You are on page 1of 27

Article

Biological Research for Nursing


1-27
Biological Processes and Biomarkers ª The Author(s) 2018
Article reuse guidelines:
sagepub.com/journals-permissions
Related to Frailty in Older Adults: DOI: 10.1177/1099800418798047
journals.sagepub.com/home/brn
A State-of-the-Science Literature Review

Jinjiao Wang, PhD, RN1, Cathy A. Maxwell, PhD, RN2,


and Fang Yu, PhD, RN, GNP-BC, FGSA, FAAN3

Abstract
The objectives of this literature review were to (1) synthesize biological processes linked to frailty and their corresponding
biomarkers and (2) identify potential associations among these processes and biomarkers. In September 2016, PubMed,
Cumulative Index to Nursing and Allied Health, Cochrane Library, and Embase were searched. Studies examining biological
processes related to frailty in older adults (≥ 60 years) were included. Studies were excluded if they did not employ specific
measures of frailty, did not report the association between biomarkers and frailty, or focused on nonelderly samples (average
age < 60). Review articles, commentaries, editorials, and non-English articles were also excluded. Fifty-two articles were
reviewed, reporting six biological processes related to frailty and multiple associated biomarkers. The processes
(biomarkers) include brain changes (neurotrophic factor, gray matter volume), endocrine dysregulation (growth hormones
[insulin-like growth factor-1 and binding proteins], hormones related to glucose and insulin, the vitamin D axis, thyroid
function, reproductive axis, and hypothalamic–pituitary–adrenal axis), enhanced inflammation (C-reactive protein,
interleukin-6), immune dysfunction (neutrophils, monocytes, neopterin, CD8þCD28T cells, albumin), metabolic imbalance
(micronutrients, metabolites, enzyme-activity indices, metabolic end products), and oxidative stress (antioxidants, telomere
length, glutathione/oxidized glutathione ratio). Bidirectional interrelationships exist within and between these processes.
Biomarkers were associated with frailty in varied strengths, and the causality remains unclear. In conclusion, frailty is
related to multisystem physiological changes. Future research should examine the dynamic interactions among these
processes to inform causality of frailty. Given the multifactorial nature of frailty, a composite index of multisystem
biomarkers would likely be more informative than single biomarkers in early detection of frailty.

Keywords
frailty, frail elderly, older adults, biomarkers, biological phenomena

Frailty, a common geriatric syndrome affecting 7 million older Kim, & Won, 2015; Collard, Boter, Schoevers, & Oude
Americans (Bandeen-Roche et al., 2015; Mather, Jacobsen, & Voshaar, 2012), and increase the risk of multiple adverse out-
Pollard, 2015), represents a state with reduced physiological comes such as disability (Ehlenbach, Larson, Randall Curtis, &
reserve for responding to stressors (El Assar et al., 2017; Fried Hough, 2015), falls and fractures (Ensrud et al., 2007), hospi-
et al., 2001; Wang, Boehm, & Mion, 2017). Frailty has been talization (Joseph et al., 2014; Kahlon et al., 2015), and death
defined and measured in a number of ways (Buta et al., 2016). (Shamliyan, Talley, Ramakrishnan, & Kane, 2013).
The most commonly used definition considers frailty to be a The World Health Organization (2015) recognized frailty as
phenotype comprising five indicators: unintentional weight an emerging public health priority and urged early recognition
loss of 5% or 10 pounds within 1 year, slow gait speed, and intervention for frailty. A first step is to understand the
reduced grip strength, exhaustion, and physical inactivity
(Fried et al., 2001). Clinically, frailty is diagnosed by the pres-
ence of three or more of these indicators and is preceded by a 1
School of Nursing, University of Rochester, Rochester, NY, USA
prodromal stage known as prefrailty, defined by having one or 2
School of Nursing, Vanderbilt University, Nashville, TN, USA
3
two indicators (Fried et al., 2001). Other common measures School of Nursing, University of Minnesota, Minneapolis, MN, USA
include the frailty index of deficits (Rockwood et al., 2005)
Corresponding Author:
and the Vulnerable Elders Survey-13 (Saliba et al., 2001). Jinjiao Wang, PhD, RN, School of Nursing, University of Rochester, Room
Frailty and prefrailty affect more than 10% and up to 50% of 2w.319, 255 Crittenden Blvd, Rochester, NY 14642, USA.
community-dwelling older adults, respectively (Choi, Ahn, Email: jinjiao_wang@urmc.rochester.edu
2 Biological Research for Nursing XX(X)

biological processes from which frailty arises and their associ- independently conducted a secondary review of titles,
ated biomarkers. Biomarkers provide objective, measurable abstracts, and article content for inclusion. After consensus was
indices for normal and/or pathogenic states as well as responses reached, J.W. reviewed the full text of eligible studies to con-
to therapeutic interventions (Colburn et al., 2001). Since altera- firm adherence to eligibility criteria.
tions in biomarkers precede clinical manifestation of frailty
(Sanchis et al., 2015), they are useful for detecting subclinical
changes that lead to observable indicators of frailty. Biomar- Data Extraction
kers are also useful for tailoring and testing intervention pro-
J.W. and C.A.M. independently extracted the following infor-
tocols for frailty, as they reflect changes in biophysiological
mation from eligible studies: geographical region, data collec-
processes that underlie and/or relate to frailty (Corwin & Fer-
tion years, study objective, design, sampling method, sample
ranti, 2016). Studies have identified biological processes
size (and power calculation), sample characteristics (age, sex,
related to frailty, such as changes in brain structure; alterations
living arrangement, chronic conditions [e.g., dementia, cancer,
in endocrinal, immune, and skeletal muscle systems; and
cardiovascular disease]), retention rate (longitudinal studies),
chronic inflammation (Clegg, Young, Iliffe, Rikkert, & Rock-
frailty measure, examined biological process and biomarkers
wood, 2013; Fougere, Vellas, Van Kan, & Cesari, 2015; Moh-
and association with frailty, data analysis, strength of associa-
ler, Fain, Wertheimer, Najafi, & Nikolich-Žugich, 2014).
tions, control of confounders, and synthesis of results. They
Researchers have often studied these processes in isolation,
assigned this information to 11 fields (Table 1). The two
without considering the interconnectivity among them (Calvani
reviewers were blinded to each other’s screening results, and
et al., 2015; Zaslavsky et al., 2013). The objectives of the
the level of agreement on data extraction was 87.9% (i.e., per-
present state-of-the-science literature review were to (1)
centage of matched fields), with discrepancy resolved by con-
synthesize the biological processes and markers linked to
sensus with a third reviewer in weekly team meetings.
frailty and (2) identify potential associations between these
processes and frailty. In this review, we use the term biomar-
Quality assessment. We assessed the quality of the reviewed
kers to refer to general laboratory-measured markers of biolo-
studies using the “Strengthening the Reporting of Observa-
gical phenomena related to frailty.
tional Studies in Epidemiology (STROBE) Statement” (von
Elm et al., 2007). First, J.W., C.A.M., and a trained research
Method assistant independently evaluated a random sample of the eli-
gible studies (n ¼ 6); interrater agreement was 90%. After we
We conducted and reported this review in three steps: data discussed disagreements and resolved them by consensus, the
acquisition, data extraction, and quality assessment of identi- research assistant evaluated the quality of the remaining
fied studies. studies.

Data Acquisition
We conducted a comprehensive literature search in September Results
2016 in four major electronic databases (PubMed, Cumulative
The initial search strategy produced 272 articles. After we
Index to Nursing and Allied Health [CINAHL], Cochrane
removed 10 duplicates and 174 unsuitable articles during the
Library, and Embase) to identify studies on biological pro-
title, abstract, and content screening, 88 data-driven articles
cesses and biomarkers related to frailty in older adults (60
remained for further review. Full-text reviews led to the exclu-
years of age). We used text terms (i.e., frail*, biomarker*,
sion of 36 studies that did not utilize a specific frailty measure
biological, mechanism) and the following Medical Subject
(n ¼ 22), did not measure specific biomarkers (n ¼ 7), focused
Headings and CINAHL headings: Frailty, Frail Elderly, Older
on nonelderly samples (n ¼ 6), or did not report the association
Adults, Biomarkers, and Biological Phenomena. We reviewed
between biomarkers and frailty (n ¼ 1). In the end, 52 studies
references of extracted studies and relevant reviews to identify
met the eligibility criteria and were included in the review
additional studies and conducted a manual search of relevant
(Figure 1).
studies. We used no limitations on the publication years in the
search process. In the final selection of articles, we included
primary studies examining the biological processes and bio-
markers related to frailty and excluded studies that lacked a
Study Characteristics
specific definition or measure of frailty, those not reporting Years and region. The majority of the studies (n ¼ 33, 63%) were
associations between biomarkers and frailty, and those focused published after 2010; 23% (n ¼ 12) did not specify data col-
on nonelderly samples (average age < 60 years). We also lection years, 56% (n ¼ 29) utilized data collected before 2007.
excluded non-data-driven articles (reviews, commentaries, and The studies were conducted primarily in North America
editorials) and articles in languages other than English. Author (United States: n ¼ 25, 48%) and Europe (n ¼ 18, 35%), with
J.W. initially screened the titles and abstracts of identified six in Asia (Australia, China, and Japan; 12%) and one in South
studies for preliminary eligibility. Author C.A.M. America (Brazil, 2%).
Table 1. Description of Reviewed Studies.
Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Walston AM: United States Cross-sectional Population-based 4,735; community; average Parkinson’s disease, stroke, Dementia with 1. Inflammation: CRP (positive) Phenotype; outcome
et al. (1989–1990) sampling age: 72.7 years; biological depression under treatment MMSE < 18 2. Coagulation: Factor VIII, D
(2002) (CHS) sex: both; race: 95% dimer (positive)
Caucasian 3. Endocrine: glucose and insulin
(positive)
4. Immune: albumin (negative)
Leng et al. AM: United States Cross-sectional Convenience 326; community; average Parkinson’s disease, cerebral Cognitive deficit 1. Endocrine: serum IGF-I, Phenotype;
(2004)a (unspecified) sampling age:81.3–84.9 years; vascular accident with with MMSE < DHEA-S (negative) exposure
biological sex: both; residual hemiparesis, 18 2. Inflammatino: IL-6 (NS)
face: 100% Caucasian symptomatic rheumatoid
arthritis or any other
inflammatory conditions,
symptomatic congestive
heart failure or coronary
artery disease, malignancy,
current use of steroids, or
other immune-modulating
agents, chronic inflammatory
diseases or disability
conditions
Puts et al. EU: the Netherlands Longitudinal (3 Population- 885; community; average Baseline frailty N 1. Inflammation: CRP (positive) Researcher-
(2005) (1995–1996) years; based, age: 74.5 years; biological 2. Endocrine: 25(OH)D developed frailty
retention stratified sex: both; race: not (negative) indicators;
82%) random (from reported outcome
the LASA)
Semba et al. AM: United States Case–control Population-based 766; community; average 2/3 of least disabled older MMSE < 18 Immune: T cell subsets (CD8þ Phenotype; exposure
(2005)a (WHAS-I: 1993– (died vs. (WHAS I & II) age: 76.9–77.3 years (for women in the community CD28 lymphocyte counts:
1997; WHAS-II: surviving in a WHAS I and II, were excluded positive)
1994–1998) longitudinal respectively); biological
cohort study) sex: female; race: 69-77%
Caucasian
Semba et al. AM: United States Longitudinal (3 Population-based 766; community; 2/3 of least disabled older N 1. Endocrine: 25(OH)D Phenotype; outcome
(2006) (1992–1995) years; (WHAS I & II) average age: 76– 80.4 women in the community (negative)
retention years; biological sex: were excluded 2. Metabolic: serum
90%) female; carotenoids, a-tocopherol,
race: 70.4–72.7% nutritional deficiencies
Caucasian (independent of weight loss;
all negative)
Ble et al. EU: Italy (1998–2000) Cross-sectional Population- 827; community; average Cancer, disabilities in basic Dementia and Oxidative stress: plasma vitamin Phenotype; outcome
(2006) based(from age: 73.6 years; biological activities of daily living severe E (negative)
the sex: both; race: not cognitive
InCHIANTI reported impairment
study) with MMSE
< 18

(continued)

3
4
Table 1. (continued)

Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Barzilay et al. AM: United States(1989/ Longitudinal (5– Population-based 3,141; community; average Having cancer under active MMSE < 18 1. Metabolic: metabolic Phenotype; outcome
(2007) 1990–1998/1999) 9 years, (from CHS age: 70.4–74.7 years; treatment, frailty, DM, syndrome total status
retention cohort) biological sex: both; race: coronary heart disease, (positive), HOMA-IR
90%) 94–97.4% Caucasian myocardial infarction, angina, (positive)
congestive heart failure, 2. Inflammation: CRP (positive),
stroke, transient ischemic Factor VIIc (positive)
attack, peripheral arterial 3. Endocrine: glucose (positive)
disease requiring
intervention, antidepressant
use, Parkinson’s disease at
baseline
Leng et al. AM: United States Cross-sectional Population-based 1,106; community; average Having disabilities in more than N 1. Inflammation: IL-6 (positive); Phenotype; outcome
(2007) (WHAS I: 1991–1994; (from the age: 67.3 years; biological one domain, acute bacterial 2. Immune: WBC count
WHAS II: 1994) WHAS I & II sex: female; race: 79.4% infection, hematological (positive)
cohorts) Caucasian cancer
Varadhan AM: United States Cross-sectional Population-based 214; community; average Having disabilities in more than N Endocrine: higher levels and Phenotype; outcome
et al. (2004–2005) (from WHAS age: 83.4–85.5 years; one domain blunted diurnal variation of
(2008) II) biological sex: female; cortisol (positive)
race: 76.1–90% Caucasian
Matteini AM: United States Cross-sectional Population-based 703; community; average N MMSE < 24 Metabolic & oxidative stress Phenotype; outcome
et al. (unspecified) (from WHAS I age:74.2–74.9 years; 1. Methylmalonic acid
(2008) & II) biological sex: female; (indicating vitamin B12
race: 79% Caucasian deficiency; positive)
2. Vitamin B12 (negative)
Hubbard EU: United Kingdom Cross-sectional Convenience 110; institution (continuing N N 1. Inflammation: TNF-a, IL-6, Phenotype; outcome
et al. (unspecified) care, day hospital, and CRP (all positive), albumin
(2008)a independent living); (negative)
average age: 82.7–84.9 2. Metabolic: drug metabolism
years; biological sex: both; (esterase activity; negative)
race: 100% Caucasian
Woo et al. Asia: China (Hong Kong; Cross-sectional Convenience 2,000; community; average N N Oxidative stress: telomere Index; outcome
(2008) unspecified) (a 4-year age: 72.02–72.75 years; length (NS)
longitudinal biological sex: both; race:
study focused all Chinese
on mortality)

(continued)
Table 1. (continued)

Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Gruenewald AM: United States Longitudinal (3 Population-based 803; community; average Manifest functional disability at Manifest Comprehensive: allostatic load Phenotype; outcome
et al. (1990–1993) years; age: 74.23 years; baseline cognitive index (positive), based on 13
(2009) retention biological sex: both; race: disability at biomarkers:
61%) 81.85% Caucasian baseline 1. Cardiovascular: systolic and
diastolic blood pressure,
HDL, total cholesterol/HDL
2. Metabolic: HbA1c, waist–hip
ratio
3. Hormonal: DHEA, urinary
cortisol, urinary epinephrine,
urinary norepinephrine
4. Inflammation: CRP, IL-6 &
fibrinogen (coagulation)
Hubbard EU: United Kingdom Cross-sectional Convenience 110; community;average age: N N Inflammation: TNF-a, IL-6, CRP Phenotype & index;
et al. (unspecified) 83.9 years; biological sex: (positive); albumin (negative) outcome
(2009) both; race: 100%
Caucasian
Leng et al., AM: United States Cross-sectional population-based 696; community; average Acute bacterial infection or N 1. Immune: WBC (positive) Phenotype; outcome
Hung (WHAS I: 1991–1994) (from WHS I) age: 77.6 years; biological hematologic malignancies 2. Endocrine: IGF-1 (negative)
et al. sex: female; race: 68.3–
(2009) 81.1% Caucasian
Leng et al. AM: United States Cross-sectional Population-based 1,106; community; average Acute bacterial infection or N 1. Inflammation: IL-6 (positive) Phenotype; outcome
(2009) (WHAS I: 1991–1994 (from WHAS age: 77.4 years; biological hematological cancer 2. Immune: neutrophil,
and Xue & 1994) I) sex: female; race: 72.6% monocyte count (positive
et al. Caucasian but NS)
(2009)
Reiner et al. AM: United States Case–control Population-based 900; community; minimal Having frailty at baseline, N 1. Inflammation: CRP (NS), IL-6 Phenotype; outcome
(2009) (1993–1998) (frail vs. (from the age: > 65 years (average Parkinson’s disease, severe (NS)
nonfrail) Women’s age not reported); autoimmune disease, 2. Coagulation: D-dimer
Health biological sex: female; multiple sclerosis, (positive), t-PA levels
Initiative race: 96.1% Caucasian amyotrophic lateral sclerosis, (positive), factor VIII (NS),
study) congestive heart failure, fibrinogen (NS)
coronary heart disease,
stroke, cancer, or use of
antidepressant medications
at baseline
Serviddioet EU: Italy (2007) Cross-sectional Convenience 62; community; average age: Physical impairment or disability Dementia with 1. Inflammation: TNF-a Phenotype; outcome
al. (2009) 76.7 years; biological sex: evaluated by MMSE < 18 (positive)
both; race: not reported Multidimensional Geriatric 2. Oxidative stress: GSH/GSSG
Assessment, disabilities in (positive), MDA (positive),
basic activity of daily living, HNE protein plasma adducts
cancer, taking antioxidant (positive)
supplementation

(continued)

5
6
Table 1. (continued)

Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Szanton et al. AM: United States Cross-sectional Population-based 728; community; average N MMSE < 24 in Comprehensive: allostatic load Phenotype; outcome
(2009) (WHAS I &II: 1991– (from WHAS I age: 74.2 years; biological one sample, (positive)
1994 & 1994) & II) sex: female; race:76% MMSE < 18 in
Caucasian another
sample
Wu et al. Asia: mainland Cross-sectional Convenience 90; community; average age: Malignancy, Parkinson’s disease, N Oxidative stress: 8-OHdG Phenotype; outcome
(2009) China(2008) 73.1–79.9 years; or stroke, not being able to (positive)
biological sex: both; race: walk
all Chinese
Hyde et al. Asia: Australia (2001– Longitudinal (7 Population-based 3,616; community: average Prostate cancer or previous N Endocrine: free testosterone FRAIL questionnaire;
(2010) 2004, 2008–2009) years; (from the age: 76.9 years; biological orchidectomy and those (negative: incident and outcome
retention Health in Men sex: male; race: receiving GnRH analogs, prevalent frailty)
44%) study) predominantly Caucasian antiandrogen therapy, or
(proportion not testosterone
reported) supplementation
Ronning EU: Norway Cross-sectional Convenience 187; cancer patients Not undergoing elective N 1. Inflammation: CRP, IL-6, Phenotype & Balducci
et al. (unspecified) (colorectal); average age: resections of tumors in colon TNF-a (positive) CGA category;
(2010)a 80 years; biological sex: or rectum 2. Coagulation: D-dimer exposure
both; race: not reported (positive)
Ensrud et al. AM: United Longitudinal (4.6 Population-based 1,128; community; average History of hip replacement, N Endocrine: 25(OH)D (negative) Phenotype; outcome
(2011) States(2000–2002) years; (from the age: 73.8 years; biological unable to walk without
retention CHS) sex: male; race: 89.7% assistance from another
76%) Caucasian person
Leng et al. AM: United States Cross-sectional Convenience 133; community; average Parkinson’s disease, stroke with Significant 1. Inflammation: IL-6 (positive) Phenotype; outcome
(2011) (unspecified) age: 84 years; biological residual hemiparesis, cognitive 2. Immune: neopterin (positive)
sex: both; race: 95.7% symptomatic congestive deficit with
Caucasian heart failure, uncompensated MMSE < 18
endocrine disorders, active
malignancy, rheumatoid
arthritis or any other
inflammatory conditions, or
use of immune-modulating
drugs including oral
corticosteroids
Tajar et al. EU: United Kingdom Cross-sectional Population-based 1,504; community; average Prevalent pituitary, testicular, N Endocrine: testosterone Phenotype & index;
(2011) (2003–2005) (the European age: 60 years; biological or adrenal diseases or use of (negative), LH (positive), FSH outcome
Male Aging sex: male; race: medications that are likely to (positive), DHEA-S
Study) predominantly Caucasian affect HPT function (negative), SHBG (negative)
(proportion not (anabolic-androgenic
reported) steroids, DHEA, anti-
androgens, GnRH agonists,
and psycholeptic agents) or
clearance of sex steroids
(e.g., anticonvulsants)

(continued)
Table 1. (continued)

Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Coelho et al. AM: Brazil Cross-sectional Convenience 48; community; Acute inflammatory condition Cognitive Brain: BDNF (negative) Phenotype; exposure
(2012)a (unspecified) (in a quasi- average age: 70.5–72.5 or active neoplasm in the alterations
experimental years; biological sex: previous 5 years; use of detected by
study) female; race: not medication with broad-based MMSE
reported action on the immune
system; amputation or
fracture of the upper or
lower limbs in the previous 6
months; presence of any
neurological diseases
Collerton EU: United Kingdom Cross-sectional Population-based 845; institution; age range: Stroke, Parkinson’s disease MMSE < 18, or 1. Inflammation: IL-6, TNF-a, Phenotype & index;
et al. (2006-2007) (from the 85þ years (average age taking drugs CRP (positive) outcome
(2012) Newcastle not reported); biological for dementia 2. Immune: neutrophils
85þ study) sex: both; race: 99.6% (positive), total lymphocyte
Caucasian count (albumin, lymphocytes,
memory/naive CD8 T cell
ratio; all negative)
Kalyani, Tian AM: United States Longitudinal (9 Population-based 329; community: average Stroke or Parkinson’s disease, N Endocrine: HbA1c (positive) Phenotype; outcome
et al. (1994–2008) years; (from WHAS age: 73.9 years; biological HbA1c < 4.5%
(2012) retention II cohort) sex: female; race: 83.9%
90%) Caucasian
Kalyani et al. AM: United States Cross-sectional Population-based 73; community; average age: N N Endocrine: glucose and insulin Phenotype; exposure
(2012a) (2008–2009) (from WHAS 86.7 years; biological sex: levels in oral glucose
II cohort) female; race: 84.9% tolerance test (positive)
Caucasian
Kalyani et al. AM: United States Cross-sectional Population-based 73; community; average age: Known diabetes or taking N 1. Metabolic: FFa (positive) Phenotype; exposure
(2012b) (2008–2009) (from WHAS 86–87 years; biological corticosteroids, having 2. Endocrine: gut- (ghrelin
II cohort) sex: female; race: not difficulty in more than 1 of 4 [negative], GLP-1 [negative])
reported domains of physical function and adipocyte-derived
hormones (leptin [positive],
adiponectin [negative],
resistin [positive]), growth
hormone (GH [positive]),
IGF-1 [negative]
3. Inflammation: IL-6 (positive)
Yeap et al. Asia: Australia (2001– Cross-sectional Population-based 3,943; community: average History of thyroid disease, N Endocrine: TSH (NS), FT4 FRAIL questionnaire;
(2012) 2004) (from the age: 75.15–76.19 years; thyroid surgery or treatment (positive) outcome
Health in Men biological sex: male; race: with radioactive iodine,
study) predominantly Caucasian previously undiagnosed
(proportion not hyperthyroidism,
reported) undiagnosed hypothyroidism

(continued)

7
8
Table 1. (continued)

Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Baylis et al. EU: United Kingdom Longitudinal (10 Population-based 254; community; average N N 1. Immune: higher baseline Phenotype; outcome
(2013) (1994–2003) years; (from the age: 66.9–67.3 years; levels of differential WBC
retention Hertfordshire biological sex: both; race: 2. Endocrine: DHEA-S
35%) Ageing Study) not reported (negative), cortisol/DHEA-S
(positive)
Dalrymple AM: United States Longitudinal Population-based 4,150; community; average Institutionalized, using a Alzheimer’s Kidney function: eGFR using Phenotype; outcome
et al. (1989–1990 & 1992– (follow-up of age: 75 years; biological wheelchair within the home, disease, serum cystatin C (negative)
(2013) 1993) 4 years, sex: both; race: 83% non- receiving chemotherapy or MMSE < 18,
retention Black radiation for cancer, or modified
unspecified) receiving hospice care, MMSE < 60
Parkinson’s disease, stroke,
prescribed medications for
Parkinson’s disease,
depression, baseline frailty
Gale et al. EU: United Kingdom Longitudinal (6 Population-based 2,146; community; average N N 1. Inflammation: CRP (positive) Phenotype; outcome
(2013) (2002–2008) years, (from the age: 70.22 years; 2. Coagulation: fibrinogen
retention English biological sex: both; race: (positive: only in women, not
46%) Longitudinal not reported in men)
Study of
Ageing)
Hart et al. AM: United States Cross-sectional Convenience (six 1,602; community; average N N Kidney function: cystatin C Phenotype; outcome
(2013) (2000–2002) clinical sites) age: 73.8 years; biological (positive), but not creatinine-
sex: male; race: 91.5% based measures
Caucasian
Tajar et al. EU: Six European Cross-sectional Stratified random 1,504; community; average N N Endocrine: 25(OH)D (negative); Phenotype & index;
(2013) countries(unspecified) (on center age: 69.5 years; biological PTH (positive) outcome
level) sex: male; race: 70–90%
Caucasian
Yeap et al. Asia: Australia (2001– Longitudinal (7 Population-based 3,447; community: average On testosterone or hormonal N Endocrine: IGF-1 (negative) and FRAIL questionnaire;
(2013) 2004, 2008–2009) years; (from the age: 75.8–78.3 years; therapies, history of prostate binding proteins (positive), outcome
retention Health In Men biological sex: male; race: cancer, metabolic syndrome total and free testosterone
48%) study) predominantly Caucasian (negative)
(proportion not
reported)
Addison AM: United States Case–control Convenience 26; community; average age: N N 1. Inflammation: IL-6 (positive), MPPT < 25
et al. (unspecified) (frail vs. 78.1–83.3 years; TNF-a (NS) (moderately frail)
(2014)a nonfrail) biological sex: both; race: 2. Metabolic (muscle): IMAT and reported little
not reported (positive) to no
planned physical
activity over the
last year; outcome

(continued)
Table 1. (continued)

Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Corona et al. EU: Italy (2009–2010) Cross-sectional Convenience 89; cancer patients (breast); N N Metabolic: amino acids CGA (Balducci);
(2014)a average age: 77 years; (positive), acylcarnitines outcome
biological sex: female; (positive), sphingo- and
race: not reported glycerol-phospholipids
(negative)
Inglés et al. EU: Spain (2006–2009 Cross-sectional Population-based 742; community; average N N Oxidative stress: MDA Phenotype; outcome
(2014)a 2013–2014) (from the age: 76.46 years; (positive), protein
Toledo Study biological sex: both; race: carbonylation (positive).
for Healthy not reported Both are circulating
Aging) oxidative-damage
biomarkers
Lai et al. China (Taiwan; (2007) Cross-sectional Convenience 386; institution (veterans Acute medical problems N Inflammation: IL-6 (positive) Phenotype; outcome
(2014) care home); average age:
81.5 years; biological sex:
male; race: all Chinese
Whitson AM: United States Cross-sectional Population-based 3,373; community; average Wheelchair bound or under N Metabolic: CML (positive) Phenotype; outcome
et al. (1996–1997) data about (from the CHS age: 78.1 years; biological active treatment for cancer
(2014) frailty in a 14- study) sex: both; race: 83.5–
year 84.5% Caucasian
longitudinal
cohort
Brouwerset EU: Belgium Cross-sectional Convenience 244; cancer patients Having new diagnosis of early or N Inflammation: IL-6 (positive) LOFS & Balducci’s
al. (2015)a (unspecified) (metastatic breast); locally advanced (i.e. CGA; exposure
average age: 76 years; nonmetastatic), primary or
biological sex: both; race: second primary breast
not reported cancer before initiation of
any chemotherapy,
radiotherapy or surgery
W. T. Chen China (Taiwan; Cross-sectional Population-based 456; community; average Major illness with limited life Dementia Brain: GMV (negative), CSF Phenotype; exposure
et al. 2011–2012) (from the I- age: 64 years; biological expectancy (less than 6 (positive)
(2015)a Lan sex: both; race: all months); having any
Longitudinal Chinese contraindication for MRI
Aging Study) such as metal implants;
having been institutionalized
for any reason; major
neuropsychiatric diseases
such as stroke, brain tumor,
or major depression
Sanchis et al. EU: Spain (2010–2012) Cross-sectional Convenience 342; hospitalized patients Prior diagnosis of heart disease N 1. Endocrine: vitamin D Phenotype; outcome
(2015) (acute coronary other than ischemic heart (negative)
syndrome survivors); disease and the indication of 2. Metabolic: hemoglobin
average age: 77–81 years; coronary surgery during (negative)
biological sex: both; race: hospitalization 3. Kidney function: cystatin-C
not reported (positive)

(continued)

9
10
Table 1. (continued)

Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Urpi-Sarda EU: Italy (1998–2000) Cross-sectional Population-based 811; community; average N N Metabolic: DTP intake and its Phenotype; outcome
et al. (from the age: 74.3 years; biological biomarker (NS), UTP
(2015) InCHIANTI sex: both; race: not (negative)
study) reported
Beben et al. AM: United States Cross-sectional Population-based 2,977; community; average In hospice or receiving radiation N Endocrine: FGF-23 (positive) Phenotype; outcome
(2016) (1996–1997) (from the CHS age: 77.9 years; biological or chemotherapy for cancer
study) sex: both; race: 83%
Caucasian
Liu et al. AM: United States Cross-sectional Population-based 1,919; community; average N N 1. Inflammation: IL-6 (positive) Phenotype; outcome
(2016) (2005–2008) (from the age: 70.94 years; 2. Oxidative stress:
Framingham biological sex: both; race: isoprostanes (positive),
Offspring not reported LpPLA2 mass (positive)
Study)
Lorenzi et al. EU: Spain (2012–2013) Cross-sectional Convenience 120; community; average Estimated life expectancy less MMSE < 18 Inflammation: plasma HtrA1 Phenotype & index;
(2016)a age: 75.4 years; biological than 6 months (positive) exposure
sex: both; race: not
reported
Namioka Asia: Japan (unspecified) Cross-sectional Convenience 140; patients with Severe disability, major Moderate-to- 1. Oxidative stress: plasma Phenotype; exposure
et al. Alzheimer’s disease (mild depression, malignancies, severe diacron reactive oxygen
(2016)a to moderate); average severe cardiac or pulmonary dementia metabolite (positive),
age: 78.2–82.3 years; disease, liver cirrsis, severe with MMSE < biological
biological sex: both; race: chronic kidney disease 18 antioxidant potential
all Japanese (negative), bilirubin
(negative), endogenous
plasma antioxidants urinary
8-OHdG (positive),
epiPGF2a (8-isoprostane;
negative), plasma albumin
(NS), uric acid (NS)
2. Inflammation: IL-6 (NS), TNF-
a (NS)
Zaslavsky AM: United States Longitudinal (4.8 Population-based 1,848; community; average Baseline frailty Dementia at Endocrine: glucose and HbA1c Phenotype; outcome
et al., (1994–1996, 2000– years, (from the age: 76 years; biological baseline or (positive)
2016 2003, 2005) retention Adult Changes sex: both; race: 85.9– during
71%) in Thought 92.8% Caucasian follow-up
study)

(continued)
Table 1. (continued)

Study Design
Article: first (Longitudinal: Sample Characteristics: Size; Frailty: Measure; As
Author, Pub- Region* (Data- Follow-Up; Setting; Age; Sex; Race/ Exclusion Based on Health Exclusion Based Biomarkers (Association With Exposure or
lication Year Collection Year[s]) Retention) Sampling Method Ethnicity Conditions on Cognition Frailty) Outcome

Zhu et al., Asia: mainland China Cross-sectional Stratified random 1,478; community; average Self-reported Parkinson’s Severe cognitive Inflammation: hsCRP (positive) Phenotype; outcome
2016 (2014) age: 75.3 years; biological disease, Alzheimer’s disease, impairment
sex: both; race: all stroke, depression, treated with revised
Chinese in emergency department Hasegawa’s
during the last 3 months or Dementia
taking anti-inflammatory Scale score <
drugs (nonsteroidal anti- 10
inflammatory drugs)
continuously for 1 month

Note. 25(OH)D ¼ 25 hydroxyvitamin D; 8-OHdG ¼ 8-hydroxy-20-deoxyguanosine; AM ¼ America; BDNF ¼ brain-derived neurotrophic factor; CGA ¼ Comprehensive Geriatric Assessment; CHS ¼ Cardiovascular
Health Study; CML ¼ carboxymethyl-lysine; CRP ¼ C-reactive protein; CSF ¼ cerebrospinal fluid; DHEA-S ¼ dehydroepiandrosterone-sulphate; DM ¼ diabetes mellitus; DTP ¼ dietary total polyphenols; eGFR ¼
estimated glomerular filtration rate; EU: Europe; F/U: follow-up; FFa ¼ free fatty acids; FGF-23 ¼ fibroblast growth factor-23; FSH ¼ follicle-stimulating hormone; FT4 ¼ free thyroxine; GlP-1 ¼ glucagon-like peptide-1;
GMV ¼ gray-matter volume; GSH ¼ glutathione; GnRH ¼ gonadotropin-releasing hormone; GSSG ¼ oxidized glutathione; HbA1c ¼ glycosylated hemoglobin; HDL ¼ high-density lipoprotein cholesterol; HNE ¼ 4-
hydroxy-2,3-nonenal; HOMA-IR ¼ homeostasis model-insulin resistance; HPT ¼ hypothalamic–pituitary–testicular; hsCRP ¼ high-sensitivity C-reactive protein; HtrA1 ¼ high-temperature requirement serine protease
A1; IGF-1 ¼ insulin-like growth factor-1; IL-6 ¼ interleukin 6; IMAT ¼ intramuscular adipose tissue; LASA ¼ Longitudinal Aging Study Amsterdam; LH ¼ luteinizing hormone; LOFS ¼ Leuven Oncogeriatric Frailty Score;
LpPLA2 ¼ lipoprotein phospholipase A2; MDA ¼ malonaldehyde; MMSE ¼ Mini-Mental State Examination; MPPT ¼ modified physical performance test score; MRI ¼ magnetic resonance imaging; NS ¼ not significant; PTH
¼ parathyroid hormones; SHBG ¼ sex hormone–binding globulin; t-PA ¼ tissue plasminogen activator; TNF-a ¼ tumor necrosis factor-alpha; TSH ¼ circulating thyrotropin (thyroid-stimulating hormone); UTP ¼ urinary
total polyphenols; WBC ¼ white blood cell; WHAS ¼ Women’s Health and Aging Study.
a
These 12 studies did not report the specific strength of identified associations between biomarkers and frailty and are, thus, not included in Table 1.

11
12 Biological Research for Nursing XX(X)

Records idenfied (n = 272)


i. database search (n = 241)

Idenficaon
ii. reference search (n = 26)
iii. -manual search (n = 5)

Duplicates excluded (n = 10)

Records aer duplicates removed


(n = 262) Records excluded (n =174)
i. topics unrelated to frailty, but
other issues in the frail
Screening

populaon (n = 114)
ii. not data driven: editorials,
Records screened for tle and leers to editors, opinions and
abstract (n = 88) discussion pieces, theorecal
papers, reviews (n = 55)
iii. no full text: conference
abstracts (n = 1)
iv. not published in English (n = 4)
Eligibility

Full-text arcles assessed for


eligibility (n =88) Records excluded (n = 36)
i. not focused on frailty or did
not measure frailty (n = 22)
ii. did not measure biomarkers (n
= 7)
iii. non-elderly sample (average
age < 60 years; n = 6)
iv. did not report the associaon
between biomarkers and frailty
Included

Studies included (N = 52) (n = 1)

Figure 1. Literature search flowchart (Preferred Reporting Items for Systematic Reviews and Meta-Analyses [PRISMA] format; Moher, Liberati,
Tetzlaff, & Altman, 2009).

Study design and data type. The majority of the studies (n ¼ 37, 58%), while 22 were restricted to one sex (male: n ¼ 8, 15%;
71%) were cross-sectional, 12 (23%) were longitudinal, and female: n ¼ 14, 27%). Among the studies (n ¼ 30, 58%) that
three (6%) were case–control. Only one study was interven- reported race and/or ethnicity of the sample, 44% (n ¼ 23)
tional (Coelho et al., 2012), and the rest were observational. comprised primarily Caucasian populations (proportion: 80–
Follow-up periods in the 12 longitudinal studies ranged from 3 100%). In 30 (58%) studies, the average age was 75 years.
to 14 years. Over, half (n ¼ 33, 52%) of the studies were The study populations were primarily convenience samples (n
secondary analyses of large-scale population cohorts, particu- ¼ 17, 33%) or population-based cohorts (n ¼ 33, 63%), though
larly the Women’s Health and Aging Study (n ¼ 12, 23%) and in two studies, researchers independently collected random
the Cardiovascular Health Study (n ¼ 6, 12%) in the United samples (6%). Sample size ranged from 26 to 4,735, with 45
States. (87%) studies having a sample size 100. Only two (4%)
studies reported power calculations. Retention rate in longitu-
Population and sample. Target populations were primarily dinal studies (n ¼ 12) ranged from 35% to 90%. Exclusion
community-dwelling older adults (n ¼ 44, 85%), followed by criteria in 37 (71%) studies included various health conditions
nursing home residents (n ¼ 3, 6%) and older adults with (e.g., cardiovascular disease, cancer, stroke, and Parkinson’s
specific health conditions including cancer (n ¼ 3, 6%), acute disease; see Table 1), while those in 18 (35%) included
cardiac conditions (n ¼ 1, 2%), and Alzheimer’s disease (n ¼ 1, impaired cognitive functioning identified by a diagnosis of
2%). Most studies included older adults of both sexes (n ¼ 30, dementia or cognitive screening tests (e.g., Mini-Mental State
Wang et al. 13

Exam [MMSE] score < 18 or 24). However, one study was that the same process was named differently in different stud-
conducted with patients with mild cognitive impairment in ies. For example, some studies grouped inflammation into
Alzheimer’s disease (MMSE > 18; Namioka et al., 2016), and immune processes while others did not.
another specifically stated that many subjects had cognitive
impairment (Hubbard, O’Mahony, Savva, Calver, & Wood- Changes in the brain. Aging- and disease-related changes in the
house, 2009). brain contribute to the development of frailty, including struc-
tural, physiological, and functional brain changes and
Measure of frailty, biomarkers, and analyses. The majority of stud- decreased brain reserve in the stress response (Buchman
ies (n ¼ 44, 85%) measured frailty using the Fried phenotype et al., 2014; Buchman, Yu, Wilson, Schneider, & Bennett,
(Fried et al., 2001) alone or with other measures, such as the 2013; Clegg et al., 2013; D. B. Miller & O’Callaghan, 2005).
frailty index of deficits (Rockwood et al., 2005; Rockwood & When these changes occur separately or in concert and accu-
Mitnitski, 2007), the FRAIL questionnaire (Morley, Mal- mulate to an extent that causes cognitive decline, frailty may
mstrom & Miller, 2012), and Balducci’s Comprehensive Ger- ensue. For example, neuronal loss due to Alzheimer’s pathol-
iatric Assessment (CGA; Balducci & Extermann, 2000). ogy (amyloid plaques, neurofibrillary tangles) and infarcts
Among studies not using the Fried phenotype as the frailty (Buchman et al., 2013, 2014) in areas of the cerebellum and
measure (n ¼ 8, 15%), three (6%) used the FRAIL question- hippocampus can cause cognitive decline (i.e., in motor plan-
naire (Hyde et al., 2010; Yeap et al., 2012, 2013), two (4%) ning) that is related to the frailty symptom of slow gait speed
used Balducci’s CGA (Brouwers et al., 2015; Corona et al., (W.T. Chen et al., 2015). As the brain changes accumulate and
2014), one (2%) used the frailty index of deficits (Woo et al., cognitive decline worsens, frailty progresses through stages of
2008), one (2%) used a modified physical performance mea- prefrailty, frailty, and eventually disability (Robertson, Savva,
sure (Addison et al., 2014), and one (2%) used a self-developed & Kenny, 2013). In addition, neurons in certain brain areas,
tool (Puts, Visser, Twisk, Deeg, & Lips, 2005). such as the hippocampal pyramid, are key to maintaining a
Biomarkers were measured as continuous variables in 28 balanced stress response (Clegg et al., 2013; Orlovsky,
(54%) studies, as categorical variables in 16 (31%) studies, and Dosenko, Spiga, Skibo, & Lightman, 2014). Thus, losing these
as both in 8 (15%) studies. Among the 24 studies that categor- neurons secondary to aging decreases the brain’s reserve in the
ized the level of biomarkers, different cut points were used to stress response, which may be at the core of frailty develop-
determine high-risk biomarker levels (e.g., in binary, tertile, ment (El Assar et al., 2017).
quartile, or quintile groups with different ranges for each). Two studies identified two brain biomarkers for frailty: gray
Linear or logistic regression and Cox proportional models were matter volume and brain-derived neurotrophic factor (BDNF;
the most common analyses (n ¼ 40, 77%) to examine the W.T.Chen et al., 2015; Coelho et al., 2012). Gray matter vol-
association between biomarkers and frailty, followed by mean ume can be measured by magnetic resonance imaging, and
comparisons across groups (e.g., analysis of variance; n ¼ 9, research has linked its decrease to slow gait speed, physical
17%) and bivariate correlations (n ¼ 3, 6%). In analysis, 42 inactivity, and reduced handgrip; all of which are clinical indi-
(81%) studies treated frailty as the outcome, the rest (n ¼ 10, cators of frailty (W.T.Chen et al., 2015). BDNF protects adult
23%) treated frailty as the exposure (e.g., comparing the means neurons from death during stress (Schäbitz et al., 2007) and
[n ¼ 9] or images [n ¼ 1] of biomarkers among frail, prefrail, promotes the development of immature neurons (Lindsay,
and nonfrail groups). Forty studies (77%) reported specific 1994). A lower level of BDNF indicates that the brain has
strength of the association between examined biomarkers and reduced reserves to respond to stressors. Due to the lack of
frailty in the forms of odds ratio, hazard ratio, correlation coef- direct measures of cerebral BDNF, researchers have used
ficient, or regression b coefficient, while 12 (23%) studies did plasma BDNF concentrations, measured using centrifuged
not report such parameters. Confounders such as demographic blood samples with sandwich enzyme-linked immunosorbent
status, body mass index, and comorbidities were controlled in assay (ELISA) kits, as a proxy measure (Inglés et al., 2017;
87% (n ¼ 45) of the studies (Table 1). Noble, Billington, Kotz, & Wang, 2011). Lower plasma BDMF
concentration is related to higher rates of frailty (Coelho et al.,
2012). Such evidence indicates the importance of the central
Biological Processes Related to Frailty and Associated
nervous system to the maintenance of stress adaptability and
Biomarkers reveals a role for the link between cognitive functioning (i.e.,
We identified six biological processes related to frailty in the stress adaption in the central nervous system) and physical
reviewed studies, including (1) brain changes, (2) endocrine functioning (i.e., physiological reserve) in the development
dysregulation, (3) enhanced inflammation, (4) immune dys- of frailty.
function, (5) metabolic imbalance, and (6) oxidative stress. Out
of all the studies, 21 (40%) examined more than one process, Endocrine dysregulation. Endocrine dysregulation entails aging-
primarily enhanced inflammation along with immune dysfunc- related disturbances of homeostasis, a balance between ade-
tion or endocrine dysregulation (Table 1). We categorized the quate protection from invading pathogens and harmful effects
biomarkers into these six processes based on how the research- of continual low-grade inflammation (Hunt, Walsh, Voegeli, &
ers had categorized them; however, it is not uncommon to see Roberts, 2010). Through the processes of apoptosis,
14 Biological Research for Nursing XX(X)

senescence, and inflammation, hormones help maintain home- cortisol/DHEA-S ratio were related to higher rates of
ostasis (Fedarko, 2011). When hormones are dysregulated, frailty (Baylis et al., 2013; Leng et al., 2004; Varadhan
internal homeostasis is disturbed, thus reducing adaptability et al., 2008). When faced with stressors, a fundamental
to stressors and increasing the risk of frailty. Hormonal dysre- response of the human body is to activate the HPA axis.
gulation is associated with poor skeletal muscle protein synth- Enhanced reaction of the HPA axis, such as that which
esis, which leads to weakness, a frailty indicator (Cooper et al., occurs during chronic, persistent stress, leading to
2012). In the present review, 20 studies examined the relation- higher levels of cortisol, epinephrine, and norepinephr-
ship between endocrine dysregulation and frailty (Table 1), ine, is linked to the etiology of several age-related
identifying the following biomarkers associated with one or health conditions that are related to frailty, such as car-
more indicators in the frailty phenotype: diovascular disease, diabetes, and osteoporosis (Ros-
mond & Björntorp, 2000; Tsigos & Chrousos, 2002).
1. Growth hormones (assessed by serum concentrations). This may be why hormones of the HPA axis, especially
Researchers found that older adults with lower levels of during abnormal stress-response patterns, are related to
insulin-like growth factor-1 (IGF-1) had higher preva- an overall status of frailty and higher incident rates of
lence rates of frailty (Leng et al., 2004; Leng, Hung, health conditions such as cardiovascular disease (See-
et al., 2009) and higher risk of incident frailty (Yeap man, Singer, Rowe, Horwitz, & McEwen, 1997). How-
et al., 2013). In contrast, higher levels of IGF-1 binding ever, no associations have been reported between these
proteins, such as IGFBP-1, were related to higher hormones and specific frailty indicators.
cerates of frailty (Yeap et al., 2013). The link may be 5. Reproductive-axis hormones, such as those related to
that reduced secretion of growth hormones is related to testosterone (assessed by plasma or urine concentra-
decreased skeletal muscle, bone mass and strength, key tions). Higher levels of circulating/free testosterone, sex
factors related to the frailty indicators of weakness and hormone–binding globulin, and DHEA-S were related
physical inactivity. to lower rates of frailty, while lower levels of luteiniz-
2. Hormones related to the vitamin D axis, including 25- ing hormone and follicle-stimulating hormone were
hydroxyvitamin D (25[OH]D), parathyroid hormone related to higher rates of frailty. In particular, low levels
(PTH), and fibroblast growth factor-23 (FGF-23; in of free testosterone were related to higher risk of inci-
serum concentrations). Lower levels of 25(OH)D and dent frailty, especially to the frailty indicators of weight
higher levels of PTH were associated with higher pre- loss and resistance (related to muscle strength; Hyde
valence rates of frailty, higher risks of incident frailty et al., 2010). These associations were independent of
(Ensrud et al., 2011; Puts et al., 2005; Sanchis et al., age (Hyde et al., 2010; Tajar et al., 2011). After adjust-
2015; Tajar et al., 2013), and higher rates of weakness ing for circulating testosterone levels, researchers found
(frailty indicator) and sarcopenia (syndrome related to that increased levels of gonadotropins were still related
frailty; Puts et al., 2005; Sanchis et al., 2015). Higher to frailty, suggesting that changes in pituitary–testicular
levels of FGF-23, which result in reductions in function might be a correlate, but not the cause, of
25(OH)D, are also related to higher rates of frailty frailty (Tajar et al., 2011). Note that DHEA-S is related
(Beben et al., 2016). These associations indicate that to both the HPA axis and secretion of reproductive
the loss of muscle mass and strength, secondary to vita- hormones (hypothalamic–pituitary–testicular), indicat-
min D deficiency, may be an underlying mechanism of ing an interrelationship between these two endocrinal
frailty. pathways. DHEA, as an anabolic hormone that facili-
3. Thyroid hormones, that is, free thyroxine. Researchers tates the building of muscle tissue, declines with age,
found higher levels of free thyroxine (even within the which may contribute to age-related loss of muscle
euthyroid range) among frail older men than among mass and strength leading to the development of weak-
their nonfrail male counterparts (Yeap et al., 2012). The ness in frailty (evidence was inconsistent; Ceci, Dur-
link between free thyroxine and frailty is particularly anti, Rossi, Savini, & Sabatini, 2011; Percheron et al.,
significant when the frailty indicators of fatigue and 2003).
weight loss are also present, suggesting that thyroid 6. Hormones related to energy metabolism and glucose–
function is probably contributing to weight loss and insulin balance, including glucose, hemoglobin A1c
reduced physical capability in frail older adults, though [HbA1c], insulin, and ghrelin (both fasting and at 2-hr
longitudinal data are needed to verify this. oral glucose tolerance test [OGTT] level; Barzilay et al.,
4. Adrenocorticoids of the hypothalamic–pituitary–adre- 2007; Kalyani, Tian, et al., 2012; Kalyani, Varadhan,
nal axis (HPA), including cortisol, epinephrine, norepi- Weiss, Fried, & Cappola, 2012a, 2012b; Walston et al.,
nephrine, dehydroepiandrosterone (DHEA), and its 2002; Zaslavsky, Walker, Crane, Gray, & Larson,
sulfated metabolite (DHEA-sulfate [DHEA-S]), 2016). Altered insulin–glucose dynamics, such as
assessed by serum concentrations. Higher levels of cor- increased fasting and 2-hr OGTT levels of glucose,
tisol (with less diurnal variation), epinephrine and nor- higher levels of HbA1c, and high levels of insulin and
epinephrine, lower levels of DHEA-S, and a higher greater degrees of insulin resistance, are related to
Wang et al. 15

higher rates of baseline frailty and greater odds of mediating effects of inflammation, as hypercoagulability both
frailty onset (Barzilay et al., 2007; Kalyani, Tian reflects and contributes to enhanced inflammation.
et al., 2012; Kalyani et al., 2012a; Walston et al.,
2002; Zaslavsky et al., 2016). Researchers also noted Immune dysfunction. An aging immune system (i.e., immunose-
lower levels of ghrelin, a fast-acting hormone produced nescence) may function well in homeostasis, but it is highly
by stomach endocrine cells that can stimulate appetite likely to fail when stressors such as acute illness, infection, and
and the release of growth hormone, among frail subjects enhanced inflammation are present. This lack of internal stress
than among nonfrail subjects (Kalyani et al., 2012b). adaptability gives rise to frailty (Clegg et al., 2013). Charac-
These results suggest that abnormal glucose response, teristics of immunosenescence include reduced stem cell pro-
such as higher levels of glucose and insulin resistance duction, slower B cell–initiated antibody response (Clegg et al.,
and lower levels of ghrelin, can reduce food intake and 2013), and changes in white blood cells (Leng, Xue, Tian,
lean body mass, which may increase the odds of frailty Walston, & Fried, 2007), such as lymphocytes (e.g., neutro-
by affecting frailty indicators related to lower energy phils and monocytes; Leng, Xue, et al., 2009). Immunosenes-
intake and muscle strength, such as weight loss, hand- cence is also related to the endocrinal and inflammatory
grip, and slow gait speed. To date, no studies have processes associated with frailty (Clegg et al., 2013; R. A.
examined the relationship between hormones related Miller, 1996). Better immune function, therefore, is related to
to energy metabolism and specific frailty indicators. lower rates of frailty.
Of the reviewed studies, eight examined the relationship of
immune function with frailty (Table 1). Identified biomarkers
Enhanced inflammation. In cases of an enhanced inflammation include albumin (Collerton et al., 2012; Hubbard et al., 2009;
process, a prolonged immune response, caused by low-grade, Walston et al., 2002), differential white cells (Baylis et al.,
chronic inflammation secondary to aging, is abnormally sensi- 2013; Leng, Hung, et al., 2009; Leng et al., 2007), neutrophils
tive and persists long after the initial inflammatory stimuli have (Collerton et al., 2012; Leng, Xue, et al., 2009), monocytes
been removed (Clegg et al., 2013; Collerton et al., 2012). (Leng, Xue, et al., 2009), neopterin (Leng et al., 2011), total
Enhanced chronic inflammation, reflected in an increased level lymphocyte counts (Collerton et al., 2012), CD8þCD28– T
of pro-inflammatory cytokines, is related to muscle weakness cells (Semba et al., 2005), and ratio of memory/naive CD8 T
of the upper and lower extremities, which manifests as the cells (Collerton et al., 2012; assessed by plasma concentrations/
frailty indicators of reduced handgrip strength and slow gait counts). Increased levels of white blood cells, including neu-
speed (Cesari et al., 2004; Zhu et al., 2016). Also related to trophils and monocytes, and neopterin, as well as CD8þ CD28–
increased levels of pro-inflammatory cytokines is enhanced T cells, were related to higher rates of frailty, while increases in
coagulation activity (i.e., hypercoagulability) that leads to sus- total lymphocyte count, albumin, and memory/naive CD8 T
tained mild clotting activity. These alterations are more com- cell ratio were related to lower rates of frailty. These associa-
mon in frail subjects (Cushman et al., 1996). tions indicate that frailty may be linked to weakened immunity,
Researchers in 23 of the reviewed studies examined the reflected by lower counts of lymphocytes, albumin, neutro-
relationship between inflammation and frailty (Table 1). They phils, monocytes, and lower levels of neopterin, a marker of
detected enhanced inflammation as indicated by increased immune system activation. Immunosenescence, indicated by a
serum concentrations of pro-inflammatory cytokines, such as decrease in the number/functionality of naive T cells and an
C-reactive protein (CRP), tumor necrosis factor a (TNF-a), and increase in the number/functionality of memory T cells and in
interleukin-6 (IL-6) as well as the secretion of related serine CD8þ CD28 T cells, is linked to frailty, although the associ-
protease (high-temperature requirement serine protease A1/ ation between memory/naive T cells and frailty went against
HtrA1). An excess of these cytokines, especially CRP, TNF- this speculation. To date, it is unclear whether biomarkers of
a, and IL-6, causes muscle breakdown and increases the risk of immunosenescence are related to specific frailty indicators.
loss of skeletal-muscle mass and strength, or sarcopenia, a
syndrome related to frailty (Hubbard & Woodhouse, 2010). Metabolic imbalance. Metabolic imbalances such as low serum
The relationship between enhanced inflammation and frailty concentrations of micronutrients (Corona et al., 2014; Urpi-
persists regardless of the presence of significant clinical comor- Sarda et al., 2015) and altered enzyme activities (Corona
bidity (Walston et al., 2002), suggesting a pathophysiological et al., 2014; Hubbard, O’Mahony, Calver, & Woodhouse,
pathway of frailty that is related to, but independent from, 2008; Matteini et al., 2008; Semba et al., 2006; Urpi-Sarda
comorbid diseases. Furthermore, markers of hypercoagulabil- et al., 2015) can disturb homeostasis and pharmacodynamics
ity are related to higher rates of frailty, such as essential pro- (Hubbard et al., 2008), preceding the development of frailty.
teins in blood clotting (Factor I: fibrinogen [Gale, Baylis, Additionally, the accumulation of metabolic end products, for
Cooper, & Sayer, 2013; Walston et al., 2002]; Factor VIII: example, carboxymethyl-lysine (CML), can cause tissue dam-
hemophilia A [Walston et al., 2002]) and an indicator of age and functional decline (Whitson et al., 2014), which are
ongoing clotting (D-dimer [Walston et al., 2002]). Although also related to frailty. Furthermore, imbalanced metabolic pro-
extant research does not clearly explain the association cesses cause dysfunction in the formation and loss of muscle
between hypercoagulability and frailty, it may be due to the cells (atrophy) and muscle fat (fat infiltration). These metabolic
16 Biological Research for Nursing XX(X)

changes in the musculoskeletal system reduce muscle mass and In general, markers of oxidative stress including lower
strength, which are related to the slowness and weakness indi- plasma levels of antioxidants, shorter telomere length, higher
cators in frailty (Visser et al., 2002, 2005). GHS/GSSG ratio, and higher levels of oxidative products are
Of the reviewed studies, 12 examined the relationship related to higher rates of frailty, though findings have been
between the metabolic process and frailty (Table 1). Identified inconsistent. In some studies, higher serum levels of oxidative
biomarkers include (a) micronutrients (carotenoids, stress markers were related to both the overall frailty status (Ble
a-tocopherol, 25(OH)D; Semba et al., 2006, polyphenols et al., 2006; Inglés et al., 2014; Matteini et al., 2008; Namioka
[Urpi-Sarda et al., 2015]); (b) metabolites (amino acids, acyl- et al., 2016; Serviddio et al., 2009; Wu et al., 2009) and the
carnitines, sphingo-, and glycerol-phospholipids [Corona et al., frailty indicator of slow gait speed (Liu et al., 2016) after
2014]); (c) indices of enzyme activity (plasma esterase [Hub- adjusting for age and sex (Inglés et al., 2014). Yet other studies
bard et al., 2008], methylmalonic acid [marker of vitamin B12 reported no association between oxidative stress (i.e., telomere
deficiency; Matteini et al., 2008]); and (d) metabolic end prod- length) and frailty (Woo et al., 2008). This finding might be due
ucts (CML [Whitson et al., 2014]). to the fact that oxidative damage on the cellular level, such as
Most micronutrients (except for polyphenols, which can be that indicated by telomere length, is largely related to genetic
assessed in urine samples), metabolites, enzyme-activity indi- polymorphisms (Gilson & Londoño-Vallejo, 2007) that may
cators, and metabolic end products are assessed by serum con- not extend to cause functional changes in multiple systems
centrations, and decreases in these micronutrients are related to (i.e., frailty). To date, no studies have reported associations
higher rates of frailty. Markers of musculoskeletal metabolism, between markers of oxidative stress and frailty indicators other
for example, loss of muscle mass and strength (sarcopenia), are than slow gait speed.
measured using intramuscular adipose tissue, magnetic reso-
nance imaging, and computed tomography ultrasonography Comprehensive biomarkers. Researchers also examined biomar-
(Addison et al., 2014; Young, Hughes, Russel, Parker, & kers representing comprehensive changes in multiple systems
Nichols, 1980). These markers of muscular metabolic imbal- for their relationships with frailty. These measures include
ance (i.e., muscle wasting, muscle strength loss, and muscle fat allostatic load, a cumulative marker representing multisystem
infiltration) are related to higher rates of frailty and usually dysregulation calculated using 13 specific biomarkers (Grue-
manifest in skeletal-muscle weakness, which is related to the newald, Seeman, Karlamangla, & Sarkisian, 2009; Szanton,
frailty indicators of reduced handgrip strength, slow gait speed, Allen, Seplaki, Bandeen-Roche, & Fried, 2009); cystatin C
and physical inactivity (Fried et al., 2001; Visser et al., 2002; (Hart et al., 2013; Sanchis et al., 2015) and glomerular filtration
Visser et al., 2005; Xue, 2011). rate (GFR; Dalrymple et al., 2013), markers of kidney function;
and hemoglobin (Sanchis et al., 2015), a marker of nutritional
Oxidative stress. Based on the free radical theory of aging (Har- status that affects multiple body systems. Among these mar-
man, 1956), accumulation of oxidative damage over time leads kers, higher alloastatic load (indicating higher stress levels),
to physiologic dysregulation, which is related to a higher inci- higher cystatin C levels, and lower GFR (indicating worse
dence of frailty (Fried et al., 2001; Howard et al., 2007). Of the kidney function) and hemoglobin levels (poor nutritional sta-
reviewed studies, eight studies examined the relationship tus) were related to higher rates of frailty, suggesting important
between oxidative stress and frailty (Table 1). Identified bio- roles of cumulative multisystem dysregulation in the develop-
markers include the following: (1) antioxidants—protectors ment of frailty.
against oxidative damage, for example, vitamin E (Ble et al.,
2006) and vitamin B12 and its marker methylmalonic acid
(Matteini et al., 2008); (2) length of telomeres, the DNA “caps”
Quality Assessment of Reviewed Studies
that protect the internal regions of the chromosomes, of which We assessed the quality of the studies using the STROBE
shortening indicates DNA replication and reduced functional- Statement (von Elm et al., 2007). According to the percentage
ity of telomerase, both reflecting oxidative damage that causes of satisfied STROBE items, we rated the quality of five (10%)
cellular aging (Woo et al., 2008); (3) glutathione (GSH)/oxi- studies “high” (>90%), of 24 (46%) studies “moderate” (75–
dized glutathione (GSSG) ratio (Serviddio et al., 2009), a mar- 90%), and of 23 (44%) studies “low” (< 75%). Low STROBE
ker of redox status and oxidative stress; and (4) oxidation end scores were caused by inadequate reporting that may or may
products, including diacron reactive oxygen metabolite not reflect a lack of rigor in methodological design. Reporting
(Namioka et al., 2016), lipoprotein phospholipase A2 (Liu inadequacies included (1) missing components: lack of infor-
et al., 2016), isoprostanes (Liu et al., 2016), 8-hydroxy-20- mation about research design in title or abstract (n ¼ 20, 38%),
deoxyguanosine (Wu, Shiesh, Kuo, & Lin, 2009), lipoperoxide lack of a flow diagram (n ¼ 48, 92%) or a detailed report of
(i.e., malondialehyde; Inglés et al., 2014; Serviddio et al., study participants per stage (n ¼ 15, 29%), or missing data in
2009), protein carbonylation, and 4-hydroxy-2,3-nonenal- pro- results (n ¼ 38, 73%); (2) insufficient data analysis without
tein plasma adducts (Inglés et al., 2014; Serviddio et al., 2009). accounting for missing data (n ¼ 40, 77%), subgroup or inter-
Researchers assessed these markers in DNA (length of telo- actions or sensitivity analysis (n ¼ 28, 54%); and (3) incom-
mere) or serum samples (concentration of other oxidative stress plete interpretation of results, for example, no discussion of the
markers). generalizability of the findings (n ¼ 23, 44%).
Wang et al. 17

Discussion Of note, mitochondrial dysfunction may be a universal bio-


logical model that underlies all of the processes related to
The primary findings in this review were (1) biological changes
frailty (Ashar et al., 2015). As we noted in a previous publica-
in the brain, endocrine, inflammatory, immune, metabolic, and
tion (Maxwell & Wang, 2017), mitochondrial function reflects
oxidative stress processes were related to frailty; however, the
cellular respiration with systemic effects across all biological
causality is unclear; (2) bidirectional interrelationships exist
systems. Recent studies reveal that mitochondrial dysfunction
within and between these biological processes, particularly
affects multiple health outcomes, such as disability (Distefano
those involving inflammation, immune, and endocrine
et al., 2017), heart failure (Rosca & Hoppel, 2010), malignant
changes; and (3) because identified biomarkers and processes
diseases (Gogvadze, Orrenius, & Zhivotovsky, 2008), and mor-
are also related to nonfrailty conditions, no single process or
tality (Ashar et al., 2015). Mitochondrial function is also
biomarker can detect changes unique to frailty.
related to the biological processes we identified in this review.
For example, mitochondria are linked to brain structure and
function through their effect on BDNF and brain plasticity
Identified Biological Processes and Frailty (Cheng, Hou, & Mattson, 2010; Raefsky & Mattson, 2017).
The reviewed studies consistently reported associations Mitochondrial function affects inflammation by regulating cel-
between biomarkers of identified biological processes and lular response to infection (Tschopp, 2011) and affects immu-
overall frailty status, but only a limited number examined how nity by modifying immune-cell activation and cellular
these biomarkers were related to specific frailty indicators, senescence (Mills, Kelly, & O’Neill, 2017). Growth and thyr-
such as slow gait speed and weakness. The causality of iden- oid hormones affect mitochondrial pathways (Clejan, Collipp,
tified associations cannot be established because the majority & Maddaiah, 1980; Sterling, Milch, Brenner, & Lazarus,
of studies were observational. In addition, since most studies 1977). Finally, mitochondrial function is linked to oxidative
were cross-sectional, it is difficult to discern whether the iden- stress through signaling to increase endogenous antioxidants
tified processes were risk factors, correlates, or consequences (i.e., GSH; J. Chen et al., 2017) and to cell and organism
of frailty. As frailty is not a unidimensional concept, and dif- metabolism through influencing the stress response (Picard
ferent frailty indicators may represent different underlying bio- et al., 2015).
logical processes (Arts et al., 2015), a deeper understanding of
the unique longitudinal dynamics of each biological process
and specific frailty indicators and how they interact over time
Interrelationships Within and Among Biological Processes
is needed (Zaslavsky et al., 2013). Although each of these processes might explain a unique aspect
Among all identified processes related to frailty, enhanced in frailty, there are intricate, bidirectional connections
inflammation, reflected by increased levels of pro- within and between these processes, as we have summarized
inflammatory cytokines (e.g., IL-6, CRP, and TNF-a), is the in Figure 2. For example, the brain–endocrine–inflammation
most studied. The close relationships between inflammation interrelationship includes the connection between brain and
and other frailty-related processes, particularly immune and endocrine changes through the hypothalamic–pituitary axis and
endocrine dysregulation as well as hypercoagulation, reveal the connection between endocrine and inflammatory changes
shared markers and regulation of the reproductive axis and through cortisol, hormones of the reproductive axis, and hor-
vitamin D axis. In fact, some studies combined these pro- mones of the vitamin D axis (Bishop, Lu, & Yankner, 2010;
cesses into an endocrine–immune–inflammation system Clegg et al., 2013). As an individual ages, the circulating level
(Baylis et al., 2013), reflecting the close interrelationship of free cortisol increases. This prolonged exposure to cortisol
among these processes. “numbs” the regulatory system in the hippocampus and miti-
Many processes, especially inflammation, are related to an gates negative feedback, leading to the release of even more
array of health conditions. Therefore, the sensitivity of using cortisol (Rohleder, Kudielka, Hellhammer, Wolf, & Kirsch-
biomarkers to detect frailty is uncertain. For example, chronic baum, 2002). Another explanation for this connection is the
inflammation is related not only to frailty but also to other decreased secretion of sex hormones secondary to aging. Sex
health conditions that are common in the elderly population, hormones can dampen the inflammatory response to pathogens
such as rheumatoid arthritis, coronary artery diseases, and dia- and prevent overreactive responses that would otherwise harm
betes (Espinoza, Jung, & Hazuda, 2012; Johansen et al., 2017; healthy body tissues (Hunt et al., 2010). Decreased levels of sex
Trevisan et al., 2017). Therefore, it is difficult to discern hormones limit the anti-inflammatory response and lead to
whether changes in inflammation biomarkers are due to the enhanced inflammation. The last explanation of this connection
underlying process of frailty or these comorbidities. Because is the PTH–vitamin D axis. Aging-related decrease in vitamin
of the lack of quantitative standards in using markers to identify D upregulates the secretion of PTH, which (1) initiates more
frailty, even when abnormal levels of biomarkers are detected secretion of IL-6 from osteoclasts (McCarty, 2005) and (2)
(according to traditional, disease-specific criteria), it is uncer- increases the level of TNF-a (Schleithoff et al., 2006).
tain (1) whether the changes in biomarkers are due to frailty or Together, reduction in these hormones and by-products of the
other health conditions and (2) how far along the individual is endocrine system (sex hormones, cortisol, and vitamin D)
in the frailty trajectory. involve brain mechanisms and increase the levels of
18 Biological Research for Nursing XX(X)

hypothalamic-pituitary axis; corsol BDNF; reduced grey maer


Brain Changes

Anoxidants; telemore length; glutathione/oxidized glutathione rao;


oxidaon end products
Oxidave Stress

Metabolic imbalance
Accmulave oxidave
Metabolic end-product: Energy metabolism; micronutrients;
damage: stressor
stressor metabolites; enzyme acvity
indicators; metabolic end-product

Enhanced CRP, TNF-alpha; IL-6; high-temperature requirement


serine protease A1/HtrA1 Frailty
inflammaon
Weakness,
slowness
Factor I, factor II,
skeletal muscle
thrombin;
reproducve Growth hormones,
IL-6, TNF-alpha
axis; vitamin hormones of Factor I, Factor
D axis reproducve axis, and VIII, D-dimer
Vitamin D axis; enhanced clong
inflammatory stressor

Total WBC counts; neutrophils; monocypte counts; neopterin; lymphocyte


immune dysfuncon counts; CD8+CD28- T cell counts; memory/naïve CD8 T cell rao
*Glucose-
insulin
Endocrine balance
Dysregulaon Growth hormones; Vitamin D axis hormones; (insulin,
hormones of hypothalamic–pituitary–adrenal axis and reproducve axis glucose,
ghrelin),

Figure 2. Interrelationships within and among biological processes. BNDF ¼ brain-derived neurotrophic factor; CRP ¼ C-reactive protein;
TNF-a ¼ tumor necrosis factor-a; IL-6 ¼ interleukin-6; WBC ¼ white blood cell.

pro-inflammatory proteins (Hunt et al., 2010), which disturbs deficiencies (Semba et al., 2006; Urpi-Sarda et al., 2015),
the intrinsic capacity necessary for a balanced stress response cause a multisystem metabolic imbalance that is related to
and increases the risk of frailty. frailty. One example of such a multisystem imbalance is the
The inflammation–coagulation connection involves shared reciprocal response between glucose and insulin that is main-
biomarkers, including coagulation factors (e.g., fibrinogen/fac- tained by endocrine function (secretion and sensitivity of
tor I) and hemostasis enzymes (e.g., thrombin, a cascading insulin and ghrelin). Once the balance of this response is
product of prothrombin, and coagulation factor II; Cushman disturbed, such as in diabetic and prediabetic subjects, imbal-
et al., 1996). Coagulation factors and enzymes (e.g., thrombin) ances in energy and protein metabolism occur (higher levels
can increase the release of pro-inflammatory cytokines (i.e., of circulating glucose and insulin with higher insulin resis-
IL-6 and TNF-a) and promote the inflammatory response tance lead to less glucose being available for energy produc-
(Davalos & Akassoglou, 2012; Robson, Shephard, & Kirsch, tion), thus contributing to frailty. Such evidence supports
1994). Inflammation, in turn, can also cause an imbalance of epidemiolgical findings of diabetes as a key risk factor for
normal coagulant that activates the hemostasis process (Mar- frailty (Espinoza et al., 2012; Johansen et al., 2017), and
getic, 2012). The bidirectional inflammation–coagulation con- ongoing trials that aim to normalize the glucose–insulin
nection contributes to greater risk of frailty. response using medications (e.g., metformin; NCT02570672).
The core of the inflammation–immunity connection Both stressors related to oxidative stress and advanced glyca-
includes an enhanced stressor (aging-related inflammation, tion end products in the metabolic process can cause enhanced
i.e., “inflammaging”) and a dysregulated immune response that inflammation (Clegg et al., 2013; Hubbard & Woodhouse,
persists (immunosenescence; Hunt et al., 2010). Reduced 2010), forming the oxidative stress—inflammation and meta-
secretions of reproductive and growth hormones as well as bolic imbalance—inflammation connections.
changes in the hormones of the vitamin D axis upregulate the Given the interconnectivity of frailty-related biological pro-
inflammatory process. Chronic inflammation, as a stressor, cesses, biomarkers representing multiple physiological impair-
also stimulates an overly reactive immune response, especially ments should be considered in frailty studies for better
when the immune system does not have a normal “finish line” prediction of adverse outcomes like disability and mortality
(such as in immunosenescence). Aging-related chronic inflam- (Hubbard & Woodhouse, 2010; Mitnitski et al., 2015; Sanchis
mation and immunosenescence thus augment each other, con- et al., 2015). These markers could include changes in inflam-
tributing to an increased risk of frailty. matory, endocrinal, cardiovascular (e.g., heart rate variability),
In addition, disturbances in the brain, endocrine, and or renal function (e.g., GFR) and markers of cumulative phy-
immune systems (Clegg et al., 2013), along with nutritional siological dysregulation (e.g., allostatic overload; Clegg et al.,
Wang et al. 19

2013; Dalrymple et al., 2013; Gruenewald et al., 2009; Hart hormones), authors should report the time and procedures for
et al., 2013; Szanton et al., 2009; Varadhan et al., 2009). sample collection.
The interrelationships within and among these processes The researchers also used different units of measure and
also make it challenging to unravel the causality underlying different risk-group categorizations based on biomarker levels
frailty. It is difficult to discern which process might be the across studies. In some studies, researchers treated levels of
trigger in the cascade of multisystem frailty development. Most biomarkers as continuous variables (i.e., used actual level, log
of the studies in the present review only examined one process or natural log-transformed z scores), while in other studies, they
using a cross-sectional design without accounting for potential used ranges of measures to create categories in a variety of
interrelationship with other processes. Although inflammation ways. For example, Hubbard, O’Mahony, Savva, Calver, and
is the most studied process involved in frailty, it is unlikely the Woodhouse (2009) treated the level of CRP as a continuous
cause of all the other processes and, consequently, is unlikely to variable in inferential analysis, while Zhu et al. (2016) categor-
be the only biological mechanism of frailty. ized the levels of CRP into low/reference (<1 mg/dl), inter-
mediate (1.01–3 mg/dl), and high (>3 mg/dl), and Collerton
et al. (2012) categorized CRP levels into low (< 1.2 mg/dl),
intermediate/reference (1.2–6 mg/dl), and high (>6 mg/dl). It
Associations Between Frailty and Biomarkers may be that the differences in the strength of the associations
Although the trends of associations between frailty and biomar- between CRP and frailty could be attributed to this variance,
kers are mostly consistent across studies—for example, that with Hubbard et al. (2009) reporting a log-transformed coeffi-
higher levels of pro-inflammatory cytokines are related to cient of 0.221 for CRP’s correlation with the frailty index, Zhu
higher rates of frailty, while higher levels of antioxidants are et al. (2016) reporting odds ratios of having a frailty phenotype
related to lower rates of frailty—variance exists in the strengths of 1.19 for intermediate versus low CRP and 1.43 for high
of reported associations (Online Supplemental Material, Table versus low CRP, and Collerton et al. (2012) reporting an odds
1). This variance might be due to a few reasons. First, it could ratio of 1.78 for high versus intermediate. The various methods
be due to variance in the measures for frailty. The choice of we found that researchers used to measure biomarkers and
frailty measure depends on the purpose of research. When the determine high-risk biomarker groups underscore the challenge
purpose is to examine the biological underpinnings and etiol- of interpreting findings across studies and present a threat to
ogy of frailty, researchers should use instruments with psycho- external validity.
metric validity that do not also include measures of disease, Third, the variance in sample characteristics (age, health
comorbidity, or disability (Buta et al., 2016), due to the inher- condition, cognitive functioning) and analytical approaches,
ent differences in the conceptualizations of frailty, comorbid- as reported earlier, might also explain why the strength of
ity, and disability (Fried, Ferrucci, Darer, Williamson, & reported associations varied across studies. For example, the
Anderson, 2004). We noted that the reviewed studies use a mean age of the samples in the reviewed studies was 75 years.
variety of frailty measures, including self-developed measures Individuals of this age might present different risk profiles
that had not been validated previously and those that also mea- related to biomarkers (e.g., different cutoff points to determine
sured diseases and comorbidities (Morley et al., 2012; Rock- high risk) than would the very old (85 years and older), who
wood et al., 2005). were included in some of the samples. The samples also varied
Second, this variance in the findings regarding the associa- across studies in their health conditions and cognitive function-
tions of biomarkers with frailty might also be due to variance in ing. In 37 (71%) studies, researchers excluded subjects with
the methods for measuring and sources of biomarkers. specific health conditions, such as acute infection, certain
Researchers in these studies took a variety of approaches to chronic diseases and functional disabilities, and in 18 (35%)
measuring the level of biomarkers, perhaps at least partially studies, researchers excluded subjects with cognitive impair-
because standards for biomarker assessment have changed over ments. Even among these studies, the specific health conditions
time. For instance, investigators used traditional ELISA kits to that were exclusionary and the criteria for determining cogni-
assess the levels of inflammatory markers, such as IL-6, TNF- tive impairment differed, limiting the comparability and gen-
a, and CRP in some studies (Coelho et al., 2012; Hubbard et al., eralizability of findings. In particular, as cognitive functioning
2008; Lai et al., 2014; Reiner et al., 2009), while in a more is closely related to frailty (as described in “Changes in the
recent study (Zhu et al., 2016), researchers used chemilumines- Brain,” above), it is possible that subjects who are cognitively
cence immunoassay, a different approach with better sensitiv- impaired may be the frailest subgroup. Thus, excluding these
ity, to measure CRP. subjects from frailty research might have artificially created a
The sources of biomarkers also differed. Some studies (e.g., sample with a decreased risk of frailty and a reduced number of
Zaslavsky et al., 2016) described the collection of samples frailty indicators, which may not represent the whole spectrum
(fasting or nonfasting blood samples and urine samples in the of frailty development.
morning or during the day), while in other studies, authors It is also worth noting that the majority of participants in the
provided no information as to when and how they collected reviewed studies were Caucasian, and frailty and related bio-
the samples. Depending on the biomarker being assessed (espe- logical phenomena may not manifest in the same ways across
cially those with circadian variation in levels, such as racial and ethnic groups. For example, older people who
20 Biological Research for Nursing XX(X)

identify as part of a minority racial group have higher levels of of many biomarkers (e.g., CRP, IL-6, insulin, and albumin) are
chronic stress and allostatic load, the “wear and tear of the readily available in electronic health records (EHRs), we
body” that accumulates in response to persistent, repeated should also consider using data-driven analytical methods,
stressors (i.e., internal or external challenges to the body), when such as machine learning and data mining, to proactively cap-
compared to older people who identify as White. Thus, those ture data from EHR, reveal novel and nonlinear relationships
who identify as members of racial minority groups may have an between biomarkers and frailty, and proactively identify high-
increased risk of frailty and development of health conditions risk patients using information from routine tests (Westra et al.,
such as cardiovascular disease. Racial disparities in the devel- 2015). Lastly, an emerging body of work has shown promise
opment of frailty, particularly among subgroups that contend for the identification of genetic biomarkers related to frailty
with persistent stressors (Seeman et al., 1997), should be fur- (Mekli et al., 2017). For example, specific genomes and epi-
ther investigated. genetic markers related to inflammatory pathways and the
Besides the variation we observed in the strength of reported stress response have been found in significantly higher levels
associations between biomarkers and frailty across studies, in frail people, highlighting the critical role of enhanced
there are additional challenges for understanding the unique inflammation in the development of frailty. Genomic research
role of biomarkers in the development of frailty. First, should be conducted in further breadth and depth to explore the
researchers have studied a variety of biomarkers for each pro- development of specific frailty indicators among older adults
cess. Second, some biomarkers are related to more than one with different chronic conditions in order to improve the pre-
process, such as IL-6, which plays a role in both inflammation cision in risk prediction and design of targeted interventions for
and immune processes; fibrinogen, which plays a role in the people who present different combinations of frailty indicators
inflammation and coagulation processes; and the glucose–insu- and different chronic conditions.
lin balance, which plays a role in endocrine and metabolic As changes in these biological processes might be undetect-
processes. Third, even within the same biological process, able in basal conditions, we need to repeatedly measure asso-
some markers of frailty also indicate diseases that are related ciated biomarkers when stressors are present (Varadhan et al.,
to frailty but have a different etiology. For example, research- 2008), such as during acute exacerbation of chronic illnesses.
ers used IL-6, CRP, and TNF-a as markers of enhanced inflam- There is also a need for more diverse patient populations, as
mation, but these markers are related to both frailty and chronic over half the reviewed studies used data from population-based
disease, as previously discussed. Due to the complexity of the cohorts that were conducted in a limited number of geographic
roles of the biomarkers selected to represent each process, regions. Additionally, despite the racial differences in biomar-
overlap in biomarkers among different processes, and lack of kers and frailty (Leng, Hung, et al., 2009; Matteini et al., 2008),
ability to determine the direction of causality between biomar- the reviewed studies were conducted primarily among Cauca-
kers and frailty, it is not possible to determine whether the sian subjects. Since interventional studies are the best approach
identified processes are drivers for frailty, compensatory for investigating causal relationship, a first step might be to
responses to frailty, or indicators of other pathophysiological analyze existing data from multiracial samples in existing
processes related to frailty. interventional studies.

Directions for Future Research Implications for Nursing Science


Future research is needed to determine which biological pro- Nurses are on the front lines for assessing geriatric syndromes
cess triggers biomarker changes that occur prior to other pro- (e.g., frailty) and coordinating interventions. Since changes in
cesses, as well as the frailty indicators that emerge first. We mechanistic biomarkers precede the development of frailty
also need to understand whether this process is modifiable and, symptomology, knowledge about frailty biomarkers provides
if so, what might be the best way to intervene in this initial an opportunity to detect frailty early on, to monitor the devel-
process. Given the close interrelationships among the identified opment of frailty, and to characterize different frailty subtypes
biological processes related to frailty, it is possible that (1) the and trajectories in both natural and interventional conditions.
causality does not reside in a single biological process but Early detection of the development of frailty is of particular
rather is related to these interrelationships and (2) instead of importance to older adults with multimorbidity (prevalence
a single biomarker, a composite index of multisystem biomar- 25%; Mercer, Smith, Wyke, O’Dowd, & Watt, 2009), as the
kers with validated diagnostic thresholds might be more sensi- processes related to frailty are also related to other conditions.
tive for the early detection of frailty and specific indicators of Biomarkers associated with these processes, such as insulin,
frailty. For example, an index including CRP, IL-6, and other albumin, CRP and IL-6, that were commonly assessed in the
inflammatory markers might be used to identify subjects with reviewed studies and are readily available in clinical nursing
higher risk of slower gait speed, weaker grip strength, and practice, should be monitored closely among patients who have
lower physical activity (Sarkisian, Gruenewald, John Boscar- already shown risk factors for frailty, such as those with dia-
din, & Seeman, 2008), while an index with albumin, glucose, betes and prediabetes. Biomarkers are also useful for identify-
insulin, and endocrinal markers might be used to detect people ing appropriate interventions specific to related biological
at risk of weight loss and exhaustion. In addition, as the levels mechanisms and evaluating the effectiveness of interventions
Wang et al. 21

(e.g., physical exercise and [pre]rehabilitation, nutritional sup- assessment of the direction of causality in the relationship
port, and psychosocial consultation) to not only slow the prog- between identified biological processes and frailty and the
ress of frailty but also proactively delay its onset and/or overlap among biomarkers used to detect changes in each pro-
decrease the risk of adverse outcomes. These interventions are cess present additional challenges for our understanding of
of particular importance for vulnerable (i.e., prefrail) older frailty. Frailty represents a lack of adaptive capacity of the
adults prior to invasive or high-risk procedures, treatment human body that occurs over time and leads to functional
(e.g., chemotherapy), or surgery (e.g., elective hip- decline and eventually death. The search for a specific biolo-
replacement surgeries) that may trigger the stress response and gical process to identify as the primary mechanism of frailty
thus precipitate the development or worsening of frailty. As may be fruitless in light of the complex and multifactorial
nurses play an essential role in the coordination of these inter- nature of frailty. Rather, a multisystem approach for assess-
ventions (Metzelthin et al., 2012), knowledge about frailty and ment and intervention is likely necessary.
related biological processes can help them identify patients
who are most likely to respond (or not) to interventions that Acknowledgment
have the potential to improve the quality of life (Nicholson, The authors wanted to thank Ms. Elizabeth Connor for her assistance
Meyer, Flatley, & Holman, 2013). with quality assessment of selected studies.

Author Contributions
Strengths and Limitations JW contributed to the conceptualization, literature search, data anal-
To our knowledge, this is the first review to synthesize knowl- ysis, structure, draft, and revision of the manuscript; CM contributed
edge about biological processes related to frailty, identify inter- to the conceptualization, data analysis, structure and revision of the
relationships among these processes, and discuss how the manuscript; FY contributed to the conceptualization, structure and
interrelationships affect the assessment of and intervention for revision of the manuscript.
frailty. Readers should interpret the findings with the following
Declaration of Conflicting Interests
limitations in mind. First, because of the scarcity of studies that
The author(s) declared no potential conflicts of interest with respect to
assessed interventions for frailty and included measures of bio-
the research, authorship, and/or publication of this article.
markers (n ¼ 1), we were not able to draw conclusions about
the direction of causality of the observed associations. Second, Funding
we did not, in our literature search, specify the biological pro- The author(s) received no financial support for the research, author-
cesses associated with frailty, such as inflammation, endocrine, ship, and/or publication of this article.
and oxidative stress. Specifically, we did not use search terms
corresponding to each biomarker in the literature search, for Supplemental Material
example, glucose or insulin, as that was beyond scope of this Supplemental material for this article is available online.
review, and we might have missed studies if they did not use
the selected Medical Subject Headings (MeSH) terms (e.g., References
biological phenomena and biomarkers). Third, in 12 of the Addison, O., Drummond, M. J., LaStayo, P. C., Dibble, L. E., Wende,
studies, authors did not report the strength of the association A. R., McClain, D. A., & Marcus, R. L. (2014). Intramuscular fat
between the biomarkers and frailty, thus we could not include and inflammation differ in older adults: The impact of frailty and
specific parameters of identified associations from those stud- inactivity. Journal of Nutrition Health and Aging, 18, 532–538.
ies in Supplemental Table 1. Lastly, since 51 of the 52 doi:10.1007/s12603-014-0019-1
reviewed studies were observational, we used the STROBE Arts, M. H., Collard, R. M., Comijs, H. C., Naude, P. J., Risselada, R.,
checklist to appraise quality, yet the checklist does not differ- Naarding, P., . . . Voshaar, R. C. (2015). Relationship between
entiate weaknesses in methodological design from those in physical frailty and low-grade inflammation in late-life depression.
reporting. Despite these limitations, our review fills a gap in Journal of American Geriatric Society, 63, 1652–1657. doi:10.
understanding about the underlying biological processes 1111/jgs.13528
related to frailty, and our findings are highly relevant for devel- Ashar, F. N., Moes, A., Moore, A. Z., Grove, M. L., Chaves, P. H. M.,
oping mechanism-specific interventions for frailty. Coresh, J., . . . Arking, D. E. (2015). Association of mitochondrial
DNA levels with frailty and all-cause mortality. Journal of Mole-
cular Medicine, 93, 177–186. doi:10.1007/s00109-014-1233-3
Conclusion Balducci, L., & Extermann, M. (2000). Management of cancer in the
Frailty involves multiple biological processes, especially those older person: A practical approach. Oncologist, 5, 224–237. doi:
related to inflammatory, immune, and endocrinal function. To 10.1634/theoncologist.5-3-224
date, the direction of causality of the relationship between these Bandeen-Roche, K., Seplaki, C. L., Huang, J., Buta, B., Kalyani, R. R.,
processes and frailty remains uncertain, and no biomarkers Varadhan, R., . . . Kasper, J. D. (2015). Frailty in older adults: A
have been identified that can differentiate between changes nationally representative profile in the United States. Journals of
related to frailty and those related to comorbid diseases. The Gerontology Series A: Biological Sciences and Medical Sciences,
lack of longitudinal and interventional studies to inform the 70, 1427–1434.
22 Biological Research for Nursing XX(X)

Barzilay, J. I., Blaum, C., Moore, T., Xue, Q. L., Hirsch, C. H., probe for real-time imaging of glutathione dynamics in mitochon-
Walston, J. D., & Fried, L. P. (2007). Insulin resistance and inflam- dria. ACS Sensors, 2, 1257–1261.
mation as precursors of frailty: The Cardiovascular Health Study. Chen, W. T., Chou, K. H., Liu, L. K., Lee, P. L., Lee, W. J., Chen, L.
Archives of Internal Medicine, 167, 635–641. doi:10.1001/ K., . . . Lin, C. P. (2015). Reduced cerebellar gray matter is a
archinte.167.7.635 neural signature of physical frailty. Human Brain Mapping, 36,
Baylis, D., Bartlett, D. B., Syddall, H. E., Ntani, G., Gale, C. R., 3666–3676. doi:10.1002/hbm.22870
Cooper, C., . . . Sayer, A. A. (2013). Immune-endocrine biomar- Cheng, A., Hou, Y., & Mattson, M. P. (2010). Mitochondria and
kers as predictors of frailty and mortality: A 10-year longitudinal neuroplasticity. ASN Neuro, 2, e00045. doi:10.1042/an20100019
study in community-dwelling older people. Age, 35, 963–971. doi: Choi, J., Ahn, A., Kim, S., & Won, C. W. (2015). Global prevalence of
10.1007/s11357-012-9396-8 physical frailty by Fried’s criteria in community-dwelling elderly
Beben, T., Ix, J. H., Shlipak, M. G., Sarnak, M. J., Fried, L. F., with national population-based surveys. Journal of the American
Hoofnagle, A. N., . . . Rifkin, D. E. (2016). Fibroblast growth Medical Directors Association, 16, 548–550. doi:10.1016/j.jamda.
factor-23 and frailty in elderly community-dwelling individuals: 2015.02.004
The cardiovascular health study. Journal of American Geriatrics Clegg, A., Young, J., Iliffe, S., Rikkert, M. O., & Rockwood, K.
Society, 64, 270–276. doi:10.1111/jgs.13951 (2013). Frailty in elderly people. Lancet, 381, 752–762. doi:10.
Bishop, N. A., Lu, T., & Yankner, B. A. (2010). Neural mechanisms of 1016/S0140-6736(12)62167-9
ageing and cognitive decline. Nature, 464, 529–535. doi:10.1038/ Clejan, S., Collipp, P. J., & Maddaiah, V. T. (1980). Hormones
nature08983 and liver mitochondria: Influence of growth hormone, thyrox-
Ble, A., Cherubini, A., Volpato, S., Bartali, B., Walston, J. D., Wind- ine, testosterone, and insulin on thermotropic effects of respira-
ham, B. G., . . . Ferrucci, L. (2006). Lower plasma vitamin E levels tion and fatty acid composition of membranes. Archives of
are associated with the frailty syndrome: The InCHIANTI study. Biochemistry and Biophysics, 203, 744–752. doi:10.1016/
Journals of Gerontology Series A: Biological Sciences and Medi- 0003-9861(80)90234-9.
cal Sciences, 61, 278–283. Coelho, F. M., Pereira, D. S., Lustosa, L. P., Silva, J. P., Dias, J. M. D.,
Brouwers, B., Dalmasso, B., Hatse, S., Laenen, A., Kenis, C., Swerts, Dias, R. C. D., . . . Pereira, L. S. M. (2012). Physical therapy inter-
E., . . . Wildiers, H. (2015). Biological ageing and frailty markers vention (PTI) increases plasma brain-derived neurotrophic factor
in breast cancer patients. Aging, 7, 319–333. (BDNF) levels in non-frail and pre-frail elderly women. Archives
Buchman, A. S., Yu, L., Wilson, R. S., Boyle, P. A., Schneider, J. A., of Gerontology and Geriatrics, 54, 415–420. doi:10.1016/j.arch-
& Bennett, D. A. (2014). Brain pathology contributes to simulta- ger.2011.05.014
neous change in physical frailty and cognition in old age. Journals Colburn, W., DeGruttola, V., DeMets, D., Downing, G., Hoth, D., &
of Gerontology Series A: Biological Sciences and Medical Oates, J., . . . Biomarkers Definitions Working Group. (2001). Bio-
Sciences, 69, 1536–1544. doi:10.1093/gerona/glu117 markers and surrogate endpoints: Preferred definitions and con-
Buchman, A. S., Yu, L., Wilson, R. S., Schneider, J. A., & Bennett, D. ceptual framework. . Clinical Pharmacology & Therapeutics, 69,
A. (2013). Association of brain pathology with the progression of 89–95.
frailty in older adults. Neurology, 80, 2055–2061. doi:10.1212/ Collard, R. M., Boter, H., Schoevers, R. A., & Oude Voshaar, R. C.
WNL.0b013e318294b462 (2012). Prevalence of frailty in community-dwelling older persons:
Buta, B. J., Walston, J. D., Godino, J. G., Park, M., Kalyani, R. R., A systematic review. Journal of the American Geriatrics Society,
Xue, Q. L., . . . Varadhan, R. (2016). Frailty assessment instru- 60, 1487–1492. doi:10.1111/j.1532-5415.2012.04054.x
ments: Systematic characterization of the uses and contexts of Collerton, J., Martin-Ruiz, C., Davies, K., Hilkens, C. M., Isaacs, J.,
highly-cited instruments. Ageing Research Reviews, 26, 53–61. Kolenda, C., . . . Kirkwood, T. B. L. (2012). Frailty and the role of
doi:10.1016/j.arr.2015.12.003 inflammation, immunosenescence and cellular ageing in the very
Calvani, R., Marini, F., Cesari, M., Tosato, M., Anker, S. D., von old: Cross-sectional findings from the Newcastle 85þ study.
Haehling, S., . . . Marzetti, E. (2015). Biomarkers for physical Mechanisms of Ageing and Development, 133, 456–466. doi:10.
frailty and sarcopenia: State of the science and future develop- 1016/j.mad.2012.05.005
ments. Journal of Cachexia, Sarcopenia, and Muscle, 6, Cooper, C., Dere, W., Evans, W., Kanis, J. A., Rizzoli, R., Sayer,
278–286. doi:10.1002/jcsm.12051 A. A., . . . Reginster, J. Y. (2012). Frailty and sarcopenia: Def-
Ceci, R., Duranti, G., Rossi, A., Savini, I., & Sabatini, S. (2011). initions and outcome parameters. Osteoporosis International,
Skeletal muscle differentiation: Role of dehydroepiandrosterone 23, 1839–1848. doi:10.1007/s00198-012-1913-1
sulfate. Hormone and Metabolic Research, 43, 702–707. Corona, G., Polesel, J., Fratino, L., Miolo, G., Rizzolio, F., Crivellari,
Cesari, M., Penninx, B. W., Pahor, M., Lauretani, F., Corsi, A. M., D., . . . Toffoli, G. (2014). Metabolomics biomarkers of frailty in
Williams, G. R., . . . Ferrucci, L. (2004). Inflammatory markers elderly breast cancer patients. Journal of Cellular Physiology, 229,
and physical performance in older persons: The InCHIANTI study. 898–902. doi:10.1002/jcp.24520
Journals of Gerontology Series A: Biological Sciences and Medi- Corwin, E. J., & Ferranti, E. (2016). Integration of biomarkers to
cal Sciences, 59, M242–M248. advance precision nursing interventions for family research across
Chen, J., Jiang, X., Zhang, C., MacKenzie, K. R., Stossi, F., Palzkill, the lifespan. Nursing Outlook, 64, 292–298. doi:10.1021/acssen-
T., . . . Wang, J. (2017). Reversible reaction-based fluorescent sors.7b00425
Wang et al. 23

Cushman, M., Psaty, B. M., Macy, E., Bovill, E. G., Cornell, E. S., phenotype. Journals of Gerontology Series A: Biological Sciences
Kuller, L. H., & Tracy, R. P. (1996). Correlates of thrombin mar- and Medical Sciences, 56, M146–M157.
kers in an elderly cohort free of clinical cardiovascular disease. Gale, C. R., Baylis, D., Cooper, C., & Sayer, A. A. (2013). Inflam-
Arteriosclerosis, Thrombosis, and Vascular Biology, 16, matory markers and incident frailty in men and women: The Eng-
1163–1169. lish Longitudinal Study of Ageing. Age, 35, 2493–2501. doi:10.
Dalrymple, L. S., Katz, R., Rifkin, D. E., Siscovick, D., Newman, A. 1007/s11357-013-9528-9
B., Fried, L. F., . . . Shlipak, M. G. (2013). Kidney function and Gilson, E., & Londoño-Vallejo, J. A. (2007). Telomere length profiles
prevalent and incident frailty. Clinical Journal of the American in humans: All ends are not equal. Cell Cycle, 6, 2486–2494. doi:
Society of Nephrology, 8, 2091–2099. doi:10.2215/cjn.02870313 10.4161/cc.6.20.4798
Davalos, D., & Akassoglou, K. (2012). Fibrinogen as a key regulator Gogvadze, V., Orrenius, S., & Zhivotovsky, B. (2008). Mitochondria
of inflammation in disease. Seminars in Immunopathology, 34, in cancer cells: What is so special about them? Trends in Cell
43–62. doi:10.1007/s00281-011-0290-8 Biology, 18, 165–173.
Distefano, G., Standley, R. A., Zhang, X., Carnero, E. A., Yi, F., Gruenewald, T. L., Seeman, T. E., Karlamangla, A. S., & Sarkisian, C.
Cornnell, H. H., & Coen, P. M. (2017). Physical activity unveils A. (2009). Allostatic load and frailty in older adults. Journal of
the relationship between mitochondrial energetics, muscle quality American Geriatrics Society, 57, 1525–1531. doi:10.1111/j.1532-
and physical function in older adults. bioRxiv, 164160. doi.org/10. 5415.2009.02389.x
1101/164160. Harman, D. (1956). Aging: A theory based on free radical and radia-
Ehlenbach, W. J., Larson, E. B., Randall Curtis, J., & Hough, C. L. tion chemistry. Journal of Gerontology, 11, 298–300.
(2015). Physical function and disability after acute care and critical Hart, A., Paudel, M. L., Taylor, B. C., Ishani, A., Orwoll, E. S., Caw-
illness hospitalizations in a prospective cohort of older adults. thon, P. M., & Ensrud, K. E. (2013). Cystatin C and frailty in older
Journal of the American Geriatrics Society, 63, 2061–2069. men. Journal of American Geriatrics Society, 61, 1530–1536. doi:
El Assar, M., Angulo, J., Carnicero, J. A., Walter, S., Garcia-Garcia, F.
10.1111/jgs.12413
J., Lopez-Hernandez, E., . . . Rodriguez-Manas, L. (2017). Frailty Howard, C., Ferrucci, L., Sun, K., Fried, L. P., Walston, J., Varadhan,
is associated with lower expression of genes involved in cellular
R., . . . Semba, R. D. (2007). Oxidative protein damage is associ-
response to stress: Results from the Toledo Study for Healthy
ated with poor grip strength among older women living in the
Aging. Journal of the American Medical Directors Association,
community. Journal of Applied Physiology, 103, 17–20.
18, 734.e731–734.e737. doi:10.1016/j.jamda.2017.04.019
Hubbard, R. E., O’Mahony, M. S., Calver, B. L., & Woodhouse, K. W.
Ensrud, K. E., Blackwell, T. L., Cauley, J. A., Cummings, S. R.,
(2008). Plasma esterases and inflammation in ageing and frailty.
Barrett-Connor, E., Dam, T. T., . . . Cawthon, P. M. (2011). Cir-
European Journal of Clinical Pharmacology, 64, 895–900. doi:10.
culating 25-hydroxyvitamin D levels and frailty in older men: The
1007/s00228-008-0499-1
osteoporotic fractures in men study. Journal of the American Ger-
Hubbard, R. E., O’Mahony, M. S., Savva, G. M., Calver, B. L., &
iatrics Society, 59, 101–106. doi:10.1111/j.1532-5415.2010.
Woodhouse, K. W. (2009). Inflammation and frailty measures in
03201.x
older people. Journal of Cellular and Molecular Medicine, 13,
Ensrud, K. E., Ewing, S. K., Taylor, B. C., Fink, H. A., Stone, K. L.,
3103–3109. doi:10.1111/j.1582-4934.2009.00733.x
Cauley, J. A., . . . Cawthon, P. M. (2007). Frailty and risk of falls,
Hubbard, R. E., & Woodhouse, K. W. (2010). Frailty, inflammation
fracture, and mortality in older women: The study of osteoporotic
and the elderly. Biogerontology, 11, 635–641. doi:10.1007/
fractures. Journals of Gerontology Series A: Biological Sciences
and Medical Sciences, 62, 744–751. s10522-010-9292-5
Espinoza, S. E., Jung, I., & Hazuda, H. (2012). Frailty transitions in Hunt, K. J., Walsh, B. M., Voegeli, D., & Roberts, H. C. (2010).
the San Antonio longitudinal study of aging. Journal of American Inflammation in aging part 1: Physiology and immunological
Geriatric Society, 60, 652–660. doi:10.1111/j.1532-5415.2011. mechanisms. Biological Research for Nursing, 11, 245–252.
03882.x Hyde, Z., Flicker, L., Almeida, O. P., Hankey, G. J., McCaul, K. A.,
Fedarko, N. S. (2011). The biology of aging and frailty. Clinics in Chubb, S. A., & Yeap, B. B. (2010). Low free testosterone predicts
Geriatric Medicine, 27, 27–37. doi:10.1016/j.cger.2010.08.006 frailty in older men: The health in men study. Journal of Clinical
Fougere, B., Vellas, B., Van Kan, G. A., & Cesari, M. (2015). Iden- Endocrinology & Metabolism, 95, 3165–3172. doi:10.1210/jc.
tification of biological markers for better characterization of older 2009-2754
subjects with physical frailty and sarcopenia. Translational Neu- Inglés, M., Gambini, J., Carnicero, J. A., Garcı́a-Garcı́a, F. J., Rodrı́-
roscience, 6, 103–110. doi:10.1515/tnsci-2015-0009 guez-Mañas, L., Olaso-González, G., . . . Viña, J. (2014). Oxida-
Fried, L. P., Ferrucci, L., Darer, J., Williamson, J. D., & Anderson, G. tive stress is related to frailty, not to age or sex, in a geriatric
(2004). Untangling the concepts of disability, frailty, and comor- population: Lipid and protein oxidation as biomarkers of frailty.
bidity: Implications for improved targeting and care. Journals of Journal of the American Geriatrics Society, 62, 1324–1328.
Gerontology Series A: Biological Sciences and Medical Sciences, Inglés, M., Gambini, J., Mas-Bargues, C., Garcı́a-Garcı́a, F. J., Viña, J.,
59, M255–M263. & Borrás, C. (2017). Brain-derived neurotrophic factor as a
Fried, L. P., Tangen, C. M., Walston, J., Newman, A. B., Hirsch, C., & marker of cognitive frailty. Journals of Gerontology Series A:
Gottdiener, J., . . . Cardiovascular Health Study Collaborative Biological Sciences and Medical Sciences, 72, 450–451. doi:10.
Research Group. (2001). Frailty in older adults evidence for a 1093/gerona/glw145
24 Biological Research for Nursing XX(X)

Johansen, K. L., Dalrymple, L. S., Delgado, C., Chertow, G. M., Leng, S. X., Xue, Q. L., Tian, J., Walston, J. D., & Fried, L. P. (2007).
Segal, M. R., Chiang, J., . . . Kaysen, G. A. (2017). Factors asso- Inflammation and frailty in older women. Journal of the American
ciated with frailty and its trajectory among patients on hemodia- Geriatrics Society, 55, 864–871.
lysis. Clinical Journal of the American Society of Nephrology, 12, Lindsay, R. M. (1994). Neurotrophic growth factors and neurodegen-
1100–1108. doi:10.2215/cjn.12131116 erative diseases: Therapeutic potential of the neurotrophins and
Joseph, B., Pandit, V., Zangbar, B., Kulvatunyou, N., Hasmi, A., ciliary neurotrophic factor. Neurobiology of Aging, 15, 249–251.
Green, D., . . . Rhee, P. (2014). Superiority of frailty over age in Liu, C. K., Lyass, A., Larson, M. G., Massaro, J. M., Wang, N.,
predicting outcomes among geriatric trauma patients: A prospec- D’Agostino, R. B. Sr., . . . Murabito, J. M. (2016). Biomarkers of
tive analysis. JAMA Surgery, 149, 766–772. doi:10.1001/jamasurg. oxidative stress are associated with frailty: The Framingham Off-
2014.296 spring Study. Age, 38, 1. doi:10.1007/s11357-015-9864-z
Kahlon, S., Pederson, J., Majumdar, S. R., Belga, S., Lau, D., Margetic, S. (2012). Inflammation and haemostasis. Biochemia Med-
Fradette, M., . . . Padwal, R. S. (2015). Association between ica, 22, 49–62.
frailty and 30-day outcomes after discharge from hospital. Mather, M., Jacobsen, L. A., & Pollard, K. M. (2015). Aging in the
Canadian Medical Association Journal, 187, 799–804. doi:10. United States. Population Bulletin, 70. Retrieved from http://www.
1503/cmaj.150100 prb.org/pdf16/aging-us-population-bulletin.pdf
Kalyani, R. R., Tian, J., Xue, Q. L., Walston, J., Cappola, A. R., Matteini, A. M., Walston, J. D., Fallin, M. D., Bandeen-Roche, K.,
Fried, L. P., . . . Blaum, C. S. (2012). Hyperglycemia and inci- Kao, W. H., Semba, R. D., . . . Stabler, S. P. (2008). Markers of B-
dence of frailty and lower extremity mobility limitations in older vitamin deficiency and frailty in older women. Journal of Nutri-
women. Journal of American Geriatrics Society, 60, 1701–1707. tion, Health, & Aging, 12, 303–308.
doi:10.1111/j.1532-5415.2012.04099.x Maxwell, C. A., & Wang, J. (2017). Understanding frailty: A nurse’s
Kalyani, R. R., Varadhan, R., Weiss, C. O., Fried, L. P., & Cappola, A. guide. Nursing Clinics of North America, 52, 349–361. doi:10.
R. (2012a). Frailty status and altered glucose-insulin dynamics. 1016/j.cnur.2017.04.003
Journals of Gerontology. Series A, Biological Sciences and Med- McCarty, M. F. (2005). Secondary hyperparathyroidism promotes the
ical Sciences, 67, 1300–1306. doi:10.1093/gerona/glr141 acute phase response—a rationale for supplemental vitamin D in
Kalyani, R. R., Varadhan, R., Weiss, C. O., Fried, L. P., & Cappola, A. prevention of vascular events in the elderly. Medical Hypotheses,
R. (2012b). Frailty status and altered dynamics of circulating 64, 1022–1026.
energy metabolism hormones after oral glucose in older women. Mekli, K., Marshall, A., Vanhoutte, B., Tampubolon, G., Nazroo, J., &
Journal of Nutrition Health and Aging, 16, 679–686. doi:10.1007/ Pendleton, N. (2017). The genetics of frailty: Summary of the
s12603-012-0369-5 results of the genetics work in the frail project. Innovation in
Lai, H. Y., Chang, H. T., Lee, Y. L., & Hwang, S. J. (2014). Associ- Aging, 1, 137–137. doi:10.1093/geroni/igx004.552
ation between inflammatory markers and frailty in institutionalized Mercer, S. W., Smith, S. M., Wyke, S., O’Dowd, T., & Watt, G. C.
older men. Maturitas, 79, 329–333. doi:10.1016/j.maturitas.2014. (2009). Multimorbidity in primary care: Developing the research
07.014 agenda. Family Practice, 26, 79–80. doi:10.1093/fampra/cmp020
Leng, S. X., Cappola, A. R., Andersen, R. E., Blackman, M. R., Metzelthin, S. F., Daniëls, R., van Rossum, E., Cox, K., Habets, H., de
Koenig, K., Blair, M., & Walston, J. D. (2004). Serum levels of Witte, L. P., & Kempen, G. I. J. M. (2012). A nurse-led interdisci-
insulin-like growth factor-I (IGF-I) and dehydroepiandrosterone plinary primary care approach to prevent disability among
sulfate (DHEA-S), and their relationships with serum community-dwelling frail older people: A large-scale process eva-
interleukin-6, in the geriatric syndrome of frailty. Aging Clinical luation. International Journal of Nursing Studies, 50, 1184–1196.
and Experimental Research, 16, 153–157. doi:10.1007/ doi:10.1016/j.ijnurstu.2012.12.016
bf03324545 Miller, D. B., & O’Callaghan, J. P. (2005). Aging, stress and the
Leng, S. X., Hung, W., Cappola, A. R., Yu, Q., Xue, Q. L., & Fried, L. hippocampus. Ageing Research Reviews, 4, 123–140. doi:10.
P. (2009). White blood cell counts, insulin-like growth factor-1 1016/j.arr.2005.03.002***
levels, and frailty in community-dwelling older women. Journals Miller, R. A. (1996). The aging immune system: Primer and prospec-
of Gerontology Series A: Biological Sciences and Medical tus. Science, 273, 70.
Sciences, 64, 499–502. doi:10.1093/gerona/gln047 Mills, E. L., Kelly, B., & O’Neill, L. A. J. (2017). Mitochondria are the
Leng, S. X., Tian, X., Matteini, A., Li, H., Hughes, J., Jain, A., . . . powerhouses of immunity. Nature Immunology, 18, 488–498. doi:
Fedarko, N. S. (2011). IL-6-independent association of elevated 10.1038/ni.3704
serum neopterin levels with prevalent frailty in community- Mitnitski, A., Collerton, J., Martin-Ruiz, C., Jagger, C., von Zglinicki,
dwelling older adults. Age and Ageing, 40, 475–481. doi:10. T., Rockwood, K., & Kirkwood, T. B. (2015). Age-related frailty
1093/ageing/afr047 and its association with biological markers of ageing. BMC Med-
Leng, S. X., Xue, Q. L., Tian, J., Huang, Y., Yeh, S. H., & Fried, L. P. icine, 13, 161. doi:10.1186/s12916-015-0400-x
(2009). Associations of neutrophil and monocyte counts with Moher, D., Liberati, A., Tetzlaff, J., & Altman, D. G (2009). Preferred
frailty in community-dwelling disabled older women: Results from reporting items for systematic reviews and meta-analyses: The
the Women’s Health and Aging Studies I. Experimental Gerontol- PRISMA statement. Journal of Clinical Epidemiology, 62,
ogy, 44, 511–516. doi:10.1016/j.exger.2009.05.005 1606–1612. doi:10.1016/j.jclinepi.2009.06.005
Wang et al. 25

Mohler, M. J., Fain, M. J., Wertheimer, A. M., Najafi, B., & Nikolich- Rockwood, K., & Mitnitski, A. (2007). Frailty in relation to the accu-
Žugich, J. (2014). The frailty syndrome: Clinical measurements mulation of deficits. Journals of Gerontology Series A: Biological
and basic underpinnings in humans and animals. Experimental Sciences and Medical Sciences, 62, 722–727.
Gerontology, 54, 6–13. doi:10.1016/j.exger.2014.01.024 Rockwood, K., Song, X., MacKnight, C., Bergman, H., Hogan, D. B.,
Morley, J. E., Malmstrom, T., & Miller, D. (2012). A simple frailty McDowell, I., & Mitnitski, A. (2005). A global clinical measure of
questionnaire (FRAIL) predicts outcomes in middle aged African fitness and frailty in elderly people. Canadian Medical Association
Americans. Journal of Nutrition, Health & Aging, 16, 601–608. Journal, 173, 489–495. doi:10.1503/cmaj.050051
Namioka, N., Hanyu, H., Hirose, D., Hatanaka, H., Sato, T., & Shi- Rohleder, N., Kudielka, B. M., Hellhammer, D. H., Wolf, J. M., &
mizu, S. (2016). Oxidative stress and inflammation are associated Kirschbaum, C. (2002). Age and sex steroid-related changes in
with physical frailty in patients with Alzheimer’s disease. Geria- glucocorticoid sensitivity of pro-inflammatory cytokine produc-
trics & Gerontology International, 17, 913–918. doi:10.1111/ggi. tion after psychosocial stress. Journal of Neuroimmunology, 126,
12804 69–77.
Nicholson, C., Meyer, J., Flatley, M., & Holman, C. (2013). The Ronning, B., Wyller, T. B., Seljeflot, I., Jordhoy, M. S., Skovlund, E.,
experience of living at home with frailty in old age: A psychosocial Nesbakken, A., & Kristjansson, S. R. (2010). Frailty measures,
qualitative study. International Journal of Nursing Studies, 50, inflammatory biomarkers and post-operative complications in
1172–1179. doi:10.1016/j.ijnurstu.2012.01.006 older surgical patients. Age and Ageing, 39, 758–761. doi:10.
Noble, E. E., Billington, C. J., Kotz, C. M., & Wang, C. (2011). The 1093/ageing/afq123
lighter side of BDNF. American Journal of Physiology—Regula- Rosca, M. G., & Hoppel, C. L. (2010). Mitochondria in heart failure.
tory, Integrative and Comparative Physiology, 300, Cardiovascular Research, 88, 40–50.
R1053–R1069. Rosmond, R., & Björntorp, P. (2000). The hypothalamic–pituitary–
Orlovsky, M., Dosenko, V., Spiga, F., Skibo, G., & Lightman, S. adrenal axis activity as a predictor of cardiovascular disease, type 2
(2014). Hippocampus remodeling by chronic stress accompanied diabetes and stroke. Journal of Internal Medicine, 247, 188–197.
by GR, proteasome and caspase-3 overexpression. Brain Research, Saliba, D., Elliott, M., Rubenstein, L. Z., Solomon, D. H., Young, R.
1593, 83–94. T., Kamberg, C. J., . . . Sloss, E. M. (2001). The Vulnerable Elders
Percheron, G., Hogrel, J. Y., Denot-Ledunois, S., Fayet, G., Forette, F. Survey: A tool for identifying vulnerable older people in the com-
, Baulieu, E. E., . . . Marini, J. F. (2003). Effect of 1-year oral munity. Journal of the American Geriatrics Society, 49,
administration of dehydroepiandrosterone to 60- to 80-year-old 1691–1699.
individuals on muscle function and cross-sectional area: A Sanchis, J., Nuñez, E., Ruiz, V., Bonanad, C., Fernandez, J., Cauli,
double-blind placebo-controlled trial. Archives of Internal Medi- O., . . . Nuñez, J. (2015). Usefulness of clinical data and biomar-
cine, 163, 720–727. kers for the identification of frailty after acute coronary syndromes.
Picard, M., McManus, M. J., Gray, J. D., Nasca, C., Moffat, C., Canadian Journal of Cardiology, 31, 1462–1468. doi:10.1016/j.
Kopinski, P. K., . . . Wallace, D. C. (2015). Mitochondrial func- cjca.2015.07.737
tions modulate neuroendocrine, metabolic, inflammatory, and tran- Sarkisian, C. A., Gruenewald, T. L., John Boscardin, W., & Seeman,
scriptional responses to acute psychological stress. Proceedings of T. E. (2008). Preliminary evidence for subdimensions of geriatric
the National Academy of Sciences of the United States of America, frailty: The MacArthur study of successful aging. Journal of the
112, E6614–6623. doi:10.1073/pnas.1515733112 American Geriatrics Society, 56, 2292–2297.
Puts, M. T., Visser, M., Twisk, J. W., Deeg, D. J., & Lips, P. (2005). Schäbitz, W. R., Steigleder, T., Cooper-Kuhn, C. M., Schwab, S.,
Endocrine and inflammatory markers as predictors of frailty. Clin- Sommer, C., Schneider, A., & Kuhn, H. G. (2007). Intravenous
ical Endocrinology, 63, 403–411. brain-derived neurotrophic factor enhances poststroke sensorimo-
Raefsky, S. M., & Mattson, M. P. (2017). Adaptive responses of tor recovery and stimulates neurogenesis. Stroke, 38, 2165–2172.
neuronal mitochondria to bioenergetic challenges: Roles in neuro- Schleithoff, S. S., Zittermann, A., Tenderich, G., Berthold, H. K.,
plasticity and disease resistance. Free Radical Biology & Medi- Stehle, P., & Koerfer, R. (2006). Vitamin D supplementation
cine, 102, 203–216. doi:10.1016/j.freeradbiomed.2016.11.045 improves cytokine profiles in patients with congestive heart fail-
Reiner, A. P., Aragaki, A. K., Gray, S. L., Wactawski-Wende, J., ure: A double-blind, randomized, placebo-controlled trial. Amer-
Cauley, J. A., Cochrane, B. B., . . . LaCroix, A. Z. (2009). Inflam- ican Journal of Clinical Nutrition, 83, 754–759.
mation and thrombosis biomarkers and incident frailty in postme- Seeman, T. E., Singer, B. H., Rowe, J. W., Horwitz, R. I., & McEwen,
nopausal women. American Journal of Medicine, 122, 947–954. B. S. (1997). Price of adaptation—allostatic load and its health
doi:10.1016/j.amjmed.2009.04.016 consequences: MacArthur studies of successful aging. Archives
Robertson, D. A., Savva, G. M., & Kenny, R. A. (2013). Frailty and of Internal Medicine, 157, 2259–2268.
cognitive impairment—A review of the evidence and causal Semba, R. D., Bartali, B., Zhou, J., Blaum, C., Ko, C. W., & Fried, L.
mechanisms. Ageing Research Reviews, 12, 840–851. doi:10. P. (2006). Low serum micronutrient concentrations predict frailty
1016/j.arr.2013.06.004 among older women living in the community. Journals of Geron-
Robson, S., Shephard, E., & Kirsch, R. (1994). Fibrin degradation tology Series A: Biological Sciences and Medical Sciences, 61,
product D-dimer induces the synthesis and release of biologically 594–599.
active IL-1b, IL-6 and plasminogen activator inhibitors from Semba, R. D., Margolick, J. B., Leng, S., Walston, J., Ricks, M. O., &
monocytes in vitro. British Journal of Haematology, 86, 322–326. Fried, L. P. (2005). T cell subsets and mortality in older
26 Biological Research for Nursing XX(X)

community-dwelling women. Experimental Gerontology, 40, Gerontology Series A: Biological Sciences and Medical
81–87. doi:10.1016/j.exger.2004.09.006 Sciences, 63, 190–195.
Serviddio, G., Romano, A. D., Greco, A., Rollo, T., Bellanti, F., Alto- Visser, M., Goodpaster, B. H., Kritchevsky, S. B., Newman, A. B.,
mare, E., & Vendemiale, G. (2009). Frailty syndrome is associated Nevitt, M., Rubin, S. M., . . . Harris, T. B. (2005). Muscle mass,
with altered circulating redox balance and increased markers of muscle strength, and muscle fat infiltration as predictors of inci-
oxidative stress. International Journal of Immunopathology and dent mobility limitations in well-functioning older persons. Jour-
Pharmacology, 22, 819–827. nals of Gerontology Series A: Biological Sciences and Medical
Shamliyan, T., Talley, K. M., Ramakrishnan, R., & Kane, R. L. Sciences, 60, 324–333.
(2013). Association of frailty with survival: A systematic literature Visser, M., Kritchevsky, S. B., Goodpaster, B. H., Newman, A. B.,
review. Ageing Research Reviews, 12, 719–736. Nevitt, M., Stamm, E., & Harris, T. B. (2002). Leg muscle mass
Sterling, K., Milch, P., Brenner, M., & Lazarus, J. (1977). Thyroid and composition in relation to lower extremity performance in
hormone action: The mitochondrial pathway. Science, 197, men and women aged 70 to 79: The health, aging and body
996–999. doi:10.1126/science.196334 composition study. Journal of the American Geriatrics Society,
Szanton, S., Allen, J., Seplaki, C., Bandeen-Roche, K., & Fried, L. 50, 897–904.
(2009). Allostatic load and frailty in the women’s health and aging von Elm, E., Altman, D. G., Egger, M., Pocock, S. J., Gøtzsche, P. C.,
studies. Biological Research for Nursing, 10, 248–256. & Vandenbroucke, J. P. (2007). The Strengthening the Reporting
Tajar, A., Lee, D. M., Pye, S. R., O’Connell, M. D., Ravindrarajah, R., of Observational Studies in Epidemiology (STROBE) Statement:
Gielen, E., . . . O’Neill, T. W. (2013). The association of frailty Guidelines for reporting observational studies. Preventive Medi-
with serum 25-hydroxyvitamin D and parathyroid hormone levels cine, 45, 247–251. doi:10.1016/j.ypmed.2007.08.012
in older European men. Age and Ageing, 42, 352–359. doi:10. Walston, J., McBurnie, M. A., Newman, A., Tracy, R. P., Kop, W. J.,
1093/ageing/afs162 Hirsch, C. H., . . . Fried, L. P. (2002). Frailty and activation of the
Tajar, A., O’Connell, M. D. L., Mitnitski, A. B., O’Neill, T. W., inflammation and coagulation systems with and without clinical
Searle, S. D., & Huhtaniemi, I. T., . . . the European Male Aging comorbidities: Results from the cardiovascular health study.
Study, G. (2011). Frailty in relation to variations in hormone levels Archives of Internal Medicine, 162, 2333–2341.
of the hypothalamic–pituitary–testicular axis in older men: Results Wang, J., Boehm, L., & Mion, L. C. (2017). Intrinsic capacity in
from the European male aging study. Journal of the American older hospitalized adults: Implications for nursing practice.
Geriatrics Society, 59, 814–821. doi:10.1111/j.1532-5415.2011. Geriatric Nursing, 38, 359–361. doi:10.1016/j.gerinurse.2017.
03398.x 06.008
Trevisan, C., Veronese, N., Maggi, S., Baggio, G., Toffanello, E. D., Westra, B. L., Clancy, T. R., Sensmeier, J., Warren, J. J., Weaver, C.,
Zambon, S., . . . Sergi, G. (2017). Factors influencing transitions & Delaney, C. W. (2015). Nursing knowledge: Big data science—
between frailty states in elderly adults: The Progetto Veneto Implications for nurse leaders. Nursing Administration Quarterly,
Anziani longitudinal study. Journal of American Geriatrics Soci- 39, 304–310. doi:10.1097/naq.0000000000000130
ety, 65, 179–184. doi:10.1111/jgs.14515 Whitson, H. E., Arnold, A. M., Yee, L. M., Mukamal, K. J., Kizer, J.
Tschopp, J. (2011). Mitochondria: Sovereign of inflammation? Eur- R., Djousse, L., . . . Zieman, S. (2014). Serum carboxymethyl-
opean Journal of Immunology, 41, 1196–1202. doi:10.1002/eji. lysine, disability, and frailty in older persons: The cardiovascular
201141436 health study. Journals of Gerontology Series A: Biological
Tsigos, C., & Chrousos, G. P. (2002). Hypothalamic–pituitary–adre- Sciences and Medical Sciences, 69, 710–716. doi:10.1093/ger-
nal axis, neuroendocrine factors and stress. Journal of Psychoso- ona/glt155
matic Research, 53, 865–871. Woo, J., Tang, N. L., Suen, E., Leung, J. C., & Leung, P. C. (2008).
Urpi-Sarda, M., Andres-Lacueva, C., Rabassa, M., Ruggiero, C., Telomeres and frailty. Mechanisms of Ageing and Development,
Zamora-Ros, R., Bandinelli, S., . . . Cherubini, A. (2015). The 129, 642–648. doi:10.1016/j.mad.2008.08.003
relationship between urinary total polyphenols and the frailty phe- World Health Organization. (2015). World report on ageing and
notype in a community-dwelling older population: The health. Geneva, Switzerland: Author. Retrieved from http://
InCHIANTI Study. Journals of Gerontology Series A: Biological www.who.int/ageing/events/world-report-2015-launch/en/
Sciences and Medical Sciences, 70, 1141–1147. doi:10.1093/ger- Wu, I. C., Shiesh, S. C., Kuo, P. H., & Lin, X. Z. (2009). High
ona/glv026 oxidative stress is correlated with frailty in elderly Chinese. Jour-
Varadhan, R., Chaves, P. H., Lipsitz, L. A., Stein, P. K., Tian, J., nal of American Geriatrics Society, 57, 1666–1671. doi:10.1111/j.
Windham, B. G., . . . Fried, L. P. (2009). Frailty and impaired 1532-5415.2009.02392.x
cardiac autonomic control: New insights from principal compo- Xue, Q. L. (2011). The frailty syndrome: Definition and natural his-
nents aggregation of traditional heart rate variability indices. Jour- tory. Clinics in Geriatric Medicine, 27, 1–15. doi:10.1016/j.cger.
nals of Gerontology Series A: Biological Sciences and Medical 2010.08.009
Sciences, 64, 682–687. doi:10.1093/gerona/glp013 Yeap, B. B., Alfonso, H., Chubb, S. A., Walsh, J. P., Hankey, G. J.,
Varadhan, R., Walston, J., Cappola, A. R., Carlson, M. C., Wand, Almeida, O. P., & Flicker, L. (2012). Higher free thyroxine levels are
G. S., & Fried, L. P. (2008). Higher levels and blunted diurnal associated with frailty in older men: The health in men study. Clinical
variation of cortisol in frail older women. Journals of Endocrinology, 76, 741–748. doi:10.1111/j.1365-2265.2011.04290.x
Wang et al. 27

Yeap, B. B., Paul Chubb, S. A., Lopez, D., Ho, K. K., Hankey, G. decade of research. Biological Research for Nursing, 15,
J., & Flicker, L. (2013). Associations of insulin-like growth 422–432. doi:10.1177/1099800412462866
factor-I and its binding proteins and testosterone with frailty Zaslavsky, O., Walker, R. L., Crane, P. K., Gray, S. L., & Larson, E. B.
in older men. Clinical Endocrinology, 78, 752–759. doi:10. (2016). Glucose levels and risk of frailty. Journals of Gerontology.
1111/cen.12052 Series A: Biological Sciences and Medical Sciences, 71,
Young, A., Hughes, I., Russell, P., Parker, M. J., & Nichols, P. J. R. 1223–1229. doi:10.1093/gerona/glw024
(1980). Measurement of quadriceps muscle wasting by ultrasono- Zhu, Y., Liu, Z., Wang, Y., Wang, Z., Shi, J., Xie, X., . . . Wang, X.
graphy. Rheumatology, 19, 141–148. doi:10.1093/rheumatology/ (2016). C-reactive protein, frailty and overnight hospital admis-
19.3.141 sion in elderly individuals: A population-based study. Archives of
Zaslavsky, O., Cochrane, B. B., Thompson, H. J., Woods, N. F., Hert- Gerontology and Geriatrics, 64, 1–5. doi:10.1016/j.archger.2015.
ing, J. R., & LaCroix, A. (2013). Frailty: A review of the first 08.009

You might also like