Chen 2018

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

The Gerontologist

cite as: Gerontologist, 2018, Vol. 00, No. 00, 1–11


doi:10.1093/geront/gny042
Advance Access publication May 17, 2018

Research Article

Surrogate Preferences on the Physician Orders for Life-


Sustaining Treatment Form
Elizabeth E. Chen, PhD,1,* Charles T. Pu, MD,2 Rachelle E. Bernacki, MD,3 Julia Ragland, MD,4
Jonathon H. Schwartz, MD,5 and Jan E. Mutchler, PhD6
The Gerontology Institute, University of Massachusetts Boston. 2Partners HealthCare, Boston, Massachusetts.
1

Psychosocial Oncology and Palliative Care, Dana Farber Cancer Institute, Boston, Massachusetts. 4Newton-Wellesley
3

Hospital, Newton, Massachusetts. 5Spaulding Hospital for Continuing Care, Cambridge, Massachusetts. 6Gerontology
Department and Institute, University of Massachusetts Boston.
*Address correspondence to Elizabeth E. Chen, PhD, Gerontology Institute, University of Massachusetts Boston, Boston, MA 02125. E-mail:
Elizabeth.E.Chen@gmail.com

Received: October 29, 2017; Editorial Decision Date: April 9, 2018

Decision Editor: Rachel Pruchno, PhD

Abstract
Background and Objectives: The purpose of this study is to compare treatment preferences of patients to those of surro-
gates on the Physician Orders for Life-Sustaining Treatment (POLST) forms.
Research Design and Methods: Data were collected from a sequential selection of 606 Massachusetts POLST (MOLST)
forms at 3 hospitals, and corresponding electronic patient health records. Selections on the MOLST forms were categorized
into All versus Limit Life-Sustaining Treatment. Multivariable mixed effects (grouped by clinician) logistic regression mod-
els estimated the impact of using a surrogate decision maker on choosing All Treatment, controlling for patient characteris-
tics (age, severity of illness, sex, race/ethnicity), clinician (physician vs non-physician), and hospital (site).
Results: Surrogates signed 253 of the MOLSTs (43%). A multivariable logistic regression model taking into consideration
patient, clinician, and site variables showed that surrogate decision makers were 60% less likely to choose All Treatment
than patients who made their own decisions (odds ratio = 0.39 [95% confidence interval = 0.24–0.65]; p < .001). This
model explained 44% of the variation in the dependent variable (Pseudo-R2 = 0.442; p < .001); mixed effects logistic regres-
sion grouped by clinician showed no difference between the models (LR test = 4.0e-13; p = 1.00).
Discussion and Implications: Our study took into consideration variation at the patient, clinician, and site level, and
showed that surrogates had a propensity to limit life-sustaining treatment. Surrogate decision makers are frequently needed
for hospitalized patients, and nearly all states have adopted the POLST. Researchers may want study decision-making pro-
cesses for patients versus surrogates when the POLST paradigm is employed.
Keywords: End-of-life care, Palliative care, Advance care planning

The Physician Orders for Life-Sustaining Treatment medical situations (National POLST Paradigm, 2017b). The
(POLST) paradigm offers a structured approach for physi- Institute of Medicine recommends the POLST when disease
cians, nurse practitioners, and physician assistants to dis- advances and when patients are facing their final year of
cuss and document patient preferences for life-sustaining life (Institute of Medicine, 2014). If a patient needs a sur-
treatments. The POLST is not a substitute for advance rogate decision maker, such as in situations when a patient
directives, which are meant to name a surrogate decision is deemed by a physician to be temporarily or permanently
maker and describe care preferences for future, unknown “not competent” to make medical decisions (Drane, 1984),

© The Author(s) 2018. Published by Oxford University Press on behalf of The Gerontological Society of America. 1
All rights reserved. For permissions, please e-mail: journals.permissions@oup.com.

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
2 The Gerontologist, 2018, Vol. 00, No. 00

the POLST reflects preferences resulting from discussions Mulvihill, & Cammer Paris, 1991; Shalowitz, Garrett-
between clinicians and patients’ surrogates. Mayer, & Wendler, 2006; Suhl, Simons, Reedy, & Garrick,
The POLST Form is a medical order and is transferrable 1994; Torke, Alexander, & Lantos, 2008). When there is
across care settings; this is particularly helpful given that no evidence of what a patient would have wanted, clini-
individuals in the last 6 months of life are often under the cians may apply the “best interest” standard, which is
care of more than 10 different physicians across multiple based on the “the pain and suffering associated with the
settings (Dartmouth Institute for Health Policy and Clinical intervention,” “the degree of and potential for benefit,”
Practice, 2014). Program efficacy studies indicate that the and “impairments that may result from the intervention”
POLST improved documentation in medical records about (American Medical Association, 2016).
end-of-life care, and that terminal interventions were largely There is evidence of strong correlation between proxy
consistent with documented preferences (Hickman et al., and self-reports for observed behaviors, such as walking
2009; Hickman et al., 2011; Richardson, Fromme, Zive, limitations after a stroke (Powell, Johnston, & Johnston,
Fu, & Newgard, 2014; Schmidt, Hickman, Tolle, & Brooks, 2007) and quality of life measures associated with physi-
2004; Tolle, Tilden, Nelson, & Dunn, 1998). The POLST cal functions, such as breathing, hearing, sleeping (Elliott,
paradigm has been, or is in the process of being imple- Lazarus, & Leeder, 2006). Socio-demographic characteris-
mented in 48 states (National POLST Paradigm, 2017a), tics and the type of chronic conditions can confound results,
and its use will likely grow in the coming years along with but controlling for these confounders has confirmed that
the numbers of individuals aging with chronic illness. proxy reports are accurate for observable behaviors, such
Previous research using POLST data showed that a as the ability to walk a certain distance, but potentially
majority of completed POLST records indicate a preference inaccurate for subjective behaviors, such as the ability to
to limit life-sustaining treatments (Fritz & Barclay, 2014; manage money or unobserved behaviors, such as difficulty
Fromme, Zive, Schmidt, Cook, & Tolle, 2014; Fromme, in using the toilet (Li, Harris, & Lu, 2015).
Zive, Schmidt, Olszewski, & Tolle, 2012; Hammes, Specific to patients nearing the end of life, family mem-
Rooney, Gundrum, Hickman, & Hager, 2012; Hickman, bers can be reliable proxy reporters of objective quality of
Keevern, & Hammes, 2014; Kim, Ersek, Bradway, & life measures, such as walking, nausea, and vomiting, but
Hickman, 2015; Schmidt, Zive, Fromme, Cook, & Tolle, with mixed results for subjective measures, such as fatigue
2014; Tarzian & Cheevers, 2017; Tolle et al., 1998) even or pain. Further, when there is a discrepancy, proxies tend
though the POLST allows patients or surrogates to indi- to report worse subjective quality of life than patients (Ferri
cate a preference to apply all treatment. One reason for the & Pruchno, 2009; Kirou-Mauro, Harris, Sinclair, Selby, &
large proportions preferring to limit treatment may be due Chow, 2006; Kutner, Bryant, Beaty, & Fairclough, 2006;
to clinicians’ interpretation of the need for a POLST form; Novella et al., 2001). In a sample of patients in an inpa-
that is, administering full treatment to sustain life is the tient palliative care unit, proxy family or physician reports
default situation, and POLST forms are thus only needed of patient symptoms were poorly correlated with patient
if individuals prefer treatment limitations. Another reason reports during early hospitalization (Day 3), improved over
may be attributed to study populations, largely comprised time (Day 6), but were still low (<35% agreement) on meas-
of non-Hispanic White individuals, who are more likely to ures of physical symptoms and psychological well-being
prefer treatment limitations than those from minority pop- (Jones et al., 2011). Family caregivers tended to overesti-
ulations (Rahman, Bressette, Gassoumis, & Enguidanos, mate the intensity or frequency of lack of energy, anxiety,
2016). Studies have not examined treatment preferences sadness, and pain distress compared with patients. On the
when surrogates or health care proxies make decisions other hand, physicians generally underestimated symptom
documented on POLST forms. burden compared with patients (Oechsle, 2013).
Almost one-half of hospitalized older adults use a sur-
rogate decision maker; and three out of four physicians in
a hospital sample reported that they have made a major Concept Model
decision with a surrogate in the past 30 days (Torke et al., This study is designed to estimate the odds that surrogate deci-
2014; Torke, Siegler, Abalos, Moloney, & Alexander, 2009). sion makers’ would choose aggressive life-sustaining treat-
Guidance from the American Medical Association (2016) ment, taking into account patient, clinician, and upstream
recommends that physicians guide surrogates to employ factors that might have channeled patients to sites of care.
the ethical principle of “substituted judgment,” that is, to We incorporated patient variables to take into account fac-
choose the treatment path the patient would have chosen tors that may influence whether a patient needed a surrogate,
if he or she were able. However, researchers have ques- such as age and severity of illness, which may also affect pref-
tioned the accuracy of substituted judgment for more than erences (Figure 1). On the other hand, clinician type, patient
25 years, and have shown in experimental vignettes that race/ethnicity, and sex were thought to affect only preferences.
surrogates are largely unable to predict patient decisions We acknowledge that prior decisions might have influenced
about life-sustaining treatments (Barrio-Cantalejo et al., a patient’s site of care, such as insurance limitations, patient
2009; Hinderer, Friedmann, & Fins, 2015; Seckler, Meier, or clinician preferences, and geography. These factors, in

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
The Gerontologist, 2018, Vol. 00, No. 00 3

practitioners, who signed nearly all (96%) of the forms;


another five clinicians outside of this palliative care practice
signed the remaining 4% of the forms at this location. Site 3
is an inpatient intensive palliative care unit with 12 beds in
a quaternary care academic medical center with more than
800 beds. This inpatient palliative care unit was staffed by
employed physician assistants under the direction of a small
number of palliative care physicians. Eleven different clini-
cians completed MOLSTs at Site 3.
Patient selection also differed from site to site, but all
patients had advanced illness and were nearing the end of
life. At Site 1 (LTACH), clinicians used the “surprise” ques-
tion, which has been validated in a number of settings as a
simple prognostic tool for patients with limited life expec-
tancy (Da Silva Gane, Braun, Stott, Wellsted, & Farrington,
2013; Javier et al., 2017; Moss et al., 2008; Vick, Pertsch,
Figure 1. Concept model—factors affecting life-sustaining treatment
preferences for hospitalized adults with limited life expectancy.
Hutchings, Neville, & Bernacki, 2016). That is, clinicians
asked themselves “Would I be surprised if this patient
died in the next 12 months?” A “no” triggered a “goals
addition to site-associated practice norms and prior advance
of care” conversation, followed by documentation on a
care plans, could affect patient or surrogate preferences for
MOLST form. Monthly tracking data over the first year of
life-sustaining treatments in ways that were not measured.
MOLST implementation at Site 1 showed that 15%–40%
of discharged patients completed MOLSTs. The palliative
Methods care consulting practice (Site 2) did not use the “surprise”
The Institutional Review Board at Partners HealthCare question for patients, but nearly all patients would have
approved this study, which uses 606 patient records from qualified had it been employed. Site 3, the inpatient pal-
three of its hospitals. Research Electronic Data Capture liative care unit, did not always use the “surprise” ques-
(REDCap), an electronic data capture tool hosted by the hos- tion because nearly all patients were expected to have less
pital system’s research management office, was used to col- than 1 year of remaining life. Clinicians completed MOLST
lect and manage study data (Harris, Taylor, Payne, Gonzalez, forms only for patients who were discharged to another
& Conde, 2009). De-identified data were exported for anal- care setting or to home, which accounted for approximately
ysis using STATA 13.1 (StataCorp. 2013, 2013). 50% of patients at this site. Despite differing patient selec-
tion criteria, patients at all sites with a completed MOLST
Setting Characteristics were expected to have limited life expectancy.
The three hospital sites were part of the same integrated care
network, and were among the first hospitals to implement
Massachusetts POLST (MOLST) in Eastern Massachusetts. Sample
As the first hospitals to implement MOLST, our data rep- The sample comprised a sequential selection of patients
resent de novo exposure to MOLST, and nearly eliminates who had completed the Massachusetts Medical Orders for
the possibility that patients had arrived with a form com- Life-Sustaining Treatment (MOLST) forms while inpatient
pleted at another hospital or in an outpatient setting. at three hospitals between July 9, 2012 and January 17,
Each site employed a slightly different staffing model, and 2014. The final analytic data set contained 593 MOLST
differed as to how broadly the MOLST was deployed. Site 1, forms after discarding 2% (N = 13 from 606) with miss-
a long-term acute care hospital (LTACH), was staffed largely ing clinician signatures, making the order invalid. Two-
by employed physicians and augmented by a few nurse hundred and eighty-eight (N = 288; 49%) of the analytic
practitioners; none were trained in palliative care. MOLST data set came from the LTACH (Site 1); 204 (34%) from
was implemented throughout the entire 180-bed hospital, the palliative care consulting practice at a community hos-
where all patients were admitted directly from an acute care pital (Site 2); and 101 (17%) from the inpatient intensive
hospital stay. LTACHs serve patients with severe illness, palliative care unit at the academic medical center (Site
who need hospital-level care for a long stay, which may 3). The sample comprised patients in both intensive care
last one or more months (American Hospital Association, (ICU) and non-intensive care units, and residing in both
2017; Center for Medicare and Medicaid Services, October long-term care settings and in the community. A small num-
2016). Thirty-one different clinicians signed MOLSTs at ber of patients completed multiple MOLST forms during
Site 1. Site 2 is a community hospital, where MOLSTs were 1 hospital stay, or completed another MOLST form dur-
largely administered by a consulting palliative care prac- ing a second or third hospitalization over the course of the
tice comprised of five palliative care physicians and nurse study period. In these circumstances, data were collected

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
4 The Gerontologist, 2018, Vol. 00, No. 00

only from the first MOLST, reserving the issue of changes Severity of Illness
to MOLST forms for another study. We used ICD-9 codes associated with the hospitalization
during which the MOLST form was signed to compute
the Charlson Score as an indicator of illness severity (Al
Dependent Variable Feghali et al., 2016; Chang et al., 2016; Tremblay, Arnsten,
The dependent variable (All vs Limit Treatment) was & Southern, 2016). The Charlson Weighted Index of
constructed from responses on Page 1 of the MOLST Comorbidity was developed as a prognostic indicator for
form, which provided a sequence of dichotomous “do individuals with chronic illnesses. Each chronic illness on the
not attempt” or “attempt” choices beginning with resus- Index has a score associated with mortality risk (1, 2, 3, or 6;
citation, followed by ventilation, then transfer to a hos- Charlson, Pompei, Ales, & MacKenzie, 1987). For example,
pital (Massachusetts Medical Orders for Life Sustaining uncomplicated diabetes is 1, diabetes with end organ damage
Treatment, 2014). Page 2 of the form seeks preferences is 2, moderate or severe liver disease is 3, and metastatic solid
about the duration of treatments on Page 1, and offers tumor is 6. Scores for comorbid conditions are summed. In
an opportunity to specify decisions regarding other life- the original Charlson validation cohort, 1-year mortality was
sustaining treatments, such as kidney dialysis and artificial 100% for patients who were hospitalized, survived to dis-
nutrition or hydration. Page 2 is not required in order for charge, and had a score greater than 5 (Charlson et al., 1987).
the MOLST to be valid, and too few were completed for Survival has improved since the original Charlson validation
meaningful analysis. cohort for many illnesses captured by its Index, but the tool
The pattern of responses generally followed logically remains useful in research as a measure for burden of illness
from the first choice, resuscitation (Figure 2). Ventilation (Crooks, West, & Card, 2015; D’Hoore, Sicotte, & Tilquin,
options were divided into two categories: invasive (intuba- 1993). Some studies have incorporated a decade-based age
tion) and noninvasive, such as continuous positive airway adjustment specified by Charlson to reflect additional mor-
pressure (CPAP); information about CPAP was not used tality risk associated with age beginning at age 50 (Charlson
due to the large number (N = 119; 20%) of forms miss- et al., 1987; Dias-Santos, Ferrone, Zheng, Lillemoe, &
ing these data. Patients were grouped into those who pre- Fernandez-del Castillo, 2015; Kaesmann, Janssen, Schild, &
ferred All Treatment (“Yes” to resuscitation, intubation, Rades, 2016; Lorenzon et al., 2017; St-Louis et al., 2015).
AND transfer) or Limit Treatment (“No” to resuscitation, Conceptually, this adjustment was meant to reflect that the
intubation, OR transfer). A separate analysis confirmed same illness in a 50-year-old and an 80-year-old represent
that there was no significant difference between the group
missing CPAP data and the group not missing CPAP data
for the primary variable of interest (surrogate vs patient
decisions; Pearson χ2 = 0.21; p = .644; Supplementary
Table 1).

Main Effect
Data for the main variable of interest (patient or surrogate
signature) were collected from the MOLST form, which
requests that signatories identify their relationship to the
patient. If the signatory did not respond, we compared the
patient name with the signature and associated printed
name. Forms with different signatory and patient names
were categorized into the surrogate group. All printed
names were legible.

Control Variables
Patient Demographics
Demographic data were obtained from patient electronic
health records, which included age at the time the MOLST
was signed and patient race/ethnicity. Logistic regression
models employed a categorical age variable approximat-
ing the bottom quartile (<60), interquartile range (60–73),
and top quartile (80 and over) at Site 1, where there was
the most variation between preferences for All versus Limit
Treatment. Figure 2. Pattern of responses on MOLST.

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
The Gerontologist, 2018, Vol. 00, No. 00 5

different risks for 1-year mortality. We did not include this palliative care practices were presented with an option for
adjustment, which is based on age decade, to allow for greater palliative care and accepted, which may be associated with a
precision using our age variable. willingness to accept limitations to life-sustaining treatment.

Clinician Type
Statistical Analysis
Clinician certification (physician vs nurse practitioner or
physician assistant) was collected from signatures on the We examined sample characteristics by site and by deci-
MOLST forms. If certification data were missing, then a sion maker using the Pearson χ2 test. Single-variable logis-
research assistant matched the signature against internal tic regression analyses were first performed to estimate the
staffing records. odds that each independent variable would predict a pref-
erence for All (vs Limit) Treatment. Multivariable logistic
Site regression models estimated the odds for All Treatment
The variables Site 1, Site 2, and Site 3 identified each hospital, first for Surrogate (vs Patient) Decision Makers, followed
and served as a means to hold constant site-associated prac- by patient characteristics, then concluding with the impact
tice variation, such as differing staffing models and the use of Site. Mixed effects logistic regression was performed
of palliative care specialists. The Site variable also took into using STATA’s xtmelogit command to account for covari-
account unmeasured factors that may have channeled patients ance due to clustering of patients within clinicians.
to each site, such as insurance coverage and patient or clini-
cian preferences farther upstream in the care continuum. For
Results
example, patients at the LTACH may not have been offered
palliative care, or were offered the service and refused; as a Baseline Patient Characteristics
result, these patients were transferred to the LTACH for high Nearly all patient characteristics differed between sites
intensity treatment. Patients under the care of clinicians in the (Table 1). Surrogates signed approximately 35% of the

Table 1. Sample Characteristics by Site and by Treatment Preference Group

MOLSTb signature

Site 1a Site 2a Site 3a Surrogate Patient


N = 593 (N = 288) (N = 204) (N = 101) (N = 253) (N = 340)

All (vs Limit) life-sustaining treatment 199 (69%) 5 (3%) 12 (12%) 58 (23%) 158 (46%)***
Surrogate decision maker 100 (35%) 119 (58%) 34 (34%) 253 (100%) —
Patient age: median (range) 68 (20–93) 83 (42–102) 60 (24–87) 80 (24–102) 68 (20–100)***
Patient age group
<60 69 (24%) 12 (6%) 50 (50%) 43 (17%) 88 (26%)**
60–79 152 (53%) 52 (25%) 41 (41%) 76 (30%) 169 (50%)***
80 and over 67 (23%) 140 (69%) 10 (10%) 134 (53%) 83 (24%)***
Charlson Scorec: median (range) 5 (0–12) 5 (0–15) 7 (2–10) 5 (0–12) 6 (0–15)*
Male (vs female) patient 162 (56%) 83 (41%) 48 (48%) 124 (49%) 169 (50%)
Race/ethnicity of patient
Non-Hispanic White 214 (74%) 191 (94%) 87 (86%) 282 (83%) 210 (83%)
Non-Hispanic Black 17 (6%) 1 (<1%) 5 (5%) 9 (4%) 14 (4%)
Non-Hispanic other and Hispanic 57 (20%) 12 (6%) 9 (9%) 34 (13%) 44 (13%)
Physician (vs NP or PA) 179 (62%) 115 (56%) 13 (4%) 135 (53%) 172 (51%)
Primary diagnoses (top 5)
Cancers 29 (11%) 87 (26%)***
Cardiovascular diseases 44 (17%) 45 (13%)
Respiratory disease/dysfunction 40 (17%) 52 (15%)
Infection/non-specific fever 34 (13%) 40 (12%)
Gastrointestinal disease/dysfunction 18 (7%) 18 (5%)

Notes: MOLST, Massachusetts Physician Orders for Life-Sustaining Treatment.


a
Site 1: 180-bed long-term acute care hospital with no palliative care specialist; Site 2: Palliative care consulting practice in 300-bed community acute care teach-
ing hospital; Site 3: Inpatient palliative care unit in 800+ bed academic medical center. bMassachusetts Medical Orders for Life-Sustaining Treatment Form.
c
Score derived from the Charlson Weighted Index of Comorbidities using discharge diagnoses (ICD-9) and does not incorporate the decade-based age adjustment
described in the original Charlson (1987) paper.
p-Value using the two-sample test of proportion comparing Surrogate and Patient groups: *≤0.05; **≤0.01; ***≤0.001.

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
6 The Gerontologist, 2018, Vol. 00, No. 00

MOLST forms at Sites 1 (LTACH without palliative care) OR = 0.26; p ≤ .001; Model 3 and 4), but site-associated
and 3 (palliative care at academic medical center) compared effects reduced the magnitude of this effect (OR = 0.59; p ≤
with 60% at Site 2 (palliative care at community hospital), .05; Model 5). Overall, patients and surrogates at the pal-
which also had the oldest patients. Only one in four surro- liative care sites had >90% lower odds for choosing All
gates chose All Treatment compared with nearly one in two Treatment than those at the LTACH (Site 2: odds ratio,
patients doing the same. Individuals at the academic medi- OR = 0.02 [95% confidence interval, CI, 0.01–0.05]; Site
cal center had higher burden of illness than the other two 3: OR = 0.06 [95% CI 0.02–0.10]; p ≤ .001).
sites as measured by the Charlson Score (median for Sites A multivariable mixed effects logistic regression model,
1 and 2 = 5; Site 3 = 7). Patient primary diagnoses were grouped by clinicians, using all observations and the full
not different between surrogate and patient-signed forms complement of independent variables was no different from
with the exception of cancers, where a larger proportion of a model that did not cluster patients by clinician (Standard
patient signers had cancers (26%) compared with patients Deviation of the Intercept = 2.58e-10 [95% CI 0.00–0.00];
who had surrogates (11%; p < .001). Likelihood Ratio Test = 4.0e-13; p = 1.00).

Multivariable Logistic Regression Post Hoc Logistic Regression Using Only Data
Surrogate decision makers were much less likely to choose from the LTACH
All Treatment than patients who made their own deci- We tested our model post hoc using only data from the
sions (OR = 0.34–0.46; p ≤ .001). This effect remained LTACH (Site 1) to eliminate the potential for confounding
unchanged as we added patient, clinician, and site char- due to clinicians channeling patients into palliative care set-
acteristics (Table 2; Models 1–5). Patients in the highest tings due to prior patient or surrogate preferences to limit
age group (≥80) were consistently less likely to choose life-sustaining treatments (Table 3). Using only LTACH
All Treatment than those in the lowest age group (<60; data, surrogate decision makers experienced 52% lower
OR = 0.14–0.22; p ≤ .001; Models 2–5). Individuals with odds for choosing All Treatment compared with patient
high Charlson Scores (>5) were less likely to choose All decision makers (OR = 0.48 [95% CI 0.28–0.84]; p = .01).
Treatment than those with lower Charlson Scores (<5; These results were substantially the same as those using

Table 2. Odds Ratios (95% Confidence Interval) for Logistic Regression Models Estimating the Odds for All Treatment versus
Limit Treatment on the Massachusetts Medical Orders for Life-Sustaining Treatment Form

N = 593 Model 1 Model 2 Model 3 Model 4 Model 5

Surrogate (vs patient 0.34*** (0.24–0.49) 0.46*** (0.31–0.67) 0.39*** (0.26–0.58) 0.36*** (0.24–0.55) 0.39*** (0.24–0.65)
signature)
Age group: (reference <60)
60–79 0.76 (0.49–1.18) 0.74 (0.47–1.16) 0.65+ (0.41–1.05) 0.41** (0.21–0.78)
80 and over 0.22*** (0.13–0.37) 0.15*** (0.09–0.26) 0.14*** (0.08–0.24) 0.19*** (0.09–0.41)
Charlson Scorea >5 0.26*** (0.17–0.39) 0.27*** (0.18–0.41) 0.59* (0.36–0.99)
(vs Charlson≤5)
Male (vs female patient) 1.71** (1.16–2.53) 1.20 (0.75–1.94)
Race/ethnicity of
patient(reference:
non-Hispanic White)
Non-Hispanic Black 2.59* (1.01–6.65) 1.37 (0.49–3.81)
 Non-Hispanic other 2.10** (1.19–3.69) 1.12 (0.59–2.12)
and Hispanic
Physician (vs NP or PA) 1.73** (1.17–2.57) 0.98 (0.59–1.63)
Site (reference site 1)
Site 2b 0.02*** (0.01–0.05)
Site 3b 0.04*** (0.02–0.10)
Pseudo-R2c 0.046++++ 0.103++++ 0.163++++ 0.195++++ 0.442++++

Notes: aScore derived from the Charlson Weighted Index of Comorbidities using discharge diagnoses (ICD-9) and does not incorporate the decade-based age
adjustment described in the original Charlson (1987) paper. bSite 1: 180-bed long-term acute care hospital with no palliative care specialist; Site 2: Palliative care
consulting practice in 300-bed community acute care teaching hospital; Site 3: Inpatient palliative care unit in 800+ bed at an academic medical center. cPseudo-R2
is an indicator of model fit and can be interpreted as the amount of variation in All versus Limit Treatment attributed to each variable.
+
p-Value of odds ratio ≤.10; *≤.05; **≤.01; ***≤.001. ++++p-Value of the model ≤.001.

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
The Gerontologist, 2018, Vol. 00, No. 00 7

Table 3. Post Hoc Analysis – Odds Ratios and (95% care practice. This post hoc analysis revealed essentially the
Confidence Interval) for a Multivariable Logistic Regression same result, that surrogate decision makers were 52% less
Model Estimating the Odds for All Treatmenta versus Limit likely to choose All Treatment than patients not using a
Treatment on the Massachusetts Medical Orders for Life- surrogate.
Sustaining Treatment Form Using Only Site 1b Data Our results are consistent with findings using POLST
data from a nursing home population (Rahman et al., 2016),
Odds ratio
which also used data gathered from a clinical setting. Other
N = 288 (95% CI) p value
studies that examined both patient and surrogate decisions
Surrogate (vs patient signature) 0.48 (0.28–0.84) .010 are not directly comparable because study methods were
Charlson Score >5c 0.56 (0.32–0.97) .038 intended to test the accuracy of surrogate decisions, and
Age: (reference ≤60) used hypothetical scenarios (Barrio-Cantalejo et al., 2009;
60–79 0.33 (0.15–0.73) .006 Shalowitz et al., 2006), which may not reflect contextual
80 and over 0.16 (0.07–0.40) ≤.001 cues that could influence patient or surrogate decisions
Male 1.14 (0.66–1.96) .638 when life-sustaining treatment preferences are sought in
Race/ethnicity (reference = clinical settings (Reamy, Kim, Zarit, & Whitlatch, 2011).
non-Hispanic White) Surrogate decision makers face intense emotional distress
Non-Hispanic Black 1.70 (0.54–5.33) .364 as they balance forces favoring more intervention against
Non-Hispanic other and Hispanic 1.05 (0.53–2.07) .889
those favoring less. Prior research suggest that negative cues
Physician (vs NP/PA signature) 1.06 (0.62–1.85) .815
are most salient when surrogates need to make decisions,
such as a patient’s unconscious state or perceived pain, and
Notes: aAll Treatment = “Yes” to All (resuscitate, intubate, transfer) versus
Limit Treatment = “No” to At Least One (resuscitate, intubate, transfer). a focus on what patients do not want instead of what they
b
Site 1 is a 180-bed long-term acute care hospital with no palliative care spe- do want (Dionne-Odom, 2015).
cialist. cScore derived from the Charlson weighted Index of Comorbidities The process of discussing end-of-life treatments with
using discharge diagnoses (ICD-9) and does not incorporate the decade-based surrogates is a delicate balance between prior patient pref-
age adjustment described in the original Charlson (1987) paper.
erences and the best interests of the patient going forward
(Torke, Moloney, Siegler, Abalos, & Alexander, 2010). We
data pooled for all three sites. Mixed effects multivariable should not assume that our results reflect prior discussions
logistic regression, grouped by 31 clinicians again showed between surrogate and patient pairs, as these pairs have
no difference between this model and a comparable model shown low agreement (62%) when asked whether dis-
that did not cluster patients within clinicians (Standard cussions about life-sustaining treatments had taken place
Deviation of Intercept: 1.58e-09 [95% CI 0.00–0.00]; (Fried, Redding, Robbins, O’Leary, & Iannone, 2011), and
p = 1.00). prior discussions also showed no impact on the accuracy
of substituted judgment (Pruchno, Lemay Jr., Field, &
Levinsky, 2005). Rather, our results seem to indicate that
Discussion surrogate decision makers are systematically different from
The goal of this study was to estimate the odds that surro- patient decision makers, which would be consistent with
gate decision makers would choose aggressive life-sustain- prior research showing that proxy respondents represented
ing treatment decisions by taking into consideration patient, individuals in worse physical health, older, and were less
clinician, and site variables, which accounted for unmeas- educated than those not using a proxy (Elliott, Beckett,
ured factors that channeled patients to each location, such Chong, Hambarsoomians, & Hays, 2008).
as patient insurance and patient or clinician preferences Prior research about the accuracy of proxy reporting
farther upstream in the care continuum. We collected data suggests that surrogates tend to overestimate the burden
from the Massachusetts version of the POLST form and of illness, especially on subjective quality of life measures
corresponding patient medical records at three hospitals in (Ferri & Pruchno, 2009; Kirou-Mauro et al., 2006; Kutner
the greater Boston area. An important strength of this study et al., 2006; Novella et al., 2001). This tendency can be
is that data represent decisions made in a health care set- problematic whether clinicians guide surrogates to employ
ting in the context of patients and families experiencing an the principle of “substituted judgment” or “best interest”
event associated with serious, life-limiting illness. when making life-sustaining treatment decisions. Clinicians
Surrogates signed 43% of the MOLST forms in our may want to take into consideration that in some situa-
study, which is consistent with prior results (47%; Torke tions, surrogates may overestimate burden, such as when
et al., 2014). Multivariable logistic regression results life-sustaining treatment decisions are based on unobserv-
showed that surrogates were 60% less likely to choose able or subjective quality of life outcomes for the patient.
All Treatment than patients who did not use a surrogate. Because the POLST is recommended for patients facing
Because decisions in palliative care settings may have con- their last year of life, clinicians may want to attend to com-
founded the results, we performed a post hoc analysis using pleting a POLST when patients are still able to make their
only data from the LTACH, which did not have a palliative own decisions and a surrogate is not yet needed.

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
8 The Gerontologist, 2018, Vol. 00, No. 00

Limitations illness. Further, dementia is assigned only a score of 1 on


Our study used a sample that completed MOLSTs at three the Charlson Index, with no further adjustment for higher
hospitals; thus, selection bias cannot be ignored. In addi- severity, even though the Index adjusts for higher severity
tion to upstream forces that channeled individuals to these of cancers, diabetes, and heart disease. This limitation in the
hospitals, clinicians at each location selected patients for Charlson Score, coupled with other research showing insuf-
MOLST, and patients or surrogates agreed to complete ficient reporting or documentation of underlying dementia
a MOLST form. We acknowledge that our data failed in older populations (Crowther, Bennett, & Holmes, 2017;
to capture patients and surrogates who were not asked van den Dungen et al., 2012; Ostbye, Taylor, Clipp, Van
to complete a MOLST form. Some clinicians may have Scoyoc, & Plassman, 2008), and a propensity for surro-
been uncomfortable discussing end-of-life care goals with gates to favor limitations when dementia is present (Feltz
patients or surrogates, whereas others may have thought & Samayoa, 2012), indicate that future studies should
their patients would survive beyond 1 year, and still others incorporate multiple measures to reflect conditions faced in
may have decided not to complete MOLST forms because the last days or weeks of life.
patients and surrogates wanted All Treatment. These biases Finally, our mixed effects models grouped by clinicians
in the data set would likely result in odds ratios being too did not show variation between clinicians even though
low for surrogates. By the same token, patients and sur- prior research point to an association between patient pref-
rogates who chose not to complete a MOLST form were erences and the clinicians’ end-of-life care preferences for
choosing All Treatment by default, and because we did themselves (Wilkinson & Truog, 2013). A larger sample
not track refusals, our estimates regarding a preference may have yielded different results. A review of the numbers
to Limit Treatment may be high. Furthermore, we lacked of MOLST forms completed by each of the 31 clinicians at
background data about patients, which would have predis- Site 1 showed that close to 50% completed just one or two
posed them to treatment versus treatment limitations, such forms while 30% completed 11–36 forms. A larger sample,
as religiosity or prior declarations in advance care directive. collected over a longer period of time, would have provided
While we have no evidence pointing to underlying system- more MOLST forms for the large proportion of clinicians
atic selection bias in our sample, we also have no evidence who had only completed one or two forms, and may have
to support a lack of systematic selection bias. yielded different results.
It is reasonable to assume that clinicians and patients
in the palliative care settings were already biased against
Conclusion
choosing All Treatment. Patients or surrogates agreed to
treatment in these settings, and the definition of palliative The results in this study show that surrogate decision mak-
care is to treat symptoms, not disease. The post hoc analysis ers for a hospitalized population with serious and advanc-
using data from the LTACH, which did not have a pallia- ing illness were more likely to indicate a preference to limit
tive care practice, allowed us to eliminate the potential for life-sustaining treatments than patient decision makers. In
this bias by examining results at one location where there multivariable logistic regression models, this effect operated
was substantial variation in the dependent variable; that independently from age, practice patterns associated with
is, 70% chose All Treatment. The result for the main effect site, and the Charlson Score. Prior research demonstrating
using single-site data was substantially the same as that that proxies tend to overestimate symptom burden com-
using the full data set, which suggests that whatever forces pared with patients presents a troubling ethical problem in
led patients to the palliative care practices as opposed to our results, which is whether decisions to limit treatment
the LTACH did not change our conclusion. If anything, one by surrogates can be construed as an accurate reflection of
would expect that clinicians in an LTACH might approach what the patient would have wanted. Given the frequency
goals of care discussions in a manner that could bias patient with which surrogates are needed for older, hospitalized
or surrogate decisions toward aggressive treatment. Again, patients, provider organizations may want to incorporate
our single-site results do not support this direction of think- new strategies to reach surrogates as an additional patient
ing; rather, they confirm that even in an aggressive treat- decision partner. Providers may also want to consider dis-
ment pathway, surrogate decision makers are less likely to cussing with patients and documenting their preferences
choose all life-sustaining treatments than patients. for temporary versus permanent incapacity. Finally, when
Using only one scale, the Charlson Score, to measure communicating with surrogate decision makers, providers
severity of illness is also an important limitation in this may want to keep in mind a surrogate’s tendency to overes-
study. Our regression models showed that the surrogate timate symptom burden..
variable acted independently from the Charlson Score,
which suggests that the Charlson Index did not fully meas-
ure conditions that would lead to needing a surrogate deci-
Supplementary Data
sion maker, such as delirium, high fever, or nausea that Supplementary data are available at The Gerontologist
often lead to hospitalization for individuals with advanced online.

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
The Gerontologist, 2018, Vol. 00, No. 00 9

Acknowledgments Crowther, G. J., Bennett, M. I., & Holmes, J. D. (2017). How well
are the diagnosis and symptoms of dementia recorded in older
Elizabeth E. Chen had full access to all the data in the study and
patients admitted to hospital? Age and Ageing, 46, 112–118.
takes responsibility for the integrity of the data and the accuracy
doi:10.1093/ageing/afw169
of the data analysis. A portion of the results in this article was pre-
Da Silva Gane, M., Braun, A., Stott, D., Wellsted, D., & Farrington,
sented at the Gerontological Society of America Research Meeting
in New Orleans, November 2016, at the International Association K. (2013). How robust is the ‘surprise question’ in predicting
of Gerontology and Geriatrics World Congress in San Francisco, short-term mortality risk in haemodialysis patients? Nephron.
July 2017, and via webinar hosted by the National POLST on Clinical Practice, 123, 185–193. doi:10.1159/000353735
January 25, 2018. We gratefully acknowledge the contributions Dartmouth Institute for Health Policy and Clinical Practice. (2014).
of Mary Fairbanks, BS, who generously volunteered hundreds of Care of chronic illness in last two years of life: Number of different
hours during the data collection phase of this study; Edward Alan physicians seen per decedent in the last 6 months of life. Retrieved
Miller, PhD, MPA, who provided important critique for earlier ver- October 22, 2017, from http://www.dartmouthatlas.org/data/
sions of this research; and Susan T. Moore, RN, MPH for facilitat- table.aspx?ind=12&loct=2&loc=2%2c3%2c4%2c5%2c6%2c7
ing the data collection process and providing important comments %2c8%2c9%2c10%2c11%2c12%2c13%2c14%2c15%2c16%
to the final draft. 2c17%2c18%2c19%2c20%2c21%2c22%2c23%2c24%2c25%
2c26%2c27%2c28%2c29%2c30%2c31%2c32%2c33%2c34%
2c35%2c36%2c37%2c38%2c39%2(TRUNCATED)
References D’Hoore, W., Sicotte, C., & Tilquin, C. (1993). Risk adjust-
Al Feghali, K. A., Robbins, J. R., Mahan, M., Burmeister, C., Khan, N. ment in outcome assessment: The Charlson comorbidity
T., Rasool, N.,…Elshaikh, M. A. (2016). Predictive capacity of 3 index. Methods of Information in Medicine, 32, 382–387.
comorbidity indices in estimating survival endpoints in women doi:10.1055/s-0038-1634956
with early-stage endometrial carcinoma. International Journal Dias-Santos, D., Ferrone, C. R., Zheng, H., Lillemoe, K. D., &
of Gynecological Cancer: Official Journal of The International Fernández-Del Castillo, C. (2015). The Charlson age comor-
Gynecological Cancer Society, 26, 1455–1460. doi:10.1016/j. bidity index predicts early mortality after surgery for pancre-
ygyno.2016.04.205. atic cancer. Surgery, 157, 881–887. doi:10.1016/j.surg.2014.
American Hospital Association. (2017). Long-term care hospitals 12.006
(LTCHs). Retrieved October 22, 2017, from http://www.aha. Dionne-Odom, J. N., Willis, D. G., Bakitas, M., Crandall, B., & Grace,
org/advocacy-issues/postacute/ltach/index.shtml P. J. (2015). Conceptualizing surrogate decision making at end of
American Medical Association. (2016). Code of Medical Ethics life in the intensive care unit using cognitive task analysis. Nursing
Opinions on Consent, Communications and Decision Making Outlook, 63, 331–340. doi:10.1016/j.outlook.2014.10.004
Section 2.2.1. Retrieved March 3, 2018, from https://www. Drane, J. F. (1984). Competency to give an informed consent.
ama-assn.org/sites/default/files/media-browser/code-of-medical- A model for making clinical assessments. JAMA, 252, 925–927.
ethics-chapter-2.pdf doi:10.1001/jama.1984.03350070043021
Barrio-Cantalejo, I. M., Molina-Ruiz, A., Simón-Lorda, P., Cámara- van den Dungen, P., van Marwijk, H. W., van der Horst, H. E., Moll
Medina, C., Toral López, I., del Mar Rodríguez del Aguila, M., van Charante, E. P., Macneil Vroomen, J., van de Ven, P. M.,
& Bailón-Gómez, R. M. (2009). Advance directives and prox- & van Hout, H. P. (2012). The accuracy of family physicians’
ies’ predictions about patients’ treatment preferences. Nursing dementia diagnoses at different stages of dementia: A system-
Ethics, 16, 93–109. doi:10.1177/0969733008097995 atic review. International Journal of Geriatric Psychiatry, 27,
Center for Medicare and Medicaid Services. (October 2016). Long 342–354. doi:10.1002/gps.2726
term care hospital prospective payment system. Retrieved Elliott, D., Lazarus, R., Leeder, S. R. (2006). Proxy respondents relia-
October 22, 2017, from https://www.cms.gov/Outreach-and- bly assessed the quality of life of elective cardiac surgery patients.
Education/Medicare-Learning-Network-MLN/MLNProducts/ Journal of Clinical Epidemiology, 59, 153–159. doi:10.1097/
Downloads/Long-Term-Care-Hospital-PPS-Fact-Sheet- DCC.0000000000000097
ICN006956.pdf Elliott, M. N., Beckett, M. K., Chong, K., Hambarsoomians, K., &
Chang, C., Yin, W., Wei, C., Wu, C., Su, Y., Yu, C., Lee, C.C., et al. Hays, R. D. (2008). How do proxy responses and proxy-assisted
(2016). Adjusted age-adjusted Charlson Comorbidity Index responses differ from what Medicare beneficiaries might have
score as a risk measure of perioperative mortality before can- reported about their health care? Health Services Research, 43,
cer surgery. PLoS One, 11, e0148076. doi:10.1371/journal. 833–848. doi:10.1111/j.1475-6773.2007.00820.x
pone.0148076 Feltz, A., & Samayoa, S. (2012). Heuristics and life-sustaining treat-
Charlson, M. E., Pompei, P., Ales, K. L., & MacKenzie, C. R. (1987). ments. Journal of Bioethical Inquiry, 9, 443–455. doi:10.1007/
A new method of classifying prognostic comorbidity in longitu- s11673-012-9396-5
dinal studies: Development and validation. Journal of Chronic Ferri, C. V., & Pruchno, R. A. (2009). Quality of life in end-
Diseases, 40, 373–383. doi:10.1016/0021-9681(87)90171-8 stage renal disease patients: Differences in patient and
Crooks, C. J., West, J., & Card, T. R. (2015). A comparison of the spouse perceptions. Aging & Mental Health, 13, 706–714.
recording of comorbidity in primary and secondary care by doi:10.1080/13607860902845558
using the Charlson Index to predict short-term and long-term Fried, T. R., Redding, C. A., Robbins, M. L., O’Leary, J. R., &
survival in a routine linked data cohort. BMJ Open, 5, e007974. Iannone, L. (2011). Agreement between older persons and their
doi:10.1136/bmjopen-2015-007974 surrogate decision-makers regarding participation in advance

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
10 The Gerontologist, 2018, Vol. 00, No. 00

care planning. Journal of the American Geriatrics Society, 59, Kaesmann, L., Janssen, S., Schild, S. E., & Rades, D. (2016). Value of
1105–1109. doi:10.1111/j.1532-5415.2011.03412.x comorbidity scales for predicting survival after radiochemother-
Fritz, Z. B., & Barclay, S. I. (2014). Patients’ resuscitation prefer- apy of small cell lung cancer. Lung, 194, 295–298. doi:10.1007/
ences in context: Lessons from POLST. Resuscitation, 85, 444– s00408-016-9857-4
445. doi:10.1016/j.resuscitation.2014.01.016 Kim, H., Ersek, M., Bradway, C., & Hickman, S. E. (2015). Physician
Fromme, E. K., Zive, D., Schmidt, T. A., Cook, J. N., & Tolle, S. W. orders for life-sustaining treatment for nursing home residents
(2014). Association between physician orders for life-sustain- with dementia. Journal of the American Association of Nurse
ing treatment for scope of treatment and in-hospital death in Practitioners, 27, 606–614. doi:10.1002/2327-6924.12258
Oregon. Journal of the American Geriatrics Society, 62, 1246– Kirou-Mauro, A., Harris, K., Sinclair, E., Selby, D., & Chow, E.
1251. doi:10.1111/jgs.12889 (2006). Are family proxies a valid source of information about
Fromme, E. K., Zive, D., Schmidt, T. A., Olszewski, E., & Tolle, S. cancer patients’ quality of life at the end-of-life? A literature
W. (2012). POLST Registry do-not-resuscitate orders and other review. Journal of Cancer Pain & Symptom Palliation, 2, 23–33.
patient treatment preferences. JAMA, 307, 34–35. doi:10.1001/ doi:10.1300/J427v02n02_04
jama.2011.1956 Kutner, J. S., Bryant, L. L., Beaty, B. L., & Fairclough, D. L.
Hammes, B. J., Rooney, B. L., Gundrum, J. D., Hickman, S. E., & (2006). Symptom distress and quality-of-life assessment at
Hager, N. (2012). The POLST program: A retrospective review the end of life: The role of proxy response. Journal of Pain
of the demographics of use and outcomes in one community and Symptom Management, 32, 300–310. DOI:10.1016/j.
where advance directives are prevalent. Journal of Palliative jpainsymman.2006.05.009
Medicine, 15, 77–85. doi:10.1089/jpm.2011.0178 Li, M., Harris, I., and Lu, Z. K. (2015). Differences in proxy-reported
Harris, P. A., Taylor, R., Thielke, R., Payne, J., Gonzalez, N., and patient-reported outcomes: Assessing health and functional
& Conde, J. G. (2009). Research electronic data capture status among medicare beneficiaries. BMC Medical Research
(REDCap)–a metadata-driven methodology and workflow pro- Methodology, 15, 62. doi:10.1186/s12874-015-0053-7
cess for providing translational research informatics support. Lorenzon, L., Costa, G., Massa, G., Frezza, B., Stella, F., & Balducci,
Journal of Biomedical Informatics, 42, 377–381. doi:10.1016/j. G. (2017). The impact of frailty syndrome and risk scores on
jbi.2008.08.010 emergency cholecystectomy patients. Surgery Today, 47, 74–83.
Hickman, S., Keevern, E., & Hammes, B. J. (2014). The effect of the doi:10.1007/s00595-016-1361-1
POLST (Physician Orders for Life-Sustaining Treatment) pro- Massachusetts Medical Orders for Life Sustaining Treatment.
gram on clinical care: A systematic review. Journal of Palliative (2014). About MOLST. Accessed October 22, 2017 http://
Care, 30, 226–227. doi:10.1089/jpm.2008.0196 molst-ma.org/about
Hickman, S. E., Nelson, C. A., Moss, A. H., Hammes, B. J., Terwilliger, Moss, A. H., Ganjoo, J., Sharma, S., Gansor, J., Senft, S., Weaner,
A., Jackson, A., & Tolle, S. W. (2009). Use of the Physician B.,…Schmidt, R. (2008). Utility of the “surprise” question to
Orders for Life-Sustaining Treatment (POLST) paradigm pro- identify dialysis patients with high mortality. Clinical Journal
gram in the hospice setting. Journal of Palliative Medicine, 12, of the American Society of Nephrology: CJASN, 3, 1379–1384.
133–141. doi:10.1089/jpm.2008.0196 doi:10.2215/CJN.00940208
Hickman, S. E., Nelson, C. A., Moss, A. H., Tolle, S. W., Perrin, National POLST Paradigm. (2017a). Programs in your state. Accessed
N. A., & Hammes, B. J. (2011). The consistency between October 22, 2017 http://polst.org/programs-in-your-state/
treatments provided to nursing facility residents and orders National POLST Paradigm. (2017b). POLST & advance
on the physician orders for life-sustaining treatment form. directives. Accessed December 17, 2017 http://polst.org/
Journal of the American Geriatrics Society, 59, 2091–2099. advance-care-planning/polst-and-advance-directives/
doi:10.1111/j.1532-5415.2011.03656.x Novella, J. L., Jochum, C., Jolly, D., Morrone, I., Ankri, J., Bureau, F.,
Hinderer, K. A., Friedmann, E., Fins, J. J. (2015). Withdrawal & Blanchard, F. (2001). Agreement between patients’ and prox-
of life-sustaining treatment: patient and proxy agreement. ies’ reports of quality of life in Alzheimer’s disease. Quality of
Dimensions of Critical Care Nursing, 34, 91–99. doi:10.1097/ Life Research, 10, 443–452. doi:10.1023/A:1012522013817
DCC.0000000000000097 Oechsle, K., Goerth, K., Bokemeyer, C., & Mehnert, A. (2013).
Institute of Medicine. (2014). Dying in America: Improving Symptom burden in palliative care patients: Perspectives of
quality and honoring individual preferences near the patients, their family caregivers, and their attending physi-
end of life. Washington, DC: National Academies Press. cians. Supportive Care in Cancer, 21, 1955–1962. doi:10.1007/
doi:10.17226/18748 s00520-013-1747-1
Javier, A. D., Figueroa, R., Siew, E. D., Salat, H., Morse, J., Stewart, Ostbye, T., Taylor, D. H. Jr, Clipp, E. C., Scoyoc, L. V., &
T. G.,…Abdel-Kader, K. (2017). Reliability and utility of the Plassman, B. L. (2008). Identification of dementia: Agreement
surprise question in CKD stages 4 to 5. American Journal of among national survey data, medicare claims, and death cer-
Kidney Diseases: The Official Journal of the National Kidney tificates. Health Services Research, 43(1 Pt 1), 313–326.
Foundation, 70, 93–101. doi:10.1053/j.ajkd.2016.11.025 doi:10.1111/j.1475-6773.2007.00748.x
Jones, J. M., McPherson, C. J., Zimmermann, C., Rodin, G., Le, L. Powell, R., Johnston, M., & Johnston, D. W. (2007). Assessing
W., & Cohen, S. R. (2011). Assessing agreement between termi- walking limitations in stroke survivors: Are self-reports and
nally ill cancer patients’ reports of their quality of life and family proxy-reports interchangeable? Rehabilitation Psychology, 52,
caregiver and palliative care physician proxy ratings. Journal of 177–183. doi:10.1037/0090-5550.52.2.177
Pain and Symptom Management, 42, 354–365. doi:10.1016/j. Pruchno, R. A., Lemay, E. P. Jr, Feild, L., & Levinsky, N. G.
jpainsymman.2010.11.018 (2005). Spouse as health care proxy for dialysis patients:

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018
The Gerontologist, 2018, Vol. 00, No. 00 11

Whose preferences matter? The Gerontologist, 45, 812–819. support is unreliable. Archives of Internal Medicine, 154, 90–96.
doi:10.1093/geront/45.6.812 doi:10.1001/archinte.154.1.90
Rahman, A. N., Bressette, M., Gassoumis, Z. D., & Enguidanos, S. Tarzian, A. J., & Cheevers, N. B. (2017). Maryland’s medical orders
(2016). Nursing home residents’ preferences on physician orders for life-sustaining treatment form use: Reports of a statewide
for life sustaining treatment. The Gerontologist, 56, 714–722. survey. Journal of Palliative Medicine, 20, 939–945. doi:10.1089/
doi:10.1093/geront/gnv019 jpm.2016.0440
Reamy, A. M., Kim, K., Zarit, S. H., & Whitlatch, C. J. (2011). Tolle, S. W., Tilden, V. P., Nelson, C. A., & Dunn, P. M. (1998). A pro-
Understanding discrepancy in perceptions of values: Individuals spective study of the efficacy of the physician order form for life-
with mild to moderate dementia and their family caregivers. The sustaining treatment. Journal of the American Geriatrics Society,
Gerontologist, 51, 473–483. doi:10.1093/geront/gnr010 46, 1097–1102. doi:10.1111/j.1532-5415.1998.tb06647.x
Richardson, D. K., Fromme, E., Zive, D., Fu, R., & Newgard, C. D. Torke, A. M., Alexander, G. C., & Lantos, J. (2008). Substituted
(2014). Concordance of out-of-hospital and emergency depart- judgment: The limitations of autonomy in surrogate decision
ment cardiac arrest resuscitation with documented end-of-life making. Journal of General Internal Medicine, 23, 1514–1517.
choices in Oregon. Annals of Emergency Medicine, 63, 375– doi:10.1007/s11606-008-0688-8
383. doi:10.1016/j.annemergmed.2013.09.004 Torke, A. M., Moloney, R., Siegler, M., Abalos, A., &
Schmidt, T. A., Hickman, S. E., Tolle, S. W., & Brooks, H. S. (2004). Alexander, G. C. (2010). Physicians’ views on the impor-
The physician orders for life-sustaining treatment program: tance of patient preferences in surrogate decision-making.
Oregon emergency medical technicians’ practical experiences Journal of the American Geriatrics Society, 58, 533–538.
and attitudes. Journal of the American Geriatrics Society, 52, doi:10.1111/j.1532-5415.2010.02720.x
1430–1434. doi:10.1111/j.1532-5415.2004.52403.x Torke, A. M., Sachs, G. A., Helft, P. R., Montz, K., Hui, S. L.,
Schmidt, T. A., Zive, D., Fromme, E. K., Cook, J. N., & Tolle, S. Slaven, J. E., & Callahan, C. M. (2014). Scope and out-
W. (2014). Physician orders for life-sustaining treatment comes of surrogate decision making among hospitalized older
(POLST): Lessons learned from analysis of the Oregon adults. JAMA Internal Medicine, 174, 370–377. doi:10.1001/
POLST Registry. Resuscitation, 85, 480–485. doi:10.1016/j. jamainternmed.2013.13315
resuscitation.2013.11.027 Torke, A. M., Siegler, M., Abalos, A., Moloney, R. M., & Alexander,
Seckler, A. B., Meier, D. E., Mulvihill, M., & Paris, B. E. G. C. (2009). Physicians’ experience with surrogate decision
(1991). Substituted judgment: How accurate are proxy making for hospitalized adults. Journal of General Internal
predictions? Annals of Internal Medicine, 115, 92–98. Medicine, 24, 1023–1028. doi:10.1007/s11606-009-1065-y
doi:10.7326/0003-4819-115-2-92 Tremblay, D., Arnsten, J. H., & Southern, W. N. (2016). A simple
Shalowitz, D. I., Garrett-Mayer, E., & Wendler, D. (2006). The accu- and powerful risk-adjustment tool for 30-day mortality among
racy of surrogate decision makers: A systematic review. Archives of inpatients. Quality Management in Health Care, 25, 123–128.
Internal Medicine, 166, 493–497. doi:10.1001/archinte.166.5.493 doi:10.1097/QMH.0000000000000096
StataCorp. 2013. (2013). Stata statistical software release 13.1. Vick, J., Pertsch, N., Hutchings, M., Neville, B., & Bernacki,
College Station, TX: StataCorp LP. R. (2016). The utility of the surprise question in identify-
St-Louis, E., Iqbal, S., Feldman, L. S., Sudarshan, M., Deckelbaum, ing patients most at risk of death [Abstract]. Journal of
D. L., Razek, T. S., & Khwaja, K. (2015). Using the age-adjusted Pain and Symptom Management, 51, 342. doi:10.1016/j.
Charlson comorbidity index to predict outcomes in emergency jpainsymman.2015.12.177
general surgery. The Journal of Trauma and Acute Care Surgery, Wilkinson, D. J., & Truog, R. D. (2013). The luck of the draw:
78, 318–323. doi:10.1097/TA.0000000000000457 Physician-related variability in end-of-life decision-making
Suhl, J., Simons, P., Reedy, T., & Garrick, T. (1994). Myth of sub- in intensive care. Intensive Care Medicine, 39, 1128–1132.
stituted judgment. Surrogate decision making regarding life doi:10.1007/s00134-013-2871-6

Downloaded from https://academic.oup.com/gerontologist/advance-article-abstract/doi/10.1093/geront/gny042/4998924


by university of winnipeg user
on 24 May 2018

You might also like