation of the tau protein.
Adapted from Maldonado (Maldonado, 2008a).
Darlington, L. G., C. M. Forrest, G. M. Mackay, R. A. Smith, A. J. Smith, N. Stoy and T. W. Stone (2010). “On the Biological Importance of the 3-hydroxyanthranilic
Acid: Anthranilic Acid Ratio.” Int J Tryptophan Res 3: 51–59.

41. D E L I R I U M • 889
Of note, a systematic literature search of all papers pub- In general, lack of sleep may lead to delirium, especially in
lished on the use of melatonin for the treatment in dementia patients at high risk (e.g., the elderly; the seriously medi-
revealed a significant improvement in sundowning/agitated cally ill); thus, promoting sleep is paramount. The first step,
behavior (de Jonghe, Korevaar, van Munster,  & de Rooij, as always, is prevention. Therefore, the implementation of
2010). This may be particularly important given the presence nonpharmacological sleep protocols as described above is
of circadian rhythm disturbances, like sundowning, in both important. If the patient still has difficulty sleeping, the use
dementia and delirium. Similarly, case reports have suggested of melatonin (or melatonin agonists) is probably the best
the usefulness of ramelteon, a melatonin agonist, in the treat- first pharmacological option. When this is not enough, clini-
ment of delirium (Kimura et al., 2011). cians should consider the use of trazodone, followed by mir-
Similarly, there are two case reports of the successful use tazapine. If needed, zolpidem may be an acceptable choice.
of ramelteon (a novel selective melatonin-receptor agonist) in Compared to conventional benzodiazepine agents, zolpidem
the treatment of patients with delirium (Kimura et al., 2011; may have a lower deliriogenic potential. Certainly, some of
Furuya et al., 2012). (See Table 41.21 for a summary of pub- the other delirium treatment alternatives already discussed
lished case reports and studies on the use of melatonin for the above may be beneficial for sleep promotion, such as the use
treatment of delirium.) of nighttime gabapentin or VPA or the use of low-dose que-
When it comes to sleep and delirium, the clinician tiapine. As a general rule, it is advisable to avoid the use of
must consider both risks and benefits of the intervention. diphenhydramine due to its anticholinergic effects. Although

Table 41.21 MELATONIN AND ANALOG FOR DELIR IUM MANAGEMENT

DELIR IUM
STUDYN = 7 POPULATION INTERVENTION DEFINITION R ESULTS

Bourne et al., 2008 S/p tracheostomy to Melatonin (MEL) Bispectral Melatonin use was associated with a 1-hour increase
N = 24 assist weaning from 10 mg PO @ 2000 index (BIS) in nocturnal sleep (P = 0.09) and a decrease in BIS
vent AUC indicating “better” sleep.
DBPCT
Melatonin use was associated with increased
nocturnal sleep efficiency.

Al-Aama, Brymer, Individuals aged ≥65 Randomized to MEL CAM Melatonin was associated with a lower risk of
et al., 2011 y/o admitted through 0.5 mg vs. PBO every delirium (12.0% vs. 31.0%, p = 0.014).
N = 145 the ED to a medical night for 14 days or
unit until discharge.

Sultan, 2010 ≥ 65 y/o scheduled Randomized to: Abbreviated The melatonin group showed a statistically
N = 300 for hip arthroplasty 1. PBO Mental Test significant decrease in the percentage of
under spinal (AMT) postoperative delirium to 9.43%.
anesthesia 2. Melatonin 5 mg
POD: PBO—32.7 %; MEL—9.4 % (p = 0.003);
3. Midazolam 7.5 mg MID—44 and (p = 0.245)
4. Clonidine 100 μg CLO—37.3 % (p = 0.629); Melatonin was
successful in treating 58.06% of patients who
suffered postoperative delirium.

de Jonghe Meta-analysis Various Nine papers, including RCTs (n = 243), and five case
et al., 2010 series (n = 87) were reviewed. Two of the RCTs found
Review a significant improvement on sundowning/agitated
behavior. All five case series found an improvement.

de Jonghe, van ≥65 y/o admitted for Randomized to: CAM Ongoing


Munster, et al., 2011 surgical repair of hip PBO
N = 452 fracture
Melatonin 3 mg @
2100

Kimura et al., 2011 Pts >59 y/o, medically Open label; ramelteon DSM-IV-TR All three cases demonstrated significant
N = 3 (case report) ill 8 mg q HS MDAS-Jap improvement in delirium scores as measured by
MDAS, with steady improvement over 7 days,
ramelteon 8 mg at HS.
Furuya et al., 2012 Mostly demented, Open label; ramelteon DSM-IV-TR Authors reported the successful treatment of five cases
N = 5 (case report) medically ill patients 8 mg of delirium within 1 day, after ramelteon 8 mg at HS.

Abbreviations: MEL=melatonin; BIS= Bispectral index (BIS) monitor.

890 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T
low-dose doxepin is an excellent alternative for many adult Finally, an open-label, prospective NRCT of patients
patients, it is unclear if it would still have enough anticholin- undergoing cardiac surgery compared placebo versus statin
ergic effect to make it wise avoid it. Certainly, the use of con- administration for prevention of delirium in elderly patients
ventional benzodiazepine agents is discouraged due to their (Katznelson et  al., 2009). The authors remarked that statin
inherit deliriogenic effect. administration has been shown to decrease morbidity and
mortality after cardiac and major noncardiac surgery, and
S L E E P R E S TOR AT ION I N DE L I R I U M
that animal and human studies have demonstrated the ben-
T R E AT M E N T
eficial effect of statins in central neural system injury. Thus
they theorized that the antithrombotic, antiinflammatory,
As described above, in delirium treatment we highly rec- and immunomodulatory properties of statins may be respon-
ommend starting with the implementation of nonpharma- sible for these protective effects, and that the same qualities
cological sleep protocols. If the patient still has difficulty may help protect against delirium. In fact, their study dem-
sleeping, the use of melatonin (or melatonin agonists) is onstrated that the administration of statins had a signifi-
probably the best first pharmacological option (e.g., 3 mg cant protective effect, reducing the odds of delirium by 46%
PO q, 2000; for prophylaxis). Several case reports have among patients 60  years old and older. Of interest, these
described the successful use of melatonin (e.g., 1–36 mg HS delirium-sparing effects did not manifest in patients younger
PO, for treatment) and melatonin agonists in treating severe than 60 years old.
postoperative delirium that is unresponsive to antipsychot-
ics or benzodiazepines (Sultan, 2010; Kimura et  al., 2011;
DE L I R I U M M A N AG E M E N T: S U M M A RY
Furuya et al., 2012).
If that fails, clinicians should consider the use of In summary, in order to prevent delirium, we recommend
non-benzodiazepine agents, such as trazodone or mirtazap- incorporating nonpharmacological and pharmacological
ine. If absolutely necessary, zolpidem may be an acceptable approaches. Key among the possible interventions is the mini-
choice, but take into consideration the moderately high mization of prolonged and uninterrupted sedation, by sched-
incidence of disordered sleep behaviors while on zolpidem uling daily sedation holidays and titration of sedation to a
and similar drugs. Given its short half-life and high seda- targeted goal, assisted by the use of sedation scales. The choice
tion effect, low-dose quetiapine may also be an acceptable of sedative and analgesic agents may also make a significant
short-term solution, especially in patients experiencing difference. It is important to be mindful of the additive effect
sundowning. that certain pharmacological agents may have, including
sedative, deliriogenic, and anticholinergic potential. A good
OT H E R AG E N T S T H AT H AV E S HOW N S OM E
prevention strategy is the routine inspection of the patient’s
PROM I S E I N DE L I R I U M M A N AG E M E N T
medication list, wise choice of agents (avoiding anticholiner-
gic and benzodiazepine agents, if possible) and eliminating
Some have theorized that an impaired serotonin metabolism all unnecessary agents. Remember the rule:  the greater the
may play a role in the development of delirium (Maldonado, number of pharmacological agents the patient is exposed to,
2008a,b). The original open-label study found that the anti- the greater the risk for delirium. The implementation of delir-
emetic agent ondansetron, a selective serotonin 5-HT3-type ium surveillance methods (such as any of the widely available
receptor antagonist, may be effective in the treatment of delirium detection tools) enables the early recognition of
“hyperactive delirium” (P = 0.001) (Bayindir, Guden Akpinar, delirium, and sometimes the detection of subsyndromal forms
Sanisoglu, & Sagbas, 2001). More recently, there was a second of delirium, allowing for prompt action before a full-blown
study (DBRCT; n = 80) comparing ondansetron (8 mg IV) syndrome has developed. Early mobilization is paramount.
to haloperidol (5 mg IV) in the treatment of post-cardiotomy As a rule, the more vertical and restriction-free a patient is,
“delirium.” The authors reported no significant baseline dif- the lower the incidence of delirium. It is important to elimi-
ference between the groups and similar clinical improve- nate restrictive devices and to reduce or eliminate sedation
ment, from delirious baseline, in both groups (p <0.01). and analgesia as soon as medically appropriate and safe, and
Unfortunately, this study did not use a standardized and to promote getting patients mobilized and out of bed as soon
validated delirium severity measure. The nonvalidated “scale” as possible. Once they are capable of leaving the bed, patients
used by these researchers (different in each case) appeared should be encouraged to eat all meals sitting at the bedside
to have been more a nonstandardized measure of agitation, chair rather than in bed. The implementation of a multicom-
rather than delirium. It is also not clear how the groups were ponent, nonpharmacological protocol (especially with train-
randomly assessed, and it is difficult to make sense pharma- ing and involvement of family and friends) may make a big
cologically of the time for assessment, given very different difference for many patients, their family, and hospital staff.
pharmacokinetics of both agents (e.g., after IV administra- When safe, family and staff should be encouraged to get the
tion, haloperidol’s half-life is 10–20 hrs., vs. ondansetron’s 5 patient out of the unit and to see that the patient receives as
hrs.; onset of action is also expected to be quite varied). Given much direct sunlight as possible. This will assist the process of
the nature of these studies, and the fact it did not rely on the normalizing the sleep–wake cycle naturally by manipulating
use of a validated diagnostic method, further, better designed the amount of ambient light. When this is not possible, con-
studies are needed. sider the use of melatonin or a melatonin-agonist, as well as

41. D E L I R I U M • 891
any of a whole host of nonpharmacological sleep-promoting Because delirium is common, it is important that physi-
interventions (e.g., avoid scheduling unnecessary blood draws cians be mindful and implement surveillance and monitoring
and vital sign checks in the middle of the night [if the patient’s strategies to allow for early detection and the implementation
medical condition allows it], minimize ambient noise and of techniques that may shorten its duration. When delirium
unnecessary light exposure, limit the amount of TV-watching occurs, it is important to immediately correct the underlying
after a certain time of day, and offer ear plugs and eye masks contributing causes and use treatment strategies directed at
to enhance the patient’s comfort). Melatonin and melatonin preventing harm to the patient and others and to shorten the
agonists may offer more benefit and fewer side effects than duration of the delirium episode, thus improving the odds of
conventional sedative agents. The selective use of prophylac- recovering baseline functional status.
tic antipsychotic agents should be considered, particularly in The psychosomatic medicine specialist has much to con-
patients at high risk or those with a prior history of develop- tribute to the diagnosis, treatment, and study of delirium. In
ing delirium in the medical environment. When delirium has the author’s experience, better results at implementing change,
already developed, the judicious use of dopamine antagonist fostering psychiatric consultation, and greater involvement of
agents seems to have the most evidence for a prompt recov- our services come from getting involved in multidisciplinary
ery. When it comes to the use of acetylcholinesterase inhibi- quality-improvement projects that involve physicians, nurses,
tors, the data available seem to suggest that they work only in and ancillary staff in the process of recognition and treat-
patients who have been on them long-term (i.e., for dementia). ment of delirious patients. Active participation in educational
Thus their use may be considered in patients with moderate to programs (e.g., continuing education, grand rounds, clinical
severe cognitive impairment who suffer from multiple medi- case conferences) and development of a multidisciplinary
cal and chronic medical issues and who are expected to repeat- task force charged with assessing the problem and develop-
edly return to the medical environment. Yet, data suggest ing an action plan are key to being recognized as an expert
that physostigmine is an underutilized therapeutic option we in the field, the one to go to for advice. At our institution, we
should consider in obvious cases of anticholinergic delirium. developed a “CNS Pharmacotherapy Task-Force” that devel-
oped protocols for the ICU in sedation, pain management,
neurological assessment, and delirium monitoring, preven-
tion, and treatment. Its success has led to this model being
S U M M A RY replicated throughout the institution, with a greater recogni-
tion of delirium as a problem and of psychosomatic medicine
Delirium is a neurobehavioral syndrome caused by the tran- specialists as the experts in the field. The author’s team has
sient disruption of normal neuronal activity secondary to implemented regular didactic sessions for nursing, regular
systemic disturbances, and is the most common psychiatric rounding sessions in the MICU and SICU, and a monthly
syndrome found in the general hospital setting. In addition “psychiatric mortality and morbidity” review in the internal
to causing distress to patients, families, and medical caregiv- medicine program that has been invaluable in helping the
ers, the development of delirium has been associated with medical staff change their perception and practice toward
increased morbidity and mortality, increased cost of care, delirium, leading to greater joint participation between psy-
increased hospital-acquired complications, poor functional chiatry and medico-surgical services.
and cognitive recovery, decreased quality of life, prolonged
hospital stays, and increased placement in specialized inter-
mediate and long-term care facilities.
C L I N IC A L   PE A R L S
Clinical experience and research data suggest that once
delirium has occurred, it is possible the patients may not
• Four important non-modifiable factors for delirium
return to their pre-delirium cognitive functional level. In fact,
are older age, baseline cognitive impairment, severity
data suggests delirium the occurrence of delirium is associated
of underlying medical illness, and preexisting mental
with an increased risk of dementia, and acceleration in the rate
disorders.
of cognitive decline in those demented patients who develop
delirium. Therefore, given increasing evidence that delirium • Among patients with preexisting mental disorders,
is not always reversible, and given the potentially numerous patients with bipolar disorder had the highest incidence
negative sequelae associated with its development, physicians of delirium.
must do everything possible to prevent its occurrence.
The data also suggest that there may be a “dose-effect” for • Opioids, corticosteroids, and benzodiazepines are major
delirium, meaning that the longer a patient is delirious, the contributors to delirium see Figure 41.7 (Gaudreau,
worse is the expected outcome. This is particularly true of Gagnon, et al., 2005b).
hypoactive delirium, which unfortunately is the most com- • A study of adult patients admitted to inpatient medicine
monly occurring type, particularly in the elderly. Therefore, wards showed that those who developed delirium had
clinicians are encouraged to implement as many prevention significantly elevated levels (i.e., above the detection
strategies as are available to them in order to improve patient limit) of IL-6 (53% versus 31%) and IL-8 (45% versus
outcome and prevent or minimize the occurrence of delirium, 22%), compared with patients who did not develop delir-
thus minimizing its morbidity. ium, even after adjusting for infection, age, and cognitive

892 • P S YC H I AT R I C C A R E O F T H E M E D I C A L PAT I E N T

You might also like