You are on page 1of 239

CARE COORDINATION FOR SENIOR PATIENTS WITH MULTIPLE CHRONIC DISEASES: EXAMINING THE ASSOCIATION BETWEEN ORGANIZATIONAL FACTORS

AND PATIENT OUTCOMES

A DISSERTATION

PRESENTED TO THE FACULTY OF THE HELLER SCHOOL FOR SOCIAL POLICY AND MANAGEMENT BRANDEIS UNIVERSITY

In Partial Fulfillment of the Requirement of the Degree Doctor of Philosophy

By Marian Ryan, M.P.H., C.H.E.S. February 2010 Stanley Wallack, Ph.D., Chairperson Professor and Director, Schneider Institutes for Health Policy, Heller School

UMI Number: 3391164

All rights reserved INFORMATION TO ALL USERS The quality of this reproduction is dependent upon the quality of the copy submitted. In the unlikely event that the author did not send a complete manuscript and there are missing pages, these will be noted. Also, if material had to be removed, a note will indicate the deletion.

UMI 3391164 Copyright 2010 by ProQuest LLC. All rights reserved. This edition of the work is protected against unauthorized copying under Title 17, United States Code.

ProQuest LLC 789 East Eisenhower Parkway P.O. Box 1346 Ann Arbor, MI 48106-1346

The signed copy of the signature page is on file at the Heller School for Social Policy and Management This dissertation of Marian Ryan entitled Care Coordination for Senior Patients with Multiple Chronic Diseases: Examining the Association between Organizational Factors and Patient Outcomes, directed and approved by the candidate's Committee, has been accepted by the Faculty of The Heller School for Social Policy and Management and the Graduate Faculty of Brandeis University in partial fulfillment of the requirements for the Degree of DOCTOR OF PHILOSOPHY

Lisa M. Lynch, Ph.D. Dean, The Heller School for Social Policy and Management February, 2010

Dissertation Committee: Stanley Wallack, Ph.D. (Chair) Professor and Director, Schneider Institutes for Health Policy, Heller School Jody Hoffer Gittell, Ph.D., Associate Professor and Director, MBA Program, Heller School Grant Ritter, Ph.D., Senior Scientist, Heller School Stuart Levine, M.D., MHA, Corporate Medical Director, Healthcare Partners, Inc., Assistant Clinical Professor of Internal Medicine at University of California, LA, David Geffen School of Medicine

ii

Copyright by Marian Ryan 2010

iii

ACKNOWLEDGMENTS First and foremost, I wish to acknowledge and thank the Agency for Healthcare Research and Quality and the Jewish Healthcare Foundation for their respective fellowships that enabled me to complete my doctoral coursework and dissertation research in residence at Brandeis University. Without such financial support this dream could not have been realized. I also express gratitude to the Brandeis Alumni and the Brandeis University Dean for their respective dissertation grants that partially funded the printing of my patient survey and my travel to conduct the on-site clinic interviews for my research. I thank my dissertation committee, Drs. Stanley Wallack, Jody Hoffer Gittell, Grant Ritter, and Stuart Levine for their guidance in the development of this research study. I thank the gifted professors at The Heller School who afforded me the opportunity to expand my quantitative skills and apply theory to health services research questions. I appreciate my site organization for allowing this study to be conducted, providing the secondary data, and allowing access to their clinics, physicians, and patients for my primary data collection. This study could not have been completed without the additional mentoring I received in multilevel modeling. I wish to thank Professor Xiaodong Liu, Assistant Professor, Psychology Department, Brandeis University for his time and mentoring in the specification and estimation of multilevel models. I must acknowledge and express my gratitude to the wonderful staff at the University of Michigans Center for Statistical Consultation and Research, especially Brady West and Zingling Zhang for their expert knowledge and guidance. Richard Congdon, senior programmer at Harvard School of Public Health, must be recognized for his assistance in helping me resolve my challenges with the HLM software. Lastly, I must acknowledge the multilevel modeling discussion list-serve group that has facilitated sharing between experts in the field of multilevel modeling and student neophytes eager to learn. I thank Drs. J. Hox and T. Snijders for taking the time to reply to my e-mails and sending me reference materials. I wish to thank my dissertation support group Dr. Andrew Ryan, Christina Marsh, and Karen Tyo for their support, encouragement, and constructive criticism throughout this process. Words cannot adequately express my gratitude to Karen Tyo for her expert SAS mentoring that enabled me to complete all of the required programming for this study. I wish to express my gratitude to the Sacred Heart Parish community in Newton for their spiritual nourishment and friendship throughout this arduous journey. I especially thank Sister Patricia Gallagher for her untiring love, encouragement, prayers, and support as my God-given, personal angel for this walk. I thank the breakfast group who provided me

iv

my daily dose of laughter, hot tea, and shared memories over these past years. I am grateful for the access to a senior population willing to pilot my patient survey and offer me feedback. I thank retired professor and statistician, Dr. Edgar Canty for his help with the manual input of my completed patient surveys. Finally, I must acknowledge and express my deepest gratitude to my family and friends for their love, encouragement, and inspiration. All the unexpected gifts and e-mails lifted my spirits especially during the times of self-doubt and discouragement. In choosing to relinquish a wonderful position in California and move 3,000 miles east to accept the predoctoral fellowship, my family especially my mother made tremendous sacrifices. I love and appreciate her more than words can adequately convey and must acknowledge her examples of faith, courage, and commitment that fed the initiation and completion of this journey. I thank God for a life-time of blessings and divine support that have brought me to this place. I ask God to continue to open doors of opportunity that allow me to use His gifts to make this world a better place for everyone!

ABSTRACT CARE COORDINATION FOR SENIOR PATIENTS WITH MULTIPLE CHRONIC DISEASES: EXAMINING THE ASSOCIATION BETWEEN ORGANIZATIONAL FACTORS AND PATIENT OUTCOMES

A Dissertation Presented to the Faculty of The Heller School for Social Policy and Management and the Graduate Faculty of Brandeis University, Waltham, Massachusetts By Marian Ryan, M.P.H., C.H.E.S. The Institute of Medicine has identified care coordination as a national priority to improve the quality of care. Care coordination is critical for senior patients who are challenged by our fragmented healthcare delivery system. Many senior patients have multiple chronic conditions, receive care from numerous providers across different care settings, and take multiple prescriptions. The primary care physician (PCP) is in a unique position to coordinate care and the Chronic Care Model (CCM) purports to optimally support the PCP.

The CCM posits that the redesign of physician practice organizations will result in effective physician-patient interactions and subsequently improved patient outcomes. Physician-patient relational coordination and trust, which are not included in the CCM, may play a significant role in facilitating these productive interactions between physicians and patients envisioned by the CCM framework. Therefore, the theories of RC and trust within the CCM framework guide this research.

vi

This study evaluated quantitatively the association between the Chronic Care Model components and PCP relational coordination and trust, and nationally recognized quality measures using patient and organization data from a single, multispecialty medical group with an Independent Practice Association division. The main research questions examined in this study were the following: 1) do the CCM components predict quality outcomes, 2) do PCP relational coordination and trust predict quality outcomes and 3) do RC and trust moderate patient risk covariates such as low levels of education, etc.?

The patient population was composed of managed care Medicare beneficiaries with diabetes and at least one additional chronic condition receiving care from this organization between 2004 and 2007. Longitudinal analyses were conducted using four years of medical claims and physician satisfaction data from the study organization, incorporating proxy variables (PCP communication and coordination scores) for relational coordination (RC) and trust. Cross-sectional analyses utilized primary data assessing CCM, RC and trust that were linked with respondents 2007 claims data. The cross-sectional analyses also examined two additional outcome variables end of life discussions with PCP and overall PCP satisfaction derived from the patient survey.

In all fitted Hierarchical Generalized Linear Models (HGLM) using longitudinal data and examining the log odds of the diabetes quality measures, PCP communication and coordination (combined as the proxy variable for RC) was a significant predictor. In the fitted HGLM using the cross-sectional survey-linked data, PCP RC moderated the

vii

negative impact on the diabetes quality composite measure from low education of the patient (p=0.04). Both RC and trust were significantly associated with the probability of patients having end of life discussions with their PCP (p=0.03). Lastly, the logisitic model fit with the CCM component scores from 24 clinics, 81 PCPs and 408 patients found the overall chronic care model score and the score for self-management support significant (p = 0.07 and 0.03 respectively). In this fitted model the combined variable for high RC and trust did not reach statistical significance although the coefficient was positive. Additionally, statistically significant correlations were found between the proxy variable of PCP coordination/ communication examined as a key predictor in the longitudinal analyses, and RC and trust examined as key predictors in the cross-sectional analyses.

In summary, this study found a strong association between high levels of PCP communication/coordination and diabetes quality composite measures in a senior population with significant disease burden. Moreover, the study found that PCP relational coordination and trust play an important role in end of life discussions with patients. Finally, the study supports previous research which highlights the importance of the self-management component within the CCM.

Given the growing prevalence of multiple chronic conditions among the elderly, this study provides evidence to support reimbursement for care coordination within primary care. The study also supports the current emphasis on the expansion of patient-centered medical homes within an infrastructure of the Chronic Care Model. Finally, the role of

viii

PCP partnership including RC and trust is critical to meaningful discussions with patients in primary care settings when patient preferences and options can be fully explored and prior to an emergent medical crisis.

ix

TABLE OF CONTENTS

INTRODUCTION CHAPTER 1: BACKGROUND


Study Rationale Theoretical Framework The Chronic Care Model (CCM) Relational Coordination Trust

1 4
4 9 9 11 15

CHAPTER 2: LITERATURE REVIEW


Gap in the Literature

17
23

CHAPTER 3: RESEARCH METHODS


Introduction Research site and population Research Questions: Four-Year Longitudinal Analyses Analytical Plan Longitudinal Analysis - Key Predictor Variables Dependent Variables Selected Patient Quality and Adherence Measures Dependent Variable Construction Patient Sample Selection and Variable Construction Longitudinal Analyses Covariates Cross-sectional Analyses (2007) Analysis Plan Key Variables of Interest Patient Survey Relational Coordination and Trust Patient Survey - Dependent Variables and Variable Construction Physician Survey Development PCP Survey - Key Domains of Interest Clinic Survey Variable Construction from Clinic Survey Assessments

26
26 28 31 32 32 38 44 46 52 53 56 56 57 58 63 67 68 70 73 75

CHAPTER 4: DESCRIPTIVE STATISTICS


Longitudinal Analyses: Final Patient Sample

76
76 76

Dependent Variables PCP Descriptive statistics Patient Descriptive statistics Descriptive Results of Hierarchical Clustering Cross Sectional Analyses 2007 Survey Linked Data Descriptive Statistics Patient Survey Responses Dependent Variable Descriptive Statistics Patient Descriptive Statistics PCP Descriptive Statistics Patient Survey Level-Two Clinic Descriptive Statistics Patient Survey - Level three PCP Survey Descriptive Statistics

77 79 81 84 85 85 87 88 91 93 94

CHAPTER 5: BIVARIATE RELATIONSHIPS LONGITUDINAL AND CROSSSECTIONAL SAMPLES 97


Results of Bivariate Associations Longitudinal Sample Results of Bivariate Associations Cross-sectional Sample Other Bivariate Relationships Explored 97 110 117

CHAPTER 6: HIERARCHICAL GENERALIZED LINEAR MODEL REGRESSION RESULTS


Tested Research Questions and Hypotheses Longitudinal Analyses HGLM Diabetes Screening Measure Outcome HGLM Diabetes Screening with Control Measure HGLM Diabetes Screening Composite Measure HGLM Results 30-day versus 90-day readmission findings HGLM Results Group practice Sample HGLM Results - Diabetes Screening Measure Group Practice Model HGLM Results - DM Measure with Control Group practice Model Cross-Sectional Analyses Patient Survey - HGLM Results Diabetes Screen Composite Measure Logistic Regression Results Patient Survey Discussing End of Life and Total PCP Satisfaction Logistic Regression Results Linked to CCM Components

119
119 120 121 137 142 148 149 150 156 157 157 160 161

CHAPTER 7. CONCLUSION AND IMPLICATIONS


Assessment of Study Hypotheses Benefits of HGLM Analyses in this Study Study Limitations Recommendations for Future Study Health Policy Implications

164
164 171 173 175 176

REFERENCES

179

xi

APPENDICES
Appendix A. Appendix B. Appendix C. Appendix D. Appendix E. Appendix F. Appendix G. Appendix H. Appendix I. The Chronic Care Model Model I Distributions for PCP Key Domains Correlation between different constructions of PCP Domains Correlation Matrix PCP Domain Scores Patient Survey Packet Correlation Matrix MD RC, trust, PCC, and PACIC PCP Survey Packet ACIC with introduction Correlation Matrix PredCMCD and RC and Trust

190
191 192 194 197 198 206 207 214 222

List of Tables

Table 1. Outcome Variables ............................................................................................. 27 Table 2. Organization's Physician Satisfaction Surve ...................................................... 40 Table 3. PCP Domains ...................................................................................................... 42 Table 4. Dependent Variables ........................................................................................... 51 Table 5. Relational Coordination Measure ....................................................................... 58 Table 6. Trust Measure ..................................................................................................... 59 Table 7. PCC and PACIC ................................................................................................ 60 Table 8. Key PCP Domains .............................................................................................. 65 Table 9. Additional Patient Survey Variables................................................................... 67 Table 10. PCP Survey ....................................................................................................... 68 Table 11. PCP Survey Covariates ..................................................................................... 72 Table 12. Assessment of Chronic Illness Care (ACIC) .................................................... 74 Table 13. Patient Sample Descriptive Statistics .............................................................. 76 Table 14. Model I Dependent Variables .......................................................................... 78

xii

Table 15. PCP Univariate Statistics ................................................................................. 81 Table 16. Patient Descriptive Statistics ........................................................................... 83 Table 17. Level 3 Model I PCP Sample with Linked Patients ..................................... 85 Table 18. Level 2 Model I Patient Sample Size............................................................ 85 Table 19. Patient Survey Responders as Compared with Non-responders ...................... 87 Table 20. Patient Survey Linked Dependent Variables (Dichotomous Variables 0/1) 88 Table 21. Patient Survey - Patient Descriptive Statistics ................................................. 90 Table 22. PCP Descriptive Statistics - Patient Survey - Level Two ................................ 92 Table 23 Clinic Descriptive Statistics linked to Patient Survey Respondents ............ 94 Table 24. PCP Descriptive Statistics PCP Survey ........................................................ 95 Table 25. Bivariate Statistics: Diabetes Screens, and Diabetes Screens with A1c and LDL control and No Acute Utilization 2004 through 2007 (significant findings) ........... 98 Table 26. Bivariate Statistics: Diabetes Composite (A1c, LDL, CR screens) and No Acute Utilization 2004 through 2007 (significant findings) ........................................... 101 Table 27. Bivariate Statistics: Medication Adherence to Oral Diabetes and No Acute Utilization 2004 through 2007 and Adherence only (significant findings) .................... 104 Table 28. Bivariate Statistics: Medication Adherence to Ace Inhibitors and Arb Medications and No Acute Utilization 2004 through 2007 (significant findings) ......... 106 Table 29. Bivariate Statistics: Diabetes Screens and DM Screens with A1c control and No Acute Utilization 2004 through 2007 Group Model Only (significant findings) ... 108 Table 30. Bivariate Statistics: Diabetes Screens and DM Screens and No Acute Utilization Patient Survey Respondents (significant findings) ................................... 111 Table 31. Bivariate Statistics: Diabetes Screening Composite and Composite with No Acute Utilization Patient Survey Respondents (significant findings) ........................ 112

xiii

Table 32. Bivariate Statistics: Diabetes Screens with A1c control and No Acute Utilization - Patient Survey Respondents (significant findings)..................................... 113 Table 33. Bivariate Statistics: End of Life and Total PCP Satisfaction Patient Survey Respondents (significant findings) ................................................................................. 115 Table 34. PCP Sample with Linked Patients Diabetes A1c and LDL Screens HGLM Model .............................................................................................................................. 122 Table 35. Patient Sample Size Repeated Measure Diabetes A1c and LDL Screens HGLM Model ................................................................................................................. 123 Table 36. Diabetes Screen Multilevel Descriptive Statistics ......................................... 123 Table 37. Diabetes Screen HGLM Unconditional Model ............................................. 125 Table 38. Diabetes Screen HGLM Level-One ............................................................... 127 Table 39. Diabetes Screen HGLM Level-One and Level-Two ..................................... 130 Table 40. Diabetes Screen HGLM Final Model ............................................................ 135 Table 41. PCP Sample with Linked Patients Diabetes A1c and LDL Control & No Acute Utilization HGLM Model ..................................................................................... 138 Table 42. Patient Sample Size Repeated Measure Diabetes A1c and LDL Control & No Acute Utilization HGLM Model ..................................................................................... 138 Table 43. Diabetes Screen & Control Multilevel Descriptive Statistics ........................ 139 Table 44. HGLM: DM Screen & Control ...................................................................... 141 Table 45. PCP Sample with Linked Patients Diabetes Composite HGLM Model ....... 143 Table 46. Patient Sample Size Repeated Measure Diabetes Composite HGLM Model 143 Table 47. Diabetes Screening Composite Multilevel Descriptive Statistics.................. 144 Table 48. HGLM: Diabetes Screening Composite ........................................................ 146

xiv

Table 49. Level 3 Group Practice Diabetes Screen HGLM PCP Sample with Linked Patients ............................................................................................................................ 151 Table 50. Level 2 Group Practice Diabetes Screen HGLM - Patient Sample Size .... 151 Table 51. Group Practice Diabetes Screen Multilevel Descriptive Statistics ................ 152 Table 52. HGLM: Group Practice Diabetes Screen....................................................... 155 Table 53. Patient Survey Final Sample Nested within Primary Care Physicians (PCPs) ......................................................................................................................................... 158 Table 54. Patient Survey Final Sample PCPs Nested within Sites/Clinics.................. 158

Table of Figures

Figure 1. Conceptual Model for Longitudinal Analyses .................................................. 33 Figure 2. Average Access, Communication, and Coordination Scores 2004 to 2007 ..... 42 Figure 3. Conceptual Model for Cross-sectional Analyses .............................................. 57 Figure 4. Overall Mean Results for Dependent Variables Over Time ............................ 79 Figure 5. Predicted DMScreens as function of time duration ......................................... 122

xv

Introduction
Diffusion of the Chronic Care Model (CCM) as a prominent strategy for redesigning the U.S. health care system promoted by professional associations, conference presentations on pilot programs and the large national pilot program funded by the Robert Wood Johnson foundation and the Center For Health Systems Change has been rapid over the past two decades. The ultimate goal of CCM diffusion was the reorganization of provider organizations that would function as bridges linking the various fragmented parts of our system, multiple care settings and providers, to improve clinical patient outcomes. The key mechanism for these desired results was productive interactions between informed, activated patients and prepared, proactive practice teams led by primary care physicians. This key operational mechanism lacked explication within the model simply described as an expectant result of the CCM components.

A great deal of health services research has found continuity of care with a primary care physician plays an important role. Medical sociologists have studied the physicianpatient relationship for decades through the evolution of the health care system from one largely treating acute care illness to one consumed greatly today by treatment for chronic conditions. This shift in emphasis necessitates ongoing, negotiated, needs-based, goaloriented treatment over time that minimizes acute exacerbations rather than producing a cure. Long-term relationships with primary physicians increase the opportunities for the exchange of critical information between physicians and patients. Given the complexity of illness found among the elderly such exchanges assist physicians with assessing

patients medical conditions, needs and treatment preferences and nurture mutual trust. Trust is essential to the development of physician-patient partnerships needed to optimize outcomes in chronic illness care.

This dissertation research was conducted in an integrated multi-specialty group practice organization which embraced the CCM philosophy of practice redesign and committed the financial and personnel resources to establish CCM components. The organization has implemented an Electronic Medical Record system with e-prescribing and registry list capacity, health education classes, dietician support, care management services, and team meetings to coordinate care for high-risk patients. While extensive senior leadership exists for the support of CCM components, variation in quality measures exists throughout their organization affiliates. This may be the result of physician adoption and consistent application of supportive CCM tools or it may be the result of differences in patient-centered care.

One objective of this study was to determine if clinic/site-specific implementation of CCM components would explain any of the variation observed in quality outcomes among its Medicare patients with diabetes. Would any specific CCM component or the CCM composite predict any of the clinical outcomes being examined?

Moreover, at the physician level of the organization would relational coordination and trust, explain any of the variation observed? Every patient in the organization is assigned a primary care physician (PCP) accountable for the assessment of health care needs and

the subsequent provision and coordination of needed health care services. Additionally, the organization has provided Spanish language classes for the physicians to facilitate communication among the many Spanish-speaking patients served.

The second objective of this study was to determine if relational coordination and trust in a PCP predicts any of the clinical outcomes being examined? Would relational coordination and trust provide measurable constructs for the productive interactions envisioned in the CCM model as a result of implementing the recommended health care system redesigned components?

Chapter 1: Background

Study Rationale
A key recommendation in the IOM report Crossing the Quality Chasm was an appeal to redesign the health care delivery system with an emphasis on improving the coordination of care in our fragmented system [1]. Care coordination is often discussed by policymakers as the next generation or extension of care management strategies aimed at reducing the ever escalating cost of health care services, reducing system fragmentation, and improving quality of care and improving consumer quality of life [2]. Nowhere are these goals as critical as they are in the delivery of care to the growing number of Medicare beneficiaries; more than half of this group have two or more chronic conditions, one-fourth have problems with mental function or cognitive impairments and more than one quarter report their health status as fair or poor [3]. The U.S. healthcare system design-fragmented, highly specialized, acute care and short-term treatment focused, impersonal, without financial reimbursement for care coordination-is a poor match for the beneficiaries just described [4]. Care coordination mechanisms are needed across the various outpatient service units where care is provided such as, primary care physician offices, specialist physician offices, diagnostic centers and laboratories, and physical/occupational therapy, and behavioral health provider offices. According to a survey conducted by the Robert Wood Johnson Foundation of 6000 Americans living with at least one of six prevalent chronic illnesses, 45 percent reported receiving no help from their doctor or their health plan in

coordinating their medical services. Only about half reported involvement with their doctor in making decisions about their care and were capable of internalizing the confidence to manage their illness between visits [5]. Additional coordination is needed during acute episodes of care involving transitions from hospitals to skilled nursing facilities to homes [6-9]. The Institute of Medicine and the National Priorities Partnership have identified care coordination as a national priority to improve the quality of care. A recent study of patients with complex chronic illness in six countries found the need for system innovations to improve outcomes for patients with complex chronic conditions [10]. Care coordination is critical for all persons but more so for vulnerable groups such as senior patients most likely to be challenged by our fragmented healthcare delivery system. Almost 80 percent of seniors have at least one chronic illness, visit multiple physicians within multiple care settings and take multiple prescriptions with little coordination of care [11]. Inadequate care coordination for seniors living with multiple chronic illnesses can lead to unnecessary nursing home placements, inappropriate hospitalizations (ambulatory care sensitive conditions), redundant procedures/tests and/or adverse drug interactions [12]. Lack of care coordination may partially explain the poor quality care received by persons with chronic illness. According to the Crossing the Quality Chasm approximately 50 percent of persons with chronic illness were not

Care coordination has numerous definitions. For the purpose of this study the IOM definition of care coordination is being examined. According to the IOM, care coordination is a set of practitioner behaviors and information systems intended to bring together access to health services, patient test results, and other vital patient information including patient and family preferences and needs to improve quality of care.

receiving evidence-based chronic illness care and only 25 to 40 percent had their chronic conditions under good control [1].

Ideally coordinated, patient-centered or relationship-based care would facilitate receipt of evidence-based care for chronic illness, and incorporate appropriate geriatric assessments [13] of long term care preferences, poly-medicine, the home environment and risk of falls, nutrition, and depression and/or substance use. The primary care physician (PCP) is in a unique position to coordinate the care within our fragmented delivery system of largely specialized care. Care coordination has been identified specifically as one of three components defining primary care [14, 15]. Furthermore, a large majority of persons with chronic illness (including 90 percent with diabetes in the U.S.) currently receive the bulk of their care in primary care offices [16]. Hence, strengthening the primary care physician-patient relationship has the potential to improve coordination of care and subsequently patient care outcomes.

The Chronic Care Model (CCM) purports to optimally support the PCP in this function by defining six organizational elements: 1) self-management support, 2) decisionsupport, 3) delivery system design, 4) clinical information systems, 5) organizational leadership and commitment to chronic illness care, and 6) linkages to the community. Theoretically these six key elements facilitate effective physician-patient interactions and subsequently improved patient outcomes [17-21]. This conceptual framework developed at the MacColl Institute for Healthcare Innovation at Group Health Cooperative is based upon scientific evidence and a thorough examination of successful health industry

leaders chronic illness programs. However, CCM fails to make explicit the relational factors between primary care physicians and their patients which have the potential to impact patient outcomes.

At the micro-level of an organization relational coordination [22-26] and trust [27-29] may play critical roles in facilitating productive interactions between prepared physicians and activated patients envisioned by the CCM. Szasz and Hollender, medical sociologists defined a relationship as an abstraction embodying the activities of two interacting [persons] [30]. The Model of Mutual Participation was described as the importance of equality between physicians and patients as human beings. The patients lived experiences assist the physician in determining the best treatment strategies as treatment for chronic illness is largely carried out by the patient over a long period of time. Failure to comply with medical recommendations and delay in seeking care may be the result of poor experiences with the health care system in general and poor physician communication specifically. The pathway connecting the six organizational elements of the CCM to the improved physician-patient interactions is currently underdeveloped in the model. If relational coordination and trust strengthen the physician-patient relationship, then research including these dimensions may elucidate this pathway. Moreover, the relational factors explored in this study have the potential to reduce health care disparities as communication ineffectiveness has been demonstrated to be partially responsible for the racial inequalities observed [31]. Social categorization of patients could be mediated through physician communication and other interpersonal skills [32]. The IOM report described a pathway elucidating the critical link between patient and

physician communication and health outcomes. The IOM model employed a linear pathway whereby effective communication led to patient satisfaction and subsequently patient adherence and improved health outcomes [33]. Improved communication and the development of a trusting relationship between physicians and patients have the potential to overcome inherent class, race and power differentials commonly found in the medical encounter. Effective patient-physician relationships may be the appropriate solution to reducing health inequalities in the U.S. Positive patient-physician encounters may reduce racial inequalities in health outcomes by bridging knowledge and power differentials, and eliminating the negative feedback loop produced by ineffective and/or disrespectful medical encounters.

Relational Coordination (RC) [26, 34] consisting of timely, frequent and accurate communication; and shared knowledge, goals, and respect are critical components to an effective physician-patient relationship. RC and trust [35-37] are the essential ingredients to developing physician-patient partnerships capable of achieving optimal health outcomes for seniors living with multiple chronic diseases [38, 39]. While many models for physician-patient communication are elucidated [40], implementation requires a trusting, established relationship.

Effective communication is an essential component of relational coordination and trust, as well as effective teamwork and the evolution of partnerships. Effective communication alone however may be inadequate for producing the relationship between

the physician and patient required for the co-production of desired patient outcomes. Therefore, this study examines the constructs of relational coordination and trust.

Theoretical Framework
The Chronic Care Model, Relational Coordination Theory and trust in personal physician provide the primary theoretical frameworks for this research study. My hypotheses assert that relational coordination and trust constitute the mechanism for producing the desirable patient care outcomes presumed by the implementation of the Chronic Care Model (Appendix A, Theoretical framework). This study views the PCP-patient as the critical dyadic team in chronic illness care whereby the patient carries out the primary self-management activities in between physician office visits. Additionally, relational coordination and trust may mediate racial inequalities in patient outcomes and adherence.

The Chronic Care Model (CCM)


Dr. Ed Wagner from the MacColl Institute for Healthcare Innovation at the Group Health Cooperative of Puget Sound, Seattle developed an analytical framework or conceptual model, The Chronic Care Model designed to improve functional and clinical outcomes for patients living with chronic illness [17-21, 41]. The changing U.S. demographics intensify the need to determine the Chronic Care Model components most influential in facilitating the proactive physician-patient relationship and care coordination.

The Chronic Care Model (CCM) provides a conceptual framework to guide organizations in making a system change to improve chronic illness care [41]. CCM identifies six

organizational components: 1) self-management support, 2) decision-support, 3) delivery system design, 4) clinical information systems, 5) organization of health care, and 6) the community. The six elements provide a framework whereby care coordination is possible. Within this framework primary care physicians have access to evidence-based medicine, comprehensive knowledge of patient medical history including care provided across healthcare settings, and receive feedback reports on their patient panels as related to their peers and organization benchmarks. Patients and families obtain consistent, reliable information and are provided the resources needed to manage their chronic condition(s).

Patients are at the center of The Chronic Care Model as they must live with their chronic illnesses every day between visits to providers; patients and families are responsible for the integration of any treatment into their daily routines. Support for self-management is an essential element of CCM and the one most empirically tested of all the CCM elements. Patients with chronic illness need the skills, knowledge, and resources to optimally manage their conditions between physician visits. Patients must modify behaviors and often medication, and take appropriate action in accordance with recognized symptoms in order to achieve optimal outcomes.

Physician support ensures appropriate care is provided and missed opportunities are rare. Decision-support denotes the easy integration of explicit, evidence-based practice guidelines into the day-to-day practice of busy primary care providers. Ideally multiple physicians treating the same patient make treatment decisions based upon evidence-based

10

medicine. Delivery system design supports care coordination through clarification of staff roles and access to patient information in a central location accessible to all providers interacting with the patient. The implementation of electronic medical records in many organizations is an example of delivery system design. The creation of chronic care teams with increased information processing capacity for the medically complex patient is another example. Clinical information systems permit the creation of patient registries to assist in patient monitoring and outreach. Reminder systems and physician feedback are other benefits of clinical information systems [19].

The last two elements of The Chronic Care Model are overarching and meant to connect the other four components discussed above. The organization of the health care refers to an organizations commitment to managing chronic illness and can be strengthened through physician champions, support for innovation, and the genuine desire to provide patient-centered care. The community component acknowledges the importance of the larger sphere in which patients live, such as local policies promoting chronic illness care. Health care organizations can form alliances with local schools, churches, community centers, senior centers and assist them in the promotion of health, prevention of chronic illness and prevention of complications for those already living with chronic disease [42].

Relational Coordination
Relational Coordination Theory extends and builds upon the design theories of coordination within organizations. Organizational structure is designed in accordance with the task-specific information processing needs; uncertainty and interdependence

11

creates a greater need for information processing or communication across units [43]. Health care organizations face great uncertainty due to the changing market demands, the evolving complexity of medicine and the increasing longevity of greater numbers of individuals living with multimorbidity. Moreover, the resulting increase in medical specialization and advanced technology has further fragmented the health care system creating higher levels of interdependencies. The Chronic Care Models elements of delivery system design and clinical information systems would support Galbraiths organizational information processing theory by increasing the capacity of the organization to process information on complex patients.

Tushman and Nadlers information processing model extends Galbraiths by discussing the additional importance of connections between sub-units, networking. They described two different approaches for cross unit information processing mechanistic, such as guidelines and protocols for routine tasks, and organistic, such as team meetings when high information-processing capability with others is required to fulfill the task [44]. Design theory suggests higher levels of performance and efficiency if such coordinating mechanisms are used in these situations. The implementation of CCM incorporating electronic health records as part of its delivery system design would support such information processing needs.

Argote posits the additional need for autonomy to permit mutual adjustment by agents in accordance with general policies rather than strict rules [45]. These theories of lateral, high bandwidth information processing and the importance of autonomy are evident in

12

the CCM. Decision-support, delivery system design and clinical information systems all contribute to efficiencies in care by facilitating the communication among providers and ensuring access to patient information and optimal care coordination. Within the context of the IOMs definition of care coordination, physician autonomy is essential in treating the medically complex, non-routine patient with multi-morbidity and incorporating patient/family needs and preferences for care.

Gittell introduced an extension to these design theories of coordination and information processing by integrating the important people component through which other coordinating mechanisms must work. Relational coordination is a communication and relationship-intensive form of coordination that are expected to be particularly important for achieving high levels of performance in settings where tasks are highly interdependent and where levels of uncertainty and complexity are relatively high [2225]. Interdependence and uncertainty are high for the physician-patient dyadic team; effective physician action depends upon the comprehensive knowledge of the patient. The treatment plan is negotiated and tailored to meet individual patient preferences, goals and lifestyle only if optimal information exchange occurs between the physician and patient. Uncertainty is high given the medical complexity among the elderly, changing life situations and the enormous time constraints placed upon outpatient medical care. Interactive, ongoing communication in an atmosphere of shared knowledge, goals, and respect is critical to achieving improvements in patient outcomes and adherence as patients must make daily decisions related to their chronic conditions.

13

Relational coordination theory suggests the gap in other coordinating mechanisms is met by measuring seven, equally weighted dimensions; assess communication for 1) frequency, 2) timeliness, 3) accuracy, and 4) problem-solving capacity, and assess relationships by 5) shared goals, 6) shared knowledge and 7) mutual respect [22-26]. Gittell developed and validated an instrument for assessing relational coordination among team members and posits high RC improves efficiency and performance under conditions of high uncertainty and high task interdependencies [22-26]. Empirical evidence found that RC mediated coordinating mechanisms such as meetings and the use of boundary spanners such as case managers, indicating these coordinating mechanisms worked through their impact on relational coordination not independent from it [22-26].

While several studies have supported this hypothesis in work teams, the hypothesis of RC leading to improved patient outcomes was recently confirmed between providers and informal caregivers within the context of hospital discharge following a surgical procedure [46]. This study extends the hypothesis to the core critical health care team in the ambulatory medical setting-physician/patient. While several teams operate in our current health care system, the critical core team is the patient and primary physician. While ancillary health providers assist busy physicians in caring for medically complex patients, this work has been demonstrated to be most effective when it is endorsed and actively supported by the primary care physician.

14

Trust
Trust as viewed through the lens of social science may be described as a relationship of reliance. Trust has been defined as the willingness of a party to be vulnerable often based upon an underlying assumption of an implicit moral duty [28]. Trust has been identified as a necessary component to the development of physician-patient relationships [47-49] . When power and control are no longer the issue, we can form partnerships with our patients in which we learn together about the mystery we call life [50]. Trust leads to sharing information and reduces the traditional imbalance in power found between the physician and patient [29]. While trust is positively related to the length of the relationship, it is also separate from related measures such as patient satisfaction [29, 51].

Within organizational theory research two types of trust have emerged, affect-based and cognitive-based trust that were later described as benevolent and competent trust [28]. Trusting a knowledge source (doctor in this case) to be benevolent and competent should increase the chance that the knowledge receiver (patient) will learn from the action [28]. Organizational studies work has measured and examined the association between both types of trust and efficiency, quality and other aspects of production. Lewin and Cross tested a dyadic knowledge exchange model and found that useful knowledge to the knowledge seeker was gained under conditions of competence and benevolence-based trust [27]. Moreover, competence-based trust was especially important for the receipt of tacit knowledge. This might be viewed in the context of the physician-patient dyad in the following manner. If the patient judges the physician competent he or she is more likely to provide the physician the type of information that can maximize and personalize the

15

recommended treatment increasing the likelihood for treatment effectiveness and patient adherence. Patients' trust in their physician and physicians' comprehensive knowledge of their patients have been found to be associated with three important outcomes of care adherence to physicians advice, patients satisfaction with their primary care physician, and change in health status [52]. Trust in this study may be found to be subsumed within RC although not explicitly measured within its empirically tested instrument. A recent study conducted to develop and test an instrument to measure trust in health care providers among minority patients with chronic illness identified six characteristics they termed collaborative trust [51]. They are the following: 1) knowledge sharing, 2) emotional connection, 3) professional connection, 4) respect, 5) honesty, and 6) partnership. Following the development and testing of a fifteen item survey three distinct domains were identified as, interpersonal connection, respectful communication, and professional partnering. This work demonstrates the multi-dimensional aspects of patient trust in providers and results in domains similar to RC. Research Questions The main research questions examined in this study were the following: 1) Do any of the Chronic Care Model Components predict quality outcomes and adherence among senior patients with chronic illness: 2) Do physician relational coordination and trust predict quality outcomes independently and/or jointly?

16

Chapter 2: Literature Review


Over a decade ago a group of physicians and researchers at MacColl Institute for Healthcare Innovation recognized that a paradigm shift was needed within the healthcare system to address the efficient and effective treatment of chronic illness given its rapidly increasing prevalence and its associated costs and the current healthcare system designed for the treatment of acute illness [41]. Previous emphasis in chronic illness care had been placed on physician behavior (adherence to evidence-based clinical practice guidelines and medical training) and patient behavior (disease and case management programs). What was needed was a systems approach to reorganizing care that would facilitate the desired outcomes [41].

The resulting theoretical paradigm (the Chronic Care Model) reflected the components found upon studying the best 72 chronic illness programs in the country. CCM was piloted in quality improvement initiatives to determine if the organizational components could be integrated into busy practices and if integration was possible, did it make a difference?

A systematic review was conducted of studies of diabetes programs using the elements of CCM [53]. Built upon a recent Cochrane review of ambulatory care diabetes management programs, 39 studies were evaluated relative to four chronic care model components (self-management, decision support, delivery system design and clinical information systems). Thirty-two of the 39 studies found improvements in at least one process or outcome measure for patients with diabetes. Because of the small number of 17

studies, it could not be determined the relative impact of multiple components on outcomes. It was noted however, that 19 of 20 interventions that included a selfmanagement component improved a process or outcome of care measure [53].

The Robert Wood Johnson Foundation funded quality improvement collaboratives within 72 organizations nationally to implement the Chronic Care Model. The Institute for Healthcare Improvement using its Breakthrough Series quality improvement strategies provided technical support and training to these organizations for the implementation. Organizations chose to focus change efforts on patients with diabetes, heart disease, asthma and/or depression. RAND was contracted as the evaluator.

Before results of these collaboratives were known, an explosive diffusion of the Chronic Care Model was observed across the health care industry that was eager for a framework to address the rising costs from the exponential growth of chronic illness. As an example, the California Department of Health Services, Medi-Cal Managed Care Division adopted the CCM framework for the implementation of its state-wide quality improvement initiatives for asthma and diabetes. The diffusion is further evidenced by the 1,345 citations identified by a recent meta-analysis of the impact of CCM on patient outcomes [54]. The final meta-analysis contained 112 studies of CCM across four disease states (asthma, heart failure, diabetes and depression). Beneficial effects on clinical outcomes and process of care measures were found for all interventions that contained at least one or more of the CCM elements [54]. Delivery system design and self-management support were the two CCM components that contributed the greatest

18

impact on outcomes. These studies did not examine CCM as a whole and its influence on patient outcomes.

A large multispecialty medical group in the Midwest (representing 600 physicians, 300,000 patients and 17 clinics) adopted CCM in 2002 and studied the implementation of CCM and its subsequent impact on quality for patients with diabetes, heart disease and depression. While significant changes over two years were identified for three of the six CCM components, only the diabetes measures were associated with two of the CCM elements decision support and clinical information systems [20]. The same organization reported several challenges to CCM implementation. While all the clinics adopted the expanded, prepared practice team concept, differences across clinics were pronounced. Finding suitable meeting times for collaboration created a universal logistic challenge. And few clinics actively engaged the physicians [55].

The RAND evaluation on the randomized controlled Improving Chronic Illness Care Collaboratives revealed mixed results. While sites in the treatment arm averaged 30 change efforts to implement CCM, the depth or intensity of such changes varied tremendously across sites from ratings of 17 percent to 76 percent of the highest rating possible [56]. Change ratings did correlate with each of the six elements of CCM. The evaluation of the impact of these collaboratives on clinical outcomes has proved more challenging due to the diversity of interventions employed and the conditions emphasized. One study evaluating four collaboratives focusing on CHF found a

19

statistically significant improvement in the composite clinical measure among the collaborative organizations as compared to controls [57].

A recent meta-analysis of articles on CCM effectiveness published since 2000 concluded that evidence supports the CCM as an integrated framework for redesigning group practice to improve patient outcomes. This study examined only published articles that cited at least one of the five seminal articles on CCM as a reference, Additionally, the intervention had to operate within the ambulatory care practice, change how care is actually delivered in some way, and incorporate at least four components of the Chronic Care Model [58].

The Chronic Care Model was intended to implicitly function as a coordinating mechanism by virtue of redesigning the delivery system. While the framework identifies a systems approach to reorganizing the structural and some of the process components required to improve outcomes, the implicit collaborative management at the heart of CCM has not been adequately measured or been made explicit. In an earlier paper by the founders of CCM, collaborative management occurs when patients and care providers have shared goals, sustained working relationship, mutual understanding or roles and responsibilities and requisite skills for carrying out their roles [59]. This implicit coordinating mechanism has yet to be tested within the CCM framework.

With the rapid diffusion of CCM within the private sector and managed care organizations positive experiences with disease and case management programs,

20

government legislated demonstration projects aimed at improvements in chronic illness care. While many of these demonstration projects utilized CCM implementation, disease and case management programs, the primary care physician had been largely excluded as an essential partner. Medicares Coordinated Care Demonstration (MCCD) and the Medicare Health Support Program focused primarily on the patient, with encouragement by contractors to engage the primary care physician.

The final report to Congress on the outcomes of the fifteen programs for MCCD following the three years of implementation was not favorable [60]. Thirteen of fifteen programs failed to reduce hospitalizations; one program actually reported an increase in hospitalizations and one program reported fewer hospitalizations. None of the programs generated net savings. Programs targeted a single chronic condition (CHF, CAD or diabetes) and three used high-risk predictive algorithms for the selection of their respective target population. Patients were the focus of most programs and while all programs sent reports to primary care physicians, only 42 percent of physicians rated such reports as very useful. These programs had no impact on adherence measures and made improvements in only a few quality measures. The authors of the report concluded that programs lacking a strong transitional component across care settings were unlikely to yield cost-savings.

The second report to Congress on the Medicare Health Support program revealed that none of the pilot programs at 18 months into the three-year program period met the three requirements in the legislation improvements in clinical care quality, patient

21

satisfaction, and cost neutrality to CMS [61]. The Medicare Health Support program was the largest population-based disease management (DM) program ever legislated. Contracts were awarded to eight experienced disease management companies or organizations with DM infrastructure and support to manage CMS identified high-risk beneficiaries. Among the programs original patient population only 16 out of 40 evidence-based process-of-care measures improved. None of the programs achieved statistically significant reductions in hospitalizations, readmissions, or ED visits in their original populations as compared with control groups. The program was suspended.

One of the largest challenges faced by these two CMS initiatives to improve chronic illness care and reduce inappropriate expenditures was the lack of direct primary care physician involvement. Care management or disease management programs that operate distinct from primary care practice have the potential to further fragment medical care [4]. A major controversy with regard to DM programs is the extent to which services they provide are integrated with the patients other medical care [4]. Kane and others believe that a high degree of integration between these support programs and physicians is essential in establishing and maintaining effective physician-patient relationships. The lack of physician involvement may have also contributed to the relatively low participation rates for these programs as well as for the earlier CMS Case Management demonstration programs of the mid 1990s [62, 63].

Evaluation of care coordination on patient outcomes has been further hampered by the lack of a consistent definition and measurement. Researchers have developed, tested, and

22

validated care coordination questions as part of overall patient satisfaction surveys of ambulatory care [64-66]. Researchers with Improving Chronic Illness Care (ICIC) developed and validated The Primary Assessment of Chronic Illness Care (PACIC) to assess concordance of care with the major tenants of the Chronic Care Model that could be accurately evaluated by the patient [67]. Too little empirical research as been employed linking measurement of care coordination to patient outcomes [52]. Furthermore, selection bias has plagued disease management program evaluations for decades and more recently care coordination evaluation. CMS utilized prospective claims-based, predictive risk modeling to identify eligible patient cohorts for the demonstration projects. Such methods have been reported to be only 14 to 20 percent reliable in predicting high cost individuals in the coming year based upon prior claim history [68]. The Medicare beneficiaries recruited and participating in the CMS Care Coordination Demonstration projects were found to be healthier than the control group beneficiaries thereby resulting in accrued benefits to those who perhaps needed them least [63]. By contrast, identifying factors that strengthen the primary care physicianpatient team would impact all Medicare beneficiaries as even those in Fee-For-Service Medicare acknowledge a personal or primary physician.

Gap in the Literature


A broader constellation of interdependent physician and patient behaviors may constitute the essential yet untested factors important to improved patient outcomes for elders living with chronic illnesses. The traditional physician-patient dyad produces dominant

23

physicians and passive patients not conducive to the development of an effective team approach to care management and the co-production of optimal patient outcomes. High quality communication, shared knowledge and shared goals (components of relational coordination) have the potential to create a bridge through which effective teamwork and partnership can evolve. To achieve positive outcomes where the goal is not curative but rather slowing disease progression, the physician must partner with the patient by working together as an interdependent team.

Although the empirical evidence on the impact of CCM on patient outcomes has been favorable given interventions that incorporate at least one CCM component, the coordinating mechanism for the productive interactions between physicians and patients has not been measured or made explicit. This research attempted to address a gap in the literature by examining simultaneously the impact of the CCM components and the relational factors of the physician-patient dyad within a single umbrella organization, thus controlling for many potential confounders. Previous research has had difficulty assessing quantitatively the relative importance of the six organizational CCM components on patient outcomes [54]. This may be the result of the inadequate consideration and measurement of the relational components required for delivering effective care for chronically ill individuals may have been the reason.

The IOM report described an untested pathway elucidating the critical link between patient and physician communication and health outcomes [1]. The mediating factor in this linear pathway is patient satisfaction, more often a quality outcome in itself [69-71].

24

The goal of this study is to incorporate CCM components, and physician relational coordination and trust to identify the relative importance of each on patient outcomes and adherence. This study seeks to explore the extension of the organizational theories of relational coordination and trust as applied to the primary care physician-patient. This dyadic team is primarily responsible for the co-production of quality outcomes for senior patients living with chronic diseases.

25

Chapter 3: Research Methods

Introduction
Longitudinal and cross-sectional analyses were conducted to evaluate quantitatively the association between the CCM components and physician relational coordination and trust, and patient outcomes and adherence using data from a single, multispecialty medical group with an Independent Practice Association division as further described. The patient population was composed of managed care Medicare beneficiaries receiving care from this organization between 2004 and 2007. Longitudinal analyses utilized four years of medical claims and physician satisfaction data from the study organization, incorporating proxy variables (PCP communication and coordination) for relational coordination (RC) and trust. Cross-sectional analyses utilized survey data assessing CCM, RC and trust that was linked with respondents 2007 claims data. All longitudinal and cross-sectional analyses incorporated the same dependent variables, nationally recognized quality process measures, intermediate outcome and outcome measures, and patient medication adherence measures (Table 1.). Additionally, the 2007 cross-sectional analyses employed two outcome variables derived from the patient survey.

26

Table 1. Outcome Variables

Outcome A1c and LDL screen A1c and LDL control Colorectal screen ACSC* hospital admission 30-day all cause readmission ED visit for chronic condition Oral diabetic medication adherence Ace inhibitor/ARB medication adherence Overall PCP satisfaction Discuss EOL preferences with PCP

Quality Measure (Type/endorsement) Process (HEDIS) Intermediate outcome (HEDIS) Process (HEDIS) Outcome (AHRQ) Outcome (AHRQ) Outcome (none) Intermediate outcome (NQF) Intermediate outcome (NQF) Intermediate outcome (NCQA) Process (none)

Model
Longitudinal & Cross-sectional Longitudinal & Cross-sectional Longitudinal & Cross-sectional Longitudinal & Cross-sectional Longitudinal Longitudinal & Cross-sectional Longitudinal & Cross-sectional Longitudinal & Cross-sectional Cross-sectional Cross-sectional

* Ambulatory Care Sensitive Condition One of five chronic illnesses-DM, CHF, COPD, asthma, and HTN

Hierarchical Linear Modeling (HLM) techniques were employed to permit the formulation of explicit structural models for processes that occur within each level of a hierarchy (patient, PCP and clinic). A major goal of this study was to determine the relative influence of the CCM components and physician relational factors on patient outcomes within a natural occurring, hierarchical nested health care structure [72-79]. This statistical approach is discussed further in the analysis plan. All analytic models represented exploratory research, which is appropriate when attempting to apply theory in a new way [80]. Although the RC index has been empirically tested and validated it has not been used to assess the physician-patient dyad.

This chapter is organized as follows. First, the study site, study population, and the research questions related to the longitudinal and cross-sectional analyses are presented.

27

Second, the analysis plan, conceptual model, and the variable construction for key predictors, dependent variables and covariates, and the patient sample for the longitudinal analyses are described. Third, the primary data collection for the cross-sectional analyses is explained including the development and implementation of the patient survey instrument and the subsequent variable construction. Lastly, the clinic instrument used to assess the Chronic Care Model components and related CCM variables are presented.

Research site and population


Study Site A large multispecialty medical group with an Independent Practice Association (IPA) division in southern California served as the study site for this research. This organization embraces the Chronic Care Model (CCM) in its operations and management philosophy and possesses many of the organizational attributes desirable for achieving optimal patient outcomes such as physician leadership and autonomy [81], and the appropriate alignment of physician financial incentives [82, 83]. Moreover, the organization is structurally more representative of physician practice organizations than the frequently studied Kaiser Permanente Model. It is also highly rated by consumers and by independent organizations such as the CA Health Foundation among IPAs in the CA market.

This organization employs many of the CCM components including electronic medical records with e-prescribing capacity, evidence-based clinical guidelines, physician-

28

feedback reports, and integrated care management and disease management programs. Despite these characteristics and practice improvements, variation was identified in quality measures and utilization across physicians, medical group clinics and IPA sites.

Study Population The study population selected were senior managed care patients (65 years of age and older) receiving medical care by providers within the organization between 2004 and 2007 with evidence of diabetes and at least one additional chronic illness. Diabetes was selected because of its high prevalence and cost, and the existence of standardized, wellestablished quality measures for patient outcomes. Multi-morbidity was an additional criterion because of its increased prevalence among older adults and its potential impact on quality and adherence. A need for more research explicitly on multi-morbidity has been identified as much of the published research to date has focused on single chronic illnesses despite the increased prevalence of multiple chronic diseases [84, 85].

The site organization identified all patients identified with diabetes and a second chronic condition. The composite list of co-morbid chronic diseases included the following: chronic hepatitis, alcohol cirrhosis, chronic renal failure/ESRD, chronic pancreatitis, Arthritis (Rheumatoid and Osteoarthritis), hypertension, atrial fibrillation, CVD/stroke/TIA, CHF, Multiple Sclerosis, Parkinsons disease, COPD/emphysema, Asthma, Ischemic Heart Disease/CIHD/CAD/PVD, stable angina, back/lumbar pain, Alzheimers disease, schizophrenia, anxiety disorder and major depressive/bipolar disorder. These chronic conditions have been previously used to quantify the burden of

29

illness from claim data [86-88]. Identified patients with a diagnosis of active cancer were excluded as attainment of selected quality and adherence measures may not be appropriate goals for a patient in treatment for active cancer.

The identified patient population for the study was also further divided into those patients receiving their care or the majority of their care within the group model clinics. This subpopulation is referred to subsequently as the group practice population. The Chronic Care Model components are known only for this sub-population.

Data provided for analyses on all identified patients included the following: 1) monthmonth enrollment data with demographic information and PCP and clinic/site assignments; 2) outpatient encounter data; 3) HEDIS-eligibility flags for diabetes measures and the colorectal screening measure; 4) Hospital and Emergency Department claims; and 5) pharmacy claims. As the organization assumes full risk contracts (financially responsible for all medical services provided) for its managed Medicare patients the claims extracts done for this study provided comprehensive information on all identified patients. The notable exception was the non-reporting of pharmacy claims by four small health plans with fewer than 10 percent of the organizations managed care patients.

This study was approved by the Brandeis University IRB and conducted under an executed Limited Data Use Agreement.

30

Research Questions:
1) Do self-management resources and electronic medical records (two components of the Chronic Care Model) predict quality outcomes and patient adherence for seniors with multiple chronic diseases? (Four-Year Longitudinal Analyses; Group practice population) 2) Do physician communication and care coordination behavior (imperfect proxy variables for the variables of interest, RC and trust) predict patient outcomes and adherence? (Four-Year Longitudinal Analyses) 3) Does the CCM predict quality outcomes and patient adherence for seniors with multiple chronic diseases? (2007 Cross-sectional Analyses) 4) Does relational coordination and trust within the physician-patient dyadic team predict patient outcomes and adherence? (2007 Cross-sectional Analyses)

Secondary Research Questions: 2007 Cross-sectional Analyses 1) Is there a correlation between the proxy variables used in the longitudinal analyses and the specifically measured RC and trust in the cross-sectional analyses and physician? 2) Is trust subsumed within the relational coordination construct or does it measure a distinct characteristic of the physician-patient relationship, distinct from RC? 3) Are differences observed in RC and trust among patients of different ethnicity/race and for those who report non-English as their preferred language?

31

Four-Year Longitudinal Analyses

Analytical Plan
Dependent variables (measured annually at the patient level) were regressed on timevarying patient covariates at level-one, time-invariant patient covariates at level-two, and physician covariates and key physician predictors at level-three. Dependent variables were constructed from claims data, years 2004 through 2007. Key predictors in the longitudinal analyses included patient continuity with the PCP at level-two; and the relational domains of physician communication, coordination and access at level-three. The subset of longitudinal models based solely upon the group assigned patients (group practice population) included two key CCM components (presence of health education resources and electronic medical records) as fixed effects at level-three. The longitudinal analyses empirically examined the IOM purported pathway to improved patient outcomes via effective physician communication. Additionally the longitudinal analyses employed physician coordination, communication and access as proxy variables for relational coordination assessed subsequently in the cross-sectional analyses. The conceptual model is depicted in Figure 1.

32

Figure 1. Conceptual Model for Longitudinal Analyses

Physician

Coordination
Communication & Access (+ covariates)

CCMGroup-only:

HE & EMR

Patient continuity with PCP


(+ covariates)

Physician-Patient Dyadic Team

Patient Quality Outcomes and Adherence

PCP visits/Yr SP Visits/Yr

Same PCP/Yr

The choice for analytic methods for clustered data can have major implications for medical practice-based, health services research [72-74]. In health care patients treated by a particular physician receive care in a common treatment setting that is influenced by physician characteristics and philosophy, and that may differ from once physician to another [72, 79]. Likewise these physicians will be influenced by the practice setting in which they work and this setting is likely to differ across settings. Ignoring group membership can lead to erroneous conclusions (inflated parameter estimates) regarding the impacts of influence on outcomes [72, 79]. As a major goal of this study was to determine the relative influence of the CCM components and PCP relational factors on

33

patient outcomes as related to these likely correlated groups (patients, physicians and clinics), HLM was the appropriate analysis method.

Wong and Mason (1985) and Longford (1993), Goldstein (1991, 2003), Snijders and Bosker (1999) and Raudenbush and Byrk (2002), have described the multilevel extension of generalized linear models [89-92]. The transformed multiple regression equation of the linear predictor defines a two-level multilevel structure as follows: nij = Y00 + Y10Xij + Y01Zi + Y11ZjXij + ujXij + u0j. For this model, nij represents the log-odds of the conditional probability of the outcome being measured for individual i nested within organization j. In a repeated measure model i would denote time period i within individual j. Y00 represents the overall intercept value or the average log odds conditional on all predictors and the level-two random error term. Y10Xij denotes the slope of X for unit j. Y01Zi is a level-two predictor variable and Y11ZjXij is a crosslevel interaction term. UjXij is the random effect of X for unit j. In a generalized linear model the variance of the level-one error term is a function of the population proportion and is not estimated separately. The value of approximately 2/3 or 3.29 specifies the variance in a logistic distribution [91]. The logit link function transforms n into the logit(n).

Using Hierarchical Generalized Linear Modeling all longitudinal analyses predict the probably of an event across years as a function of time, covariates and key predictors, and random effects. The logit link transforms each linear structural model into the log of the odds of success (represented by nij) or a positive event [90, 91, 93]. Predicted log-odds

34

may be converted into odds by taking the exp(nij) or into a predicted probability by estimating ij = 1/1 + exp(-nij).

HGLM analysis allows the inclusion of all the patients meeting study criteria regardless of the number of measurement periods as unbalanced panel data can be modeled successfully. HGLM estimation methods enable unbiased estimates of the Beta coefficients in each unit as it borrows strength from the fact that the estimation of Beta is being repeated across a number of units [93]. HGLM allows the cluster variance to be modeled in addition to correcting the standard errors for fixed effects parameters as compared with other approaches such as Generalized Estimating Equations (GEE). GEE analysis corrects the standard errors of the fixed effects parameter estimates ensuring proper interpretation of parameter significance however the cluster variation is treated as a nuisance [91].

Additionally, as previously discussed HGLM permits predictors at each level of the hierarchical data structure (e.g. time-varying predictors across years, individual patient predictors and physician predictors). HGLM allows for estimation of covariance structures between levels and estimates predictors from one level that potentially explain variance at another lower level. Most importantly, HGLM estimates the reduction in variation at the highest level of the data structure that is explained by the key predictors.

HGLM assumptions include the following. The transformed function forms are linear at each level. Residuals at level two and three are assumed to be normally distributed and

35

not correlated with level two or three predictors. Observations at the highest level are independent of each other.

SAS 9.2 software was used for all variable construction, and for the examination of descriptive statistics and bivariate relationships. HLM 6.06 software (SSI Scientific Software International) designed specifically for hierarchical linear modeling was used to conduct all final analyses, and produce fixed effects parameter and variance reduction estimates. Users may select the option for estimation of the hierarchical models for a binary dependent variable, Laplace transformation or penalized quasi-likelihood (PQL) techniques (full or restricted). Laplace transformation technique involves the maximization of approximations to maximum likelihood (ML) parameter estimates, cluster variation estimates, and produces the associated model deviance statistics. PQL technique (the default estimation procedure for non-linear models in HLM 6.06) is based on approximations of the joint posterior densities of the model. PQL produces empirical Bayes, generalized least squares estimates for the random coefficients, fixed effects and variance components but does not generate a reliable deviance statistic that can be used to compare competing models.

Laplace estimation was used for the majority of analyses; research suggests the estimation provides better performance in terms of bias especially as related to the estimation of the variance components [94]. Penalized quasi-likelihood (PQL) estimation was used in cases of model non-convergence and in cross-sectional analyses with fewer level two units to produce unit-level fixed effects parameter and variation reduction

36

estimates. Fixed effect estimates and variance components are negatively biased with PQL techniques if the level-two variances and/or level-three variances are very large and if the targeted probability is very small or very large [91, 93, 94]. Estimation through the Laplace transformation overcomes this bias and produces a reliable deviance statistic for binary response models however lack of model convergence may be a problem for some data sets. For analyses using the Laplace estimation method, goodness-of-fit for a given model was assessed by the comparison of deviance statistics between nested binary models. HLM 6.06 software incorporates the use of Fisher scoring to estimate maximum-likelihood estimates (MLE) using Laplace transformation in three-level models and Expectation Maximization (EM) to estimate MLE using Laplace transformation in two-level models.

All models were constructed using a sequential model-building approach described by Raudenbush as: 1) specify the unconditional model, the probability of an event given the level one, two, and three units of analysis without any predictors (and with the inclusion of time if appropriate); 2) specify the compositional model by adding covariates at level one, two and three and 3) specify the contextual model by adding key predictors to the compositional model ([90] .

The intra-class correlation coefficient ([ICC] is the estimate of the between-cluster variance in the model. The total variation in patient outcomes at each level may be partitioned into two variance components, within-level and between-level variance. If there is no correlation among observations at a given level the ICC is equal to zero

37

whereas if all the observations within a level were identical the ICC is equal to one. The ICC is calculated on the unconditional or empty model based upon the between-level variance estimate divided by the total variance estimate.

The ICC calculation is more complicated in the HGLM with the logit link as the level-1 residual is constrained by the predicted probability. Snijders [91] suggested formula for calculating the ICC in a HGLM with a logit link was used; it incorporates the implied variance for the level-one residual in a logistic distribution (2/3 = 3.29). The equation for the ICC becomes U00(level-three variance estimate)/U00 + R0 (level-two variance estimate) + 3.29. For each analytic model presented, significant fixed effects parameter estimates and the reduction in between-unit variation (derived via U00 empty modelU00restricted model/U00empty model) were reported.

Longitudinal Analysis - Key Predictor Variables


Proxy variables were constructed for the key variables of interest (RC and trust) using the quarterly physician satisfaction survey data collected by the organization across this time period (2004 to 2007). Additionally, two components of the Chronic Care Model (CCM), presence of on-site health education or care management resources, and days of electronic health records were examined in the sub-models for the group practice population.

38

A key construct of relational coordination and trust is communication [25, 26, 29, 31, 35, 39, 95-98]. Effective physician communication must be preceded by access patient must be able to access the office, the medical practice, etc. Coordination is the hypothesized latent and currently unmeasured component within the CCM that impacts positively on patient outcomes [12, 14, 15, 23-26, 46, 60, 64, 69, 99-107]. The organization mailed surveys quarterly to all managed care patients (on a rotating basis) that were designed to assess the patients PCP and aspects of care of care using the fivepoint rating scale of excellent, very good, good, fair and poor. Specific survey questions linked to hypothesized domains are presented in Table 2.

Surveys from patients younger than fifty years of age and physicians with fewer than ten total surveys were removed from the database containing raw scores for survey questions for each PCP by quarter and year. Restricting the analysis of physician satisfaction surveys to patients at least 50 years of age was done to better reflect the patient population for this study. The perception of physician behavior by younger patients and the medical needs of younger patients may differ from those of older patients [108, 109]. The threshold of a minimum number of surveys (10 surveys) was used to improve measure stability in the resulting domain scores.

39

Table 2. Organization's Physician Satisfaction Survey Physician Satisfaction Survey Data Rate your Primary Care Physician on. Hypothesized Variable Survey Question Domain Name Access CAREND Access to medical care from the DOCTOR/CLINICIAN when needed. Access AFTRHRS Access to medical care when needed after regular business hours or on weekends.. Access Access PHNCNT DRSPK Ease of contacting the medical group by phone

Ability to speak with the DOCTOR/CLINICIAN or nurse on the phone. The number of days you waited for your appointment How well DOCTOR/CLINICIAN explained what he/she was doing and why How well DOCTOR/CLINICIAN used words that were easy to understand. How well DOCTOR/CLINICIAN listened to your concerns or questions. How well DOCTOR/CLINICIAN answered your questions.. Warmth and caring demonstrated to you by the DOCTOR/CLINICIAN. The courtesy and supportiveness shown to you by the nurse(s) and/or medical assistants. The way you were informed about the results of lab or x-ray tests.. The ease in having prescription renewals reauthorized by the DOCTOR/CLINICIAN The amount of time you had with the DOCTOR/CLINICIAN Ability to get a referral when you felt it was needed.

Access Communication Communication Communication Communication Communication Coordination Coordination Coordination Coordination Coordination

DAYSWT DREXPL DRWORD DRLSTN DRANSR DRWARM NURCOR INFLAB RX VSTIM REFND

The mean number of surveys per PCP was 271 (median, 252). The majority of survey respondents over the time period were female (60%) and 4.8% of surveys were completed in the Spanish language. Only 11 physicians had fewer than 30 surveys. The mean number of surveys among PCPs linked to the group practice models is 315 with a

40

range of 27 to 910 surveys. A total of 84.6% of PCPs had at least 12 quarters of survey data used for the calculation of domain scores.

Principal components analysis using the average raw scores and the polychoric correlation matrix from 52,599 surveys was conducted without differences in factor loading results from the two approaches. As survey response data consist of rating scales and cannot be considered interval data, some have argued that analysis of such data should be conducted on the data transformed polychoric correlation or Spearman correlation matrices. Polychoric and Spearman correlation matrices consider the ordinal nature of the data and the lack of equal distances between each response level [110, 111]. When a three-component solution was not specified, PCA produced a two-component solution with communication as a single component, and access and coordination combined as a component. However, to maximize distinct uncorrelated components and assist interpretation with a three-component solution (hypothesized to have distinct construct validity) orthogonal transformation was done. [112]. PCA with orthogonal transformation resulted in factor loadings above 0.40 on all items indicating item convergence and item discrimination (communication variables range from 0.75 to 0.87, access variables range from 0.68 to 0.79, and coordination variables range from 0.61 to 0.66). Internal reliability for the three components was assessed by calculating the alpha Cronbach coefficients for each domain (Table 3). Standardized Cronbach coefficient alphas exceeded the recommended minimum threshold of 0.70 [112-114].

41

Table 3. PCP Domains


Physician Domain Standardized Cronbach Coefficient Alpha

Access Communication Coordination

0.903 0.969 0.843

PCP domain variables of access, communication, and coordination were constructed by an iterative process. The mean of the item responses per survey respondent for each domain (excluding surveys with more than one item missing per domain) was calculated initially. Subsequently, the overall mean of each domain for all survey respondents for a given PCP for each calendar year was calculated. Domain scores for each PCP were then standardized using z-score transformation with a mean of zero and a standard deviation of one for each year of survey data as the distribution and variance of the raw domain scores differed across years with the final two years trending upward (Figure 2.). Additionally, the PCP peer-group was not consistent across the years.
Figure 2. Average Access, Communication, and Coordination Scores 2004 to 2007
PCP Domain Averages Raw Scores
4.60 4.50 4.40 4.30 4.20 4.10 4.00 3.90 3.80 3.70 3.60 2004 2005
Year

ACS CM CD

2006

2007

42

The best estimate for the overall physician domain scores over the period of study was determined to be the predicted value based upon yearly z-transformed domain scores and random error (Appendices B & C). Hierarchical Linear Modeling was used to regress the z-transformed annual PCP scores on the PCP cluster and random error to produce a dataset of predicted PCP domain values for access, communication, and coordination. (As an example the equation for the coordination domain was zCDti = B00 + r0i + eti.) Dichotomous variables for each PCP domain were created by assigning a one for a score above zero and a zero otherwise. (Although all annual PCP domain scores were standardized to a mean of zero and a standard deviation of one the predicted PCP domain scores from HLM resulted in less than 50% of PCPs having scores above zero as a result of accounting for the random variance at the PCP level - 46.9% of PCPs have overall access and communication scores of greater than zero and 48.6% of PCPs have an overall coordination score of greater than zero.)

As high correlation was found among the three PCP domain scores (Appendix D), additional variables were created to capture the influence of these key predictors as follows: zPDCMCD as the sum of the z-transformed predicted scores for communication and coordination and HPDCMCD as a dichotomous variable with one indicating above average predicted score on communication and coordination and zero otherwise. The standardized Cronbach coefficient alpha was 0.8895 for predicted access, predicted communication and predicted coordination and with the deletion of the predicted access scores the alpha coefficient increased to 0.9081. Moreover, communication and

43

coordination constitute the key behaviors of interest and those exclusively under the control of the PCP.

Two variables were constructed to represent two key components of the Chronic Care Model electronic medical records (EMR) and the presence of on-site care management and disease management resources. As HLM 6.06 software does not accommodate a fourth level data structure these two key variables were modeled as fixed effects at level three, the PCP level. The EMR variable represented the total number of EMR days available to impact clinic operations through the end of the study period (12/31/2007 go-live EMR implementation date). As this variable EHRDays had a normal distribution (Kolmogorov-Smirnov test statistic = 0.14961393, p > 0.105) it was created as a continuous predictor variable. The presence of care management or disease management services on-site at a clinic location was created as a dichotomous variable with one indicating presence of resource and zero otherwise. These CCM components were examined in the group practice longitudinal analyses only as information for these variables was known only for the group practice population.

Dependent Variables Selected Patient Quality and Adherence Measures


The dependent variables included recognized measurements of quality and adherence among patients with diabetes and other morbidities. The process and intermediate outcome variables for diabetes are nationally recognized by the National Quality Forum, the Ambulatory Quality Alliance, and the National Committee for Quality Assurance as important in determining quality care for the optimal management of diabetes and have

44

been a long-standing component of the HEDIS measurement set for all populations. As persons with diabetes are at an increased risk for colon cancer [115-117], the receipt of colon cancer screening was added as part of a diabetes composite screening measure.

Many researchers examine hospital admissions and ED visits for ambulatory care sensitive conditions as quality measures for inadequate primary care access or services [118, 119] and readmissions [120, 121] as either a failure of care within the hospital, failure in the transfer of patient to another setting or failure in physician follow-up in the ambulatory care setting. Readmissions are common among senior patients with multiple chronic conditions and increase cost to the Medicare program [122-125]; recently NQF has endorsed 30-day all cause hospital readmission measures. Ideally optimal patientcentered, coordinated care can avoid ambulatory care sensitive admissions and reduce readmissions. Within a managed care environment a common goal is to provide the right amount of care at the appropriate time and in the right care setting. If the desired levels of disease control are to be achieved in chronic illness care and especially in persons with diabetes and the often accompanying cardiovascular disease patients must also adhere to treatment regimes [126-128]. The Medication Possession Ratio (MPR) for oral diabetic medications and the MPR for ACE inhibitor/ARB medications were calculated. MPR is the preferred measure given pharmacy claim data [127, 129, 130] and is now an endorsed medication adherence measure by NQF (http://www.qualityforum.org/Measures List.aspx).

45

Dependent Variable Construction


Eligibility flags following HEDIS (The Healthcare Effectiveness Data and Information Set) specifications[131] determined the eligible population for modeling A1c and lowdensity lipoprotein (LDL) screens and control, and colorectal screens annually. HEDIS requires 12 months of continuous eligibility with no more than one gap up to 45 days for A1c and LDL screens, and 24 months of continuous eligibility for colorectal screens using the measurement year and previous year, with the same gap allowance.

Adherence to the annual diabetes screen measure was assessed by evidence of at least one A1c and LDL screening test per patient per eligible year, without applying the HEDIS age ceiling of 75 years and the HEDIS two-year allowance for LDL screening for years 2004, 2005, and 2006. The optimal level for A1c was not specified by HEDIS until 2007 when the measure for good A1c control was introduced. A1c control in years 2004 through 2006 was defined as less than 8% for years 2004 through 2006 consistent with some recommendations for elder populations [132] and less than 7% in 2007 (consistent with the newly introduced A1c measure for good control). The HEDIS specified measure during the period from 2004 through 2006 was for poor control defined as an A1c result of above 9%. LDL control was assessed in accordance with HEDIS specifications as less than 130 mg/dL for 2004, 2005, and 2006, and less than 100 mg/dL for year 2007. LDL control was introduced as a new measure in the HEDIS diabetes measure set beginning in 2004 [131]. In accordance with HEDIS specifications the control measure was determined by the value of the lab result for the test taken last during a given year.

46

Evidence of a colorectal screen (CR) consisted of a claim for any of the following: a fecal occult blood test (FOBT) during the measurement year, a flexible sigmoidoscopy during the measurement year or the four years prior to the measurement year, a double contrast barium enema (DCBE) during the measurement year or the four years prior to the measurement year or a colonoscopy during the measurement year or the nine years prior. As the colorectal screen measure requires a significant look-back period annual rates for this measure were likely underestimated for years prior to 2007. As many physicians caring for senior patients do not recommend CR screens for patients older than 80 years of age, the age ceiling was applied for this measure. All control and screening variables were coded as dichotomous variables with one indicating control or evidence of a screen and zero otherwise. Additionally, a screening composite measure was created as a dichotomous variable indicating the evidence of all three screenings A1c, LDL and CR in a given year.

A combination of McCalls 22 conditions and the AHRQ Preventive Quality Indicator (PQI) conditions (http://www.qualityindicators.ahrq.gov) were used to define and identify ambulatory care sensitive condition (ACSC) hospitalizations from hospital claim data. McCall, principal investigator on a CMS project to identify trends in ACSC admissions among Medicare beneficiaries, convened a panel of experts to identify ambulatory sensitive conditions specifically for a senior population [133, 134]. The AHRQ PQI measures were intended to address all population groups. The combined list of potentially preventable hospitalization included the following discharge diagnoses:

47

amputation-lower extremity, angina, asthma, cellulites, CHF, COPD, dehydration, diabetes, grand mal/epileptic convulsions, gastroenteritis, hypertension, hypoglycemia, hypokalemia, immunizable conditions, influenza, invasive cervical cancer, iron deficiency anemia, kidney/urinary infection, malnutrition, perforated/bleeding ulcer, pneumonia, PVD with diabetes, ruptured appendix, severe ear/nose/throat infections and tuberculosis. All conditions were observed in the hospital claims of the study population except immunizable conditions. The total number of ACSC hospitalizations/year, and a dichotomous variable indicating any ACSC/year for each patient were coded. Using hospital admitting and discharge dates readmission variables were created indicating a hospital admission for any cause within 30 days and 90 days of a previous admission, excluding a subsequent admission within 24 hours of a discharge date. Total count variables and dichotomous variables (one indicating any 30-day or 90-day readmission) were coded for each year.

As AHRQ does not endorse the use of its PQI measures for ED utilization, ED visits with the principal diagnosis of five chronic conditions were coded as ED visit for chronic illness consistent with other research [135]. AHRQ explicitly states that the PQI measures rely solely on inpatient hospital administrative data and do not assess quality of care for patients who are not admitted to a hospital; a known limitation of these indicators. However, the optimal management of the selected five chronic conditions COPD, asthma, hypertension, diabetes, or congestive heart failure should result in the avoidance of an ED visit for treatment of an acute exacerbation. A broader list of conditions as potentially preventable for ED visits among the elderly may be more

48

difficult to identify in terms of preventability [119]. Clearly early recognition of the symptoms of a possible heart attack or stroke, and seeking immediate treatment in the ED would be considered prudent behavior. The total number of ED visits for the five chronic conditions each, and a dichotomous variable indicating any ED visit for one or more of the five conditions each year were coded for each patient.

The medication possession ratio (MPR) for oral diabetic medications was calculated from prescription claim data for patients in the study sample with prescription drug information. The NDC codes were mapped to the National Drug Data File that is annually updated and maintained by First Data Bank (http://www.firstdatabank.com/Products/national-drug-file.aspx) to obtain the major and minor drug classifications for all prescription claims data.

As combined diabetic medication therapy is common in the treatment of diabetes a second claim for oral diabetic medications dispensed on the same day was deleted. Single prescribing events were also deleted. The following classes of oral diabetic medications were identified: sulfonylureas (43.6% of the oral diabetic claims), biguanides (39.4%), thiazolidinediones (13.5%), and current combined oral medications available (less than 5.0%). The total number of days of medication was obtained by summing the days of medication supplied for each dispensed event for each patient per year. The total number of prescribed days was calculated by subtracting the first fill date of each year from the last fill date and adding the days supplied from the last prescription filled. The MPR was derived by dividing the total number of days of supplied medication by the

49

total number of prescribed days for each patient for each year. Consistent with the newly NQF-endorsed oral diabetic medication adherence measure a dichotomous variable for diabetes medication adherence was coded; one indicating the MPR was greater than or equal to 0.80 and zero otherwise.

As diabetes clinical practice guidelines recommend the use of angiotensin-classified medications in persons with diabetes as first-line treatment for hypertension given their renal-protective properties [136] and 92 percent of the patient population was identified as having hypertension the medication possession ratio (MPR) for these medication prescriptions was also calculated. Prescriptions for angiotensin converting enzyme inhibitors (ACE inhibitors) and angiotensin receptor blocker (ARB drugs) were identified by flagging any of four medication classifications: ACE inhibitors, ARB drugs, and ACE with diuretics and ARB with diuretics. The MPR was calculated in the same manner as was done for oral diabetic medications.

Although initially planning to examine each quality and adherence measure separately, as a result of the lack of variation found at the PCP level in some individual measures, and in consideration of the trend in quality reporting toward meaningful composite measures, quality process measures were combined with intermediate outcome (A1c and LDL control) and outcome measures (ACSC admission, ED visit) as composite outcome measures (dependent variables) used in final model analyses The Institute of Medicine (IOM) has further defined quality as the degree to which health services for individuals

50

and populations increase the likelihood of desired health outcomes and are consistent with current professional knowledge [1]. Dependent variables are presented in Table 4.
Table 4. Dependent Variables

Name/Variable Code DM Screen (A1c & LDL) DMScreen DM Screen with A1c Control and without Acute UtilizationScrUTA1c DM Screen with A1c and LDL Control and without Acute UtilizationScrUTCL DM Screening Composite (A1c, LDL, & Colorectal Screen [CR])DMC DM Composite (A1c, LDL, & CR) and without Acute Utilization DMCUT Medication Adherence with Oral Diabetic Medication and without Acute Utilization DMMedAdUT Medication Adherence with Ace/Arb Medications and without Acute UtilizationMedAdUT
* Medication Possession Ratio

Context HEDIS quality process measures Linking process and intermediate outcome measure to the prevention of potentially unnecessary acute utilization Linking process and intermediate outcome measures to the prevention of potentially unnecessary acute utilization HEDIS quality process measures combined into a composite Linking process measure and intermediate outcome measure to prevention of potentially unnecessary acute utilization Linking adherence to prevention of potentially unnecessary acute utilization Linking adherence to prevention of potentially unnecessary acute utilization

Construction
1 = patient had an A1c and LDL screen in eligible year ; else=0 1 = patient had an A1c in eligible year and A1c result was in control and NO ACSC hospitalization or ED visits for chronic conditions; else 0 1 = patient had an A1c and LDL screen in eligible year and both values were in control and NO ACSC hospitalization or ED visits for chronic conditions; else= 0 1= patient had an A1c and LDL screen/eligible year and CR screen following HEDIS specifications; else 0 1= patient had an A1c and LDL screen/eligible year and CR screen following HEDIS specifications and NO ACSC hospitalization or ED visits for chronic conditions; else= 0 1= MPR* = > 0.80 and no ACSC hospitalizations or ED visits for chronic conditions; else=0 1= MPR = > 0.80 and no ACSC hospitalizations and no ED visits for chronic conditions; else=0

51

Patient Sample Selection and Variable Construction


Patients were linked to their assigned primary clinic and primary care physician (PCP) using specified attribution rules and the organizations enrollment files over the four years of the study. All identified patients with less than 12 months of eligibility with the organization were removed from the sample and the time in months for all remaining patients assigned to a PCP (or PCPs) and a site (or sites) was summed. The linked PCP per patient was the PCP with the highest number of months for each patient in the sample. The linked site per patient is the site with the highest number of months. In case of ties (only 15 patients) the PCP and site later in the study period was selected. Two ratio variables, PCP time (MDT) and site time (SiteT), were created by dividing the total number of months assigned to a PCP or site by the total number of eligible months with the organization. A minimum of 0.60 for the MDT was established for retention in the sample as patients with less than 0.6 would have been linked to their primary care physician for less than 60% of the total time with the group (excluded 572 patients). In assessing health care performance, the entity being evaluated must have had an adequate opportunity to affect the aspect of quality that is being measured [137]. All patients with long-term care stays of greater than 90 days per year (451 patients) and patients identified as receiving hospice care (146 patients) were also excluded from the sample as these patient sub-populations represent a different patient population than the identified study sample community-dwelling seniors actively managing diabetes and other chronic conditions. Lastly, patients linked to PCPs with fewer than three patients in the study population were excluded from the sample.

52

Although multilevel modeling techniques are available to address patients linked to more than one physician, cross-classified multilevel models [90, 138-141] as an example, such approaches have not been developed for longitudinal models. The use of cross-classified models is restricted currently to cross-sectional analyses involving a single period of time.

Longitudinal Analyses Covariates


As continuity of care with a primary care physician contributes to the potential for establishing an ongoing relationship and partnership [142-145], four continuity variables were developed. Dichotomous time-varying variables were coded to reflect continuity with the PCP and site each year (PCPCS and SiteCS respectively). Continuity with the PCP and site over the entire period of the study were coded as time-invariant covariates (CtPCP and CtSite).

Patient demographic variables included age at the start of the study coded as a continuous variable, and patient sex coded as a dichotomous variable with one equivalent to male patient. By linking patient zip code to median household income using the 2000 U.S. Census Data three dichotomous income variables were created using the following definitions. Variable, FPL200 indicated patients with zip code linked median household incomes equivalent to 200% of the Federal Poverty Level in 2000 for a couple ($24,000). Variable, Pov10 indicated patient households with income at the 10th percentile ($27,471) of the study patient sample. Variable, Pov25, indicated patient households with income

< 1% of the patient sample (n=116) had missing median household income data. For these few cases, median household income was imputed at the city level with data from the American Fact Finder.

53

at the 25th percentile ($34,824). As race and ethnicity were not observed in the organization enrollment or claim data race/ethnicity variables were not constructed.

Several variables were created that qualify and quantify patient severity. Continuous variables included the following: average medications as average number of distinct medication classifications (as opposed to the average number of prescriptions dispensed), the total number of chronic conditions (AHRQ chronic condition indicators), patient risk factor (Hierarchical Condition Category [HCC] assignment in 2007), overall patient risk (demographic, disease interaction and HCC in 2007), and baseline A1c and LDL lab values. Dichotomous variables included: diagnosis of a high cost condition-COPD, HF, stage III or IV renal disease, or emphysema; diagnosis of anxiety, drugs, alcohol or depression; diagnosis of dementia; patient had a long-term care stay(s) of less than 90 days/year; patient left the organization in 2008 (non-death); diagnosis of tobacco dependence. Patient risk factors for 2007 were used as patient-level, time-invariant covariates because risk factor data was not available for 2004 and among patients terminating the organization in 2008 risk factor data was known for year 2007 only. Average medication values were imputed for 129 patients with missing pharmacy claims data using mean values plus adjustment via ordinary least squares regression with the identified key predictors of patient age, overall risk factor, termed status and Medicaid insurance.

Dichotomous variables were created for SCAN health plan and Medicaid (Medi-Cal) benefits. SCAN is a well-known health plan in southern CA dedicated to the

54

comprehensive, coordinated care for senior patients and the health plan providing coverage for 25% of the patient sample. Medicaid coverage in addition to Medicare denotes patient poverty and/or disability and/or costly chronic medical conditions. Medicaid benefit eligibility was known only for year 2007 and was therefore coded as a dichotomous patient level time-invariant variable.

Time-varying covariates to employ at level one (repeated measures) were created using office visit records (encounters) to produce yearly totals for office visits to primary care, specialists, endocrinologists, health education and urgent care. Using these same records variables were created as time-invariant level two (patient level) covariates; average yearly visit totals for primary care, specialists, endocrinologists, health education, urgent care, and other care (representing pharmacy, therapy, etc. as a measure of additional resource intensity) were coded. Dichotomous variables were also created with one indicating any use of behavioral health, endocrinology, health education, nephrology, and urgent care and zero otherwise.

Physician sex was coded as a dichotomous variable with male physician equivalent to one. Additional physician covariates that included the following: the number of years in practice as categorical variables less than 10 years, 10 to 20 years, and greater than 20 years; average age and average risk of assigned study patients, ratio of referrals to patient caseload, ratio of high cost patients to assigned study patients and ratio of patients diagnosed with anxiety, alcohol, drugs, depression and/or dementia to assigned study patients, the average number of annual referrals, and average patient caseload as

55

continuous variables. Dichotomous variables included: doctor of osteopathy or nurse practitioner or physician assistant, physician trained outside the U.S., geriatric training or board-certification, physician with additional specialty other than primary care, internal medicine board-certified, physician works for affiliated IPA, and physician speaks Spanish.

Cross-sectional Analyses (2007)


The cross-sectional analyses incorporated primary data collection (a patient survey and clinic interviews) to assess relational coordination, trust, and the Chronic Care Model components. A PCP survey was also developed, implemented, and analyzed to assess and model physician-reported care coordination behavior.

Analysis Plan
Cross-sectional analyses examined the association between the key variables of interest, the CCM components, relational coordination and trust, and patient outcomes in 2007. Hierarchical generalized linear modeling (HGLM) was used for final analyses for the same reasons it was used in the longitudinal analyses naturally, nested data structure (theoretically non-dependent observations) with patients linked to PCPs and provided care at assigned clinics/sites. Moreover, as the cross-sectional HGLM analyses included the measured variables of the Chronic Care Model and RC and trust, the influence of these key predictors at the clinic and PCP level were examined directly. Survey response data was linked to quality outcome measures and adherence measures using 2007 patient

56

claims data. Additionally, the HGLM analyses sought to determine if CCM components influenced physician relational coordination and trust and if physician RC and trust moderated or mediated any patient risk covariates. Lastly, what impact would physician self-reported care coordination have on the model? The conceptual model for the crosssectional analyses is depicted in Figure 3.

Figure 3. Conceptual Model for Cross-sectional Analyses Physician Coordination

Physician

Relational Coordination & Trust


(+ covariates)

Chronic Care Model CCM Score

Patient continuity with PCP


(+ covariates)

Physician-Patient Dyadic Team

Patient Quality Outcomes and Adherence

Key Variables of Interest


Key variables of interest were collected via surveys administered in the first quarter of 2008; a mailed patient questionnaire, a web-based PCP survey, and in-person clinic interviews assessing the Chronic Care Model components.

57

Patient Survey
The patient survey was developed over ten iterations drawing upon the survey development for coordination and trust over the past decade. The seven-question Relational Coordination (RC) index [26] was modified for relevance to the patientphysician dyadic team and incorporated the bidirectional nature of the index by expanding the index to fourteen questions. The first seven questions asked the patient to assess his/her PCP with respect to communication, respect and shared goals, and the last seven questions asked the patient to self-report their behavior on these same indices relative to their PCP (Table 5.). The RC index used a five-point frequency rating scale worded appropriately for the question. Responses to one set of questions were never, rarely, sometimes, often and very often. Responses to the other set of questions were not at all, very little, some, a lot, and completely.
Table 5. Relational Coordination Measure

Patient Survey Relational Coordination Index


Subset Question

MD MD MD MD MD MD MD

When you have a medical need how often does your doctor communicate with you? Does your doctor communicate with you in a timely way? Does your doctor communicate with you accurately and honestly? When problems arise does your doctor work with you to solve the problem? Does your doctor understand you and the role you play in your own treatment? Does your doctor respect you?

Does your doctor share your goals for treatment or your plan of care? Patient When you have a medical need how often do you communicate with your doctor? Patient Do you communicate with your doctor in a timely way?

Rating Scale 5 point Never.Very often 5 point Never.Very often 5 point Never.Very often 5 point Never.Very often 5 point Not at all..Completely 5 point Not at all..Completely 5 point Not at all..Completely 5 point Never.Very often 5 point Never.Very often

58

Patient Patient Patient Patient

Patient Survey Relational Coordination Index 5 point Do you communicate with your doctor accurately and Never.Very often honestly 5 point When problems arise do your work with your doctor to Never.Very often resolve them? 5 point Do you understand your doctor and the role he or she Not at all..Completely plays in your treatment 5 point Do you respect your doctor?
Not at all..Completely 5 point Not at all..Completely

Patient Do you share your doctors goals for your treatment?

The key predictor of patient trust in the PCP was measured using an abbreviated validated survey of trust in ones personal physician [29]. The abbreviated survey was selected to minimize patient survey response burden [114] given the insignificant loss of internal reliability. The rating scale consisted of a five-point Likert scale, labeled points of strongly disagree, disagree, neutral, agree, and strongly agree. The first question is reverse-scored (Table 6.).
Table 6. Trust Measure

Patient Trust in Their Primary Care Physician 5-point rating scale: Strongly disagee to Strongly agree with point labels Statement Sometimes your doctor cares more about what is convenient for him or her than about your medical needs. Your doctor is extremely thorough and careful. You completely trust your doctors decisions about which treatments are best for you. Your doctor is honest in telling you about all of the treatment options available for you. All in all, you have complete trust in your doctor. Additional validated indices on patient coordination (Table 7.) were included to further examine care coordination by the PCP as evaluated by the patient. Care coordination indices included the care coordination components of Wagners Patient Assessment of Chronic Illness Care (PACIC) instrument [67] and the Primary Care Assessment Survey (PCC) [64]. The questions from the PACIC asked the patient to rate the frequency of

59

specific behaviors done by their doctors and observable to them. Questions of this nature are found to be less-biased than questions asking patients to rate their doctors in terms of qualitative adjectives such as good, excellent, etc. The items taken from the PCC also asked the patient to rate the frequency of physician behavior, but not all behaviors were observable by the patient. An example of the latter is Your doctor communicates with the other doctors or health providers that you see. A five-point Likert rating scale was used for both domains and consisted of never, rarely, sometimes, often and very often.
Table 7. PCC and PACIC

Patient Survey Additional Coordination of Care Domains 5-point frequency rating scale: never, rarely, sometimes, often, very often Domain Question PCC Your doctor knows when you are due for a check-up. PCC Your doctor keeps track of all your health care. PCC Your doctor follows up with you about visits to other doctors or health providers. PCC Your doctor helps you understand your lab tests, x-rays or information from visits to other doctors. PCC Your doctor communicates with the other doctors or health providers that you see. PACIC How often did your doctor ask you about your preferences about treatment to (CCM) manage your conditions? PACIC How often did your doctor review with you all of the medicine you are (CCM) taking? PACIC How often were you contacted after a visit to see how things were going? (CCM) PACIC How often were you encouraged to attend programs in the community that (CCM) could help you? PACIC How often were you referred to a dietician, health educator or a counselor? (CCM) PACIC How often were you told how your visits with other doctors such as the eye (CCM) doctor helped your treatment? PACIC How often were you asked how your visits with the other doctors were (CCM) going? The patient survey captured other patient factors that have been demonstrated independently to impact patient outcomes and adherence allowing for them to be controlled in model analyses. These included the following: a validated question set for 60

patient self-efficacy in managing chronic illness [146-149] and medication adherence (Morisky Index) [149-153], and individual questions assessing overall health status [154], health literacy [155-158], religiosity [159-163], exercise frequency [164, 165], and difficulty paying for medications [166]. Two questions from the ACES [167] patient satisfaction survey were combined with the support of the author doctor knowledge about patients values and beliefs that may impact treatment. The frequency of telephone calls from the organizations nurse care managers was assessed as this intervention was not observable in the claims data. As ethnicity and language preferences were not collected by the organization these questions were included on the last page of the survey with other demographic questions including the patients living situation. Lastly, two additional outcome variables were collected overall satisfaction with PCP and discussion of end of life wishes with PCP both were hypothesized to be strongly predicted by RC and trust.

The survey was edited following feedback from professional nurses and health educators, and the reading level was reduced to the fifth grade, consistent with Medicare policy. A convenience sample of sixty seniors from the Boston area and representing a more homogeneous, higher educated elder population than the target population participated in the pilot test of the survey. Pilot respondents were instructed to complete the survey without feeling obliged to answer questions honestly but to provide feedback that guided additional revision noted the time it took to complete the survey, suggested language to increase respondent motivation to participate, circled words and questions that were not clear, and noted recommended changes on the survey tool. The final patient survey

61

incorporated findings from the pilot and feedback from professionals. The patient survey and supporting documents were translated into Spanish by a Medicare/Medicaid health plan certified translator to facilitate the participation by all patients identified to receive the survey.

Patient Survey Sampling Framework and Implementation The patient sample for the patient survey linked cross-sectional analyses for 2007 was selected using stratified, random, proportional sampling (12,039 patients identified by the organization December 30, 2007 using study patient criteria) by the 40 clinic/site locations and by linked primary care physician (PCP). PCPs and associated patients were removed from the sampling frame if the PCP had fewer than 10 patients (689 patients were removed as a result leaving a potential sample of 11,350). Three (25%) samples of patients from each clinic/site were selected using the random process in SAS and exporting the associated sampling weights. Three samples were selected in anticipation of the need to execute three survey waves to attain an acceptable patient survey response rate. A maximum of 150 patients per site was imposed to ensure a balanced representation of patients across all the sites.

The patient surveys were mailed in waves between January and April 2008. To frame the reference period of assessment for the patient, the introduction of the survey asked the patients to respond to questions from the perspective over the past 12 months or the past year. Patients were also informed that their medical claims from 2007 would be linked to their survey responses as required by the Brandeis IRB. The survey packet (Appendix

62

E) included a prepaid envelope, a cover letter signed by one of the organizations medical directors, a study protocol introducing myself and describing the study and their potential contribution to it, and two copies of the patient questionnaire (English and Spanish). Additionally to improve the survey response rate an incentive was offered, a drawing for one of 20 WalMart gift certificates. Due to cost constraints only the first two waves of patient survey mailings (wave one=2098 patients and wave two=2356 patients) were executed. A three-step survey process was implemented consisting of the following: 1) mailed the survey packets during first quarter of 2008; 2) mailed reminder post cards two to three weeks later; and 3) mailed a second reminder post card two to three weeks later. Due to cost constraints the entire survey packet was not mailed a second time although the reminder post card instructed patients to request another packet by phone if needed. Approximately 25 additional survey packets were mailed via priority mail upon patient request.

Relational Coordination and Trust


Although the RC index has empirical validation [22, 23] principal components analysis and cluster analysis (SAS Institute Clustering Algorithm) were conducted and a single component solution was confirmed using both raw values and polychoric correlation matrix. Internal reliability of the scale was confirmed via Cronbach alpha (alpha coefficient =0.9495 standardized). If greater than three questions in the RC index (14 questions) or two questions within the PCP portion of the RC index (seven questions) were observed as missing the surveys were excluded; 2.2% (n=12) of remaining surveys had any missing variables ipsative mean replacement was used to impute values [168].

63

Six variables from the RC index were created as follows: patient RC score (average of patient reported behavioral questions on a five-point scale), physician RC scores (average of the average of the fourteen bi-directional RC questions for each respondent linked to the PCP), z-transformed (standardized) patient RC scores, and z-transformed physician RC score (zMDRC). Additional dichotomous variables included high physician RC score (coded one if the z-transformed physician RC score is greater than zero and zero otherwise).

Similarly, principal components analysis was conducted and the Cronbach alpha coefficient was derived on the set of questions constituting the trust domain [29, 98, 169] following reverse scoring on question one. A single construct resulted although the alpha coefficient (at 0.79) was lower than the initially tested 5-question abbreviated trust survey, alpha coefficient of 0.87 [169]. If greater than one question was missing from the trust scale surveys were excluded (5 patients). Less than 5% (n=28) of the trust domains were missing one question and values were imputed using ipsative mean replacement [168]. Three trust variables were created: 1) the mean trust score for each PCP as aggregated from linked patients; 2) the z-transformed (standardized) trust score for each PCP; and 3) a dichotomous variable, high PCP trust with one indicating that the standardized trust score was above average and zero otherwise. Lastly, a composite domain variable (RC and Trust) was created as the high correlation between these variables may prevent the inclusion of both RC and trust in the final HGLMs. RC and trust was combined into a single dichotomous variable with one indicating above average

64

score in trust and above average score in RC and zero otherwise. Both are important to an ongoing relationship with a primary care physician.

Other Patient Survey PCP Domains as Covariates The other care coordination domains examined within the patient survey were processed in the same manner conducting principal components analysis on raw values and the polychoric correlation matrices for the questions comprising the coordination domains of the Primary Care Component survey [64] and the Patient Assessment of Chronic Illness Care survey [67], and the five questions assessing patient self-efficacy [149]. Single component solutions resulted from all hypothesized question sets and internal reliability was confirmed by alpha Cronbach coefficients (Table 8.). The Cronbach coefficient alpha for all PCP survey domains exceeded the lower recommended bound of 0.70 [112114].
Table 8. Key PCP Domains
Patient Survey PCP Domains Standardized Cronbach Coefficient Alpha

PCP RC Trust PCC PACIC

0.9495 0.7944 0.8794 0.8095

All four PCP domains RC, PCC, PACIC and trust were combined into a single dichotomous variable, High MD with one indicating above average score in all four of the measured domains from patient respondents assigned to them. Statistically significant correlations were found among all the care coordination indices. The standardized cronbach alpha for RC, trust, PCC, and PACIC was 0.892. (Appendix F)

65

Ratio variables were also created accounting for the number of survey respondents reporting physician screen for alcohol use, fall risk and depression over the total number of survey respondents and dichotomous variables for each ratio indicating the screen was done for more than half of the PCPs linked survey respondents.

Construction of Other Patient Survey Covariates At the patient level, the patient RC index was constructed as a standardized, continuous variable derived from the average raw score of the responses to the patient subset of the RC index (questions 8 to 14). High patient RC was created as a dichotomous variable denoting that the patient RC index was above average. The Cronbach coefficient alpha was 0.9069 for the patient subset of the RC index. Patient self-efficacy was a continuous variable equivalent to the average score on five 10-point Likert scaled responses to selfefficacy assessment questions; Cronbach coefficient alpha was 0.8430.

Several dichotomous variables were coded to capture additional demographic and other patient information such as non-English speaking patients, living alone without adult children nearby, African-American, Latino, Asian, education less than eight years, education less than high school, low health literacy, regular exercise, high religiosity, help completing the survey, difficulty paying for medication, low medication adherence, physician-patient language discordance and poor health (Table 9.). Self-reported medication adherence was assessed using the validated Morisky medication adherence scale; answering two questions affirmatively indicated low medication adherence.

66

Table 9. Additional Patient Survey Variables

Patient Survey Additional Covariates Coded as dichotomous variables: 1 = response indicated/ 0 otherwise Variable Survey question Patient response DrKnow How much does your doctor know about you as a A lot or most person including your values and beliefs that may everything be related to your medical care? (ACES) DrLangDc Does your primary doctor speak your language? No LowMedAd Do you ever forget to take your medicine? Are you Two or more careless at times taking your medicine? When you questions answered feel better do you sometimes stop taking your as yes medicine? Sometimes if you feel worse when you take your medicine, do you stop taking it? (Morisky index) AA How do you describe yourself? African-American Asian How do you describe yourself? Asian-American Latino How do you describe yourself? Latino/Hispanic EDLT8 How many years of school were you able to Less than 8yrs complete? EDLTHS How many years of school were you able to Less than 8 yrs or 8complete? 11 yrs NonEng What language do you feel the most comfortable Any response other speaking? than English LowHthLt How do you find reading the health information Very difficult or that you receive? difficult Hlpcsurv Did you have help filling out this survey? Yes LvAlone What is your living situation? Live alone children not nearby or live alone no children Rexerc How often do you walk or do some other kind of Often or very often exercise? PoorHlth In general would you say your health is (SF-12) Poor or fair Relig Do you consider yourself a religious or spiritual Moderately or Very person? religious/spiritual DfPayMed How often do you find that you cannot afford to Often or very often pay for your medicine? CM How often did a care manager from XX call you? Often or very often

Patient Survey - Dependent Variables and Variable Construction


The same dependent variables created in the 4-year longitudinal analyses were created using 2007 claim data only. Two additional dependent variables were coded from the 67

patient survey as dichotomous variables: end of life wishes shared with your doctor (1= a lot or most everything), and visit satisfaction with PCP (1= often or very often).

Physician Survey Development


The physician survey was developed via an iterative process using the principles of measurement development [113, 170] to identify facilitators and barriers to care coordination from the physician perspective and self-reported physician behaviors indicative of care coordination (Appendix G). Unlike the patient survey previously validated questions about care coordination as assessed by primary care physicians were not identified. A physician consultant in health quality (Richard Helmer, NCQA surveyors) and physician leaders (Arnold Milstein, Medical Director, Business Group on Health; Robert Blackman, co-founder of XX; Stuart Levine, Medical Director of XX) provided feedback on individual items within hypothesized key domain headings. Feedback from ten health care professionals working in ambulatory care and health care quality was incorporated in finalizing the physician survey.

The physician survey assessed the following key domains of interest that have been associated with care coordination: transfer of information between care settings; communication with patients, style of practice (referrals, screenings, following-up on care provided by other providers), medication management, decision-support tools, and health plan relationships.
Table 10. PCP Survey

PCP Survey Key Domains of Interest 6-point frequency rating scale: never, rarely, sometimes, fairly often, very often, always Domain Question How often did you

68

PCP Survey Key Domains of Interest 6-point frequency rating scale: never, rarely, sometimes, fairly often, very often, always Transfer of Find that the information you needed about your patient was available to Information you at the time of the first visit? Transfer of Information Transfer of Information Transfer of Information Transfer of Information Transfer of Information End of Life End of Life Find that the information you needed about your patient was available to you at the time of the follow-up visits? Find that information was timely following a hospitalization of one of your patients? Find that information was accurate following a hospitalization of one of your patients? Obtain feedback from behavioral health providers on patients referred? Obtain feedback from specialists on patients referred?

Ask about patient preferences for end of life care? For how many patients have you discussed end of life care wishes? (none, few, some, many, most, all) Coordination Refer patients to Disease Management or the dietician within XX? of Care Coordination Refer patients to XX Care Management program? of Care Coordination Refer patients to available community resources to assist with support of Care and/or education? Coordination Initiate follow-up with specialists seeing your patients? of Care Coordination Make sure that the information from specialists and other providers are of Care put into the patients medical chart? Coordination Use the physician portal (EHR) with patient intervention reports to help of Care you manage your patients? Coordination Ensure that advance directives are recorded for your patients? of Care Coordination My senior patients with multi-morbidity and myself are a team working of Care together to ensure positive health outcomes. Medication Review all the medications that your patients are taking with your patients Management during the visit? Medication Ask about over the counter medications and alternative medicine use? Management Medication Simplify the medication regimen of your patients to increase patient Management adherence to your treatment plan? Medication Assess how well medications are working and ask about side effects? Management

69

Questions on the PCP survey under the headings of Organizational Support/Tools and Health Plan Relationships were not included as the responses failed to factor into respective components. Descriptive statistics for these domains are reported under the survey analyses.

PCP Survey Implementation All primary care physicians associated with the identified patient sample were selected to receive the PCP survey with the exception of IPA-affiliated physicians (excluded by the study organization). All remaining primary care physicians provided medical care to associated patients at the group model sites. The total number of PCPs identified was 112. PCPs were contacted via an e-mail invitation, informed about the study, and offered two ways to participate either via the completion of a paper survey or electronically via a link to the secured, web-hosted Survey Monkey survey. All PCPs were informed that administrative claims data from their patient panels would be linked to survey data and to answer questions as related only activities performed in 2007. To improve the physician survey response rate two incentives were offered and an upfront cash incentive of $10 given with a letter inviting study participation, and a chance to win one of six $50 bills upon completing the survey.

PCP Survey - Key Domains of Interest


Principal components analysis and cluster analysis were conducted on survey response variables from the hypothesized domains on the PCP survey and calculate Cronbach coefficient alphas were calculated to confirm internal reliability. Initially the first domain

70

resulted in a two factor solution (patient information at the office visit and follow-up visit versus all other sources of information), eigenvalue for factor one is 2.54 and 1.16 for factor two. However, all items loaded satisfactorily on factor one (factor loadings of 0.55 to 0.75) with the exception of the question, How often do you obtain feedback from behavioral health providers on patients referred? After orthogonal rotation this item was negatively correlated with the others, factor loading of -0.05. Therefore, a single composite variable was created by taking the average among the other items in the information transfer domain. The standardized Cronbach coefficient alpha for the remaining five items in the information transfer domain was 0.7565. The two end of life questions were combined into a single variable as the average of both items, standardized Cronbach coefficient alpha of 0.8587. All of the hypothesized coordination of care domain responses, resulted in a single factor solution, eigenvalue of 3.409 (factor loadings of 0.52 to 0.75) and a standardized Cronbach coefficient alpha of 0.8047. The eight questions were combined into a single care coordination domain variable as the average of all items. The four medication management questions resulted in a single factor solution, eigenvalue of 2.33 (factor loadings of 0.71 to 0.82) and a standardized Cronbach coefficient alpha of 0.7629. A medication management domain variable was created as average of all items.

Few missing values were observed within these domain sets (n=4, 0, 3 and 2 respectively for single items) and imputation was done using ipsative mean replacement [168]. All the domain variables were standardized using z-transformation with a mean of zero and standard deviation of one.

71

Additional Key Variables of Interest and PCP Covariates Additional dichotomous variables were created to capture additional facilitators and barriers to care coordination and physician demographic information (Table 11.). As only six PCPs reported always assessing patient recall of care instructions, the responses of fairly often, very often or always were combined into a single variable, MAssRcll. As no PCP reported health literacy or language discordance to always be a barrier the responses of fairly often and often for each question were combined into the single variables of HthLitB and LangDC. Other dichotomous variables were coded as follows: age greater than 60 years and age less than 35 years, PCP sex (male coded as one), family practice, new with the organization (less than three years coded as one), and high percentage of minority patients, high percentage of patients enrolled in care management, and high percentage of patients for whom physician is actually working with the caregiver rather than the patient.
Table 11. PCP Survey Covariates

PCP Survey Additional Covariates Coded as dichotomous variables: 1 = response indicated/ 0 otherwise Variable Survey question How often did you Patient response MAssRcll Assess recall of your care instructions by asking Fairly often or very patient to tell you what they will do? often or always HthLitB Find health literacy to be a barrier to meeting your Fairly often or very goals during a patient visit? often or [always] LangDC Find language discordance to be a barrier to Fairly often or very meeting your goals during a patient visit? often or always Survey question FP What is your medical specialty? Family Practice MDLE5 How many years have you practiced medicine? < 3yrs or 3 to 5yrs NEW How many years have you worked for XX or the < 3yrs group assumed by XX? AgeGT60 What is your age? >60yrs AgeLE35 What is your age? <30yrs or 30-35yrs PCPSex What is your gender? Male

72

PCP Survey Additional Covariates Coded as dichotomous variables: 1 = response indicated/ 0 otherwise MinPts What percentage of your patients was ethnic >50% or >75% minorities (2007)? CMHigh What percentage of your patients was enrolled in >20 to 33% or >33% Care Management program (2007)? CarGivH For what percentage of your patients did you work >20 to 33% or >33% primarily with the care-giver rather than the patient?

Clinic Survey
The validated Assessment of Chronic Illness Care (ACIC ) tool developed by the MacColl Institute for Healthcare Innovation, version 3.5 (Appendix H) was used to examine the impact of CCM components on patient outcomes and PCP predictors of patient outcomes. This tool assesses each of the six CCM components by rating three to six sub-components per component scored on a scale of zero to 11 (Table 12.). In-person interviews at the organizations group-model clinics with clinic administrators and other invited team members such as nursing managers, patient liaisons, etc. were conducted as previous research finds variable interpretation in the ACIC [20]. Clinic interviews were restricted to the organizations group model clinics per request of the organization. Key words within each sub-components assessment were highlighted for consistent interpretation and emphasis across all the clinics. Face-to-face meetings allowed for clarification of questions regarding the assessment of any sub-component. All clinic interviews were completed by the end of the first quarter of 2008; clinic teams scored each sub-component based upon the 2007 year. Additional information collected as an introduction to the ACIC instrument included the following: the estimated demographic

73

profile of patients served, percentage of visits attributed to seniors, and the presence of urgent care, care managers and health education services on-site.
Table 12. Assessment of Chronic Illness Care (ACIC)

Assessment of Chronic Illness Care


Part Component Sub-components

Organization of healthcare delivery system chronic care

Community Linkages

3a

Self-Management Support

3b

Decision Support

3c

Delivery System Design

3d

Clinical Information Systems

Integration of Chronic Care Model Components

Overall organizational leadership Organizational goals Improvement strategy Incentives and regulations Senior leaders Benefits Linking patients to resources Partnerships with CBOs Health plans Assessment and documentation Self-management support Address concerns pts & family Effective behavior change interventions Evidence-based guidelines Involvement of specialists Provider education Informing patients of CPGs Practice team functioning Practice team leadership Appointment system Follow-up Planned chronic illness care Continuity of care Registry Reminders to providers Feedback Information patient subgroups Patient treatment plans Information patients of CPGs Registries Community Programs Organizational planning Routine f/u for patient planning Guidelines for chronic care

74

Variable Construction from Clinic Survey Assessments


Scores were summed for each sub-component and the total number was divided by the number of items in the component in obtaining a component score. The total of all individual component scores were averaged to derive an overall CCM score ranging from zero to 11 for each clinic. Variables created for each clinic included each individual component score, the overall CCM score, and the overall CCM score minus Integration, (the newest section of the instrument). Additionally, a variable averaging all the subcomponent scores for Part 3 was developed as the practice-level components have been identified previously are key drivers in improving patient outcomes [20, 56, 171, 172].

Additionally, continuous variables were created for the average clinic visits in 2007 and the percentage of minority patients. Dichotomous variables were created for the presence of on-site health education or care management and the presence of on-site urgent care.

75

Chapter 4: Descriptive Statistics


Longitudinal Analyses:
Final Patient Sample
The patient is the unit of analysis and sample size fluctuated per model analysis as data was not available on all outcome measures for identified patients. Additionally, about 40% of the identified patient sample were linked to PCPs without data on the key physician domain variables of access, communication and coordination (n=4357). As these patients had to be excluded from all longitudinal analyses, the two patient groups were examined for differences with the significant differences reported (Table 13.). Overall, the patients retained in the final patient sample have a higher disease burden, lower continuity with a primary care physician, fewer annual average primary care visits and higher use of other services.
Table 13. Patient Sample Descriptive Statistics

Variable

Label

AvgOther

Average annual use of other services AvgPCVis Average annual primary care visits AvgSpVis Average annual visits with specialists FirstA1c Baseline A1c MDT Time-majority PCP/total time PhyMos Months-majority PCP SiteT Time-majority site/total time Smos Months-majority site TotChrDz Chronic conditions-observable in claims
**

Final Pt Not in Final Sample Pt Sample n=5760 n=4357 Mean [standard deviation] 2.9 [4.0] 2.4 [3.7] 6.5 [3.4] 5.5 [4.8] 7.0 [1.4] 0.86 [0.1] 36.3 [8.3] 0.97 [0.1] 42.1 [9.5] 15.2 [6.5] 6.9 [3.8] 5.1 [4.9] 6.8 [1.3] 0.93 [0.1] 40.5 [10.2] 0.98 [0.1] 42.6 [9.8] 14.6 [6.5]

P Value
P Value ** < .0001 < .0001 < .0001 < .0001 < .0001 < .0001 < .0001 0.009 < .0001

Students t-test

76

Variable
TotRAF ADDA

Label
Overall risk factor for 2007

Alcohol drugs depression or anxiety AnyBhlth Any visits to behavioral health AnyEndo Any visits to endocrinologist AnyHE Any visits to health education AnyUCare Any visits to Urgent Care CtPCP All time with same PCP CtSite All time at same site FPL200 Median household inc < $24,000 Group Patient in group model care Pov10 Median household inc < $27,471 Pov25 Median household inc < $34,824 Scan SCAN insurance plan (member) Dementia Dementia HCostDz End-stage renal disease, CHF, COPD and/or emphysema TobDep Tobacco dependence

Final Pt Not in Final Sample Pt Sample 2.7 [1.4] 2.5 [1.4] % Patients (n) 26.5 (1503) 23.5 (1022) 3.0 (168) 9.9 (562) 7.3 (412) 1.0 (57) 37.1 (2106) 87.6 (4968) 4.6 (262) 86.0 (4876) 10.0 (570) 25.0 (1407) 25.5 (1444) 9.0 (516) 57.4 (3254) 3.0 (178) 2.1 (93) 5.1 (224) 3.2 (140) 3.0 (135) 71.9 (3133) 90.0 (3920) 3.4 (150) 8.8 (383) 4.7 (205) 16.4 (714) 33.2 (1445) 11.0 (470) 50.0 (2180) 1.9 (84)

P Value
< .0001 P Value 0.0005 0.0098 < .0001 < .0001 < .0001 < .0001 0.0002 0.0032 < .0001 < .0001 < .0001 < .0001 0.0049 < .0001 0.0002

Dependent Variables
The study patient population had low rates of acute utilization as assessed by ambulatory care sensitive hospitalizations, hospital readmissions and Emergency Department visits (Table 14). Ambulatory care sensitive hospital admissions were under seven percent with a rate of 66.9/1000 and hospital readmission rates even up to 90 days post discharge were under five percent with rates of 28.6/1000 for 30-day readmissions, 36.2/1000 for 60-day readmission, and 40.5/1000 for 90-day readmissions. ED visits for the common chronic conditions of diabetes, CHF, COPD/asthma, and/or hypertension were two percent with a rate of 21.9/1000. Rates of compliance with annual A1c and LDL diabetes screens concurrently were 72 percent. The annual A1c screening rate was 92% (not

Pearson chi-square test

77

shown as compliance was so high there was no between provider variation to model). Diabetes control rate, an intermediate outcome of quality, was slightly lower at 82 percent using the more liberal control criteria and fell to 66% when using the 2007 A1c control definition of less than seven percent across all years. The rate for A1c and LDL control was 67 percent and fell to 55 percent when applying the 2007 control criteria for A1c across all years. The diabetes screening composite variable (DMC) consisting of an A1c, LDL screen and a colorectal screen was very low at 29%. Both medication adherence measures were relatively high with adherence to oral diabetes medications at 83 percent and adherence to Ace Inhibitors/Arb drugs at 84 percent. In combining each positive quality and adherence dependent variable with the absence of an ambulatory care sensitive hospitalization or ED visit for one of the five common chronic conditions measure rates fell by two to six percent.
Table 14. Model I Dependent Variables

Variable
A30Rmit A90Rmit AnyACSC AnyEDChr AnyED DMScreen DMScrUT DMC DMCUT HEDA1cCL A1c_UT HEDISCL HEDIS_UT DMRXAdh DMRxUT MedAdh MedAdUT

Label
Any hosp readmission 30 days after D/C Any hosp readmission 90 days after D/C Any ACSC Any ED visit for DM, HF, COPD, asthma or HTN Any ED visit A1c and LDL Screens A1c and LDL Screen No Acute Ut* A1c, LDL & Colorectal screens A1c, LDL, CR and No Acute Ut* A1c control** A1c control and No Acute Ut* A1c & LDL control** A1c & LDL control and No Acute Ut* Adherence to oral diabetes meds Adherence to oral DM meds & No Acute Ut* Adherence to Ace/Arb meds Adherence to Ace/Arbs & No Acute Ut*

N
20436 20436 20436 20436 20436 19905 19905 15013 15013 18568 18568 16629 16629 12159 12159 10854 10854

Mean

SD
0.17 0.20 0.25 0.15 0.38 0.45 0.47 0.45 0.44 0.38 0.43 0.47 0.48 0.37 0.42 0.37 042

0.03 0.04 0.07 0.02 0.18 0.72 0.66 0.29 0.27 0.82 0.76 0.67 0.62 0.83 0.77 0.84 0.77

CV 582.54 486.95 373.35 667.21 216.85 62.69 71.12 157.98 166.18 46.49 56.15 70.13 77.74 45.08 55.16 44.12 55.13

* No Ambulatory Care Sensitive Condition hospitalization and no ED visit for DM, CHF, COPD, asthma or hypertension (HTN) ** A1c control defined as < 8% for 2004, 2005, & 2006 and < 7% for 2007

78

The change in overall means for these dependent variables was examined over time (Figure 4.). The screening process measures and acute utilization rates increased linearly with time. However, the composite variables examining A1c and LDL control without acute utilization, and medication adherence without utilization did not demonstrate a linear trend. These results informed the use of time in final model analyses.

Figure 4. Overall Mean Results for Dependent Variables Over Time


Mean Dependent Variables Over Time
1 0.9 0.8 0.7 0.6

Percent

Time 0 Time 1 Time 2 Time 3

0.5 0.4 0.3 0.2 0.1 0


SC An yE D C hr An yE D D M Sc re en D M sc rU T L R xA dh Ac ea dU T Ac eA dh R M IT IT L M C U T ut T ed is C D M C yA C IS _ R XU T M a1 C R M c_ U A1 D

A3 0

A9 0

ED

ED

An

M D

Quality and Adherence Outcomes

PCP Descriptive statistics


The predicted values for the key PCP domain variables of access, communication and coordination were not normally distributed with the exception of predicted access. Predicted communication was skewed to the right (-1.6 skewness statistic) with a long 79

flat tail extending negatively (kurtosis statistic of 5.7). Predicted coordination assumed a similar distribution with a skewness statistic of -1.7 and kurtosis statistic of 7.2. Consequently, the sum of predicated communication and coordination had the same distribution (Appendix E: PCP Domain Distribution Histograms). The dichotomous variables for access, communication and coordination were slightly less than 50 percent as a result of the predicted values for key domains derived via hierarchical generalized linear model with random error. The dichotomous variable for the sum of predicted ztransformed values for communication and coordination was 38 percent.

Sixty percent of all PCPs were male and 24 percent speak Spanish. Twenty nine percent of all PCPs were associated with an IPA rather than being employees of the study organization. Fifty-seven percent were board-certified in Internal Medicine, six percent reported an additional specialty, four percent had training in geriatrics, and 34 percent were foreign-trained physicians. Forty percent of PCPs reported more than 20 years in practice, 33 percent ten years to less than 20 years and 27 percent less than 10 years. PCP patient panels varied from 40 to 2640 patients (mean of 913 patients); referral ratio of average annual referrals over average patient caseload ranged from 0.28 to 1.9 (mean of 0.60 referral/patient). Contextual PCP variables created by averaging key patient characteristics for linked patients (age, risk burden, high cost illness and mental health and/or substance) are also reported (Table 15).

80

Table 15. PCP Univariate Statistics Variable AvgACS AvgCM AvgCD PredACS PredCM PredCD PredCMCD HPredACS HPredCM HPredCD HPdCMCD PCPSex Lt10yrs Yrs10_20 Yrs_GE20 SP IPA IM Spec Geri FTr DOPANP AvRef Panel RefRat SptPn AvgTRAF AvgAge AvgRsk HCostPts ADDDAPts Label Average Access Domain Average Communication Domain Average Coordination Domain Predicted z-score Access Predicted z-score Communication Predicted z-score Coordination Pred z-scores, Comm + Coord Pred z-score Access > average Pred z-score Communication > avg Pred z-score Coordination > avg Pred z-score Comm + Coord > avg Male PCP In practice less than 10 years In practice 10 to 20 years In practice greater than 20 years Spanish-speaking PCP Independent Practice Association Board-certified Internal Medicine Specialty beyond IM or FP Geriatric training or board-certified Foreign-trained PCP D.O., P.A. or nurse practitioner Average annual referrals Average pt panel or caseload Ratio of avg referrals/pt panel Ratio of study pts/pt panel Avg total risk of study pts in 2007 Avg Age of study pts in panel Avg HCC of study pts in 2007 Ratio of high cost pts/all study pts Ratio of mental health pts/study pts N 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 179 Mean 3.94 4.41 4.20 -0.04 -0.08 -0.07 -0.15 0.47 0.47 0.49 0.38 0.60 0.27 0.33 0.40 0.24 0.29 0.57 0.06 0.04 0.34 0.08 500.23 913.40 0.60 0.05 2.56 73.86 1.91 0.53 0.36 Min. 2.97 3.15 2.78 -2.70 -4.51 -4.86 -8.86 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 29.50 40.00 0.28 0.01 0.88 63.00 0.18 0.00 0.00 Max. 4.60 4.82 4.67 2.17 1.52 1.77 3.26 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1220.25 2640.00 1.94 0.50 5.05 87.00 4.02 1.00 1.00 Std 0.22 0.22 0.20 0.77 0.87 0.82 1.62 0.50 0.50 0.50 0.49 0.49 0.44 0.47 0.49 0.43 0.46 0.50 0.23 0.19 0.47 0.28 286.85 556.47 0.25 0.05 0.55 2.84 0.48 0.22 0.20

Patient Descriptive statistics


Group assigned patients constituted the majority of the patient population at 86 percent and one fourth were members of the SCAN insurance plan, a health plan providing an array of comprehensive supportive services and resources to frail seniors. While the final study population was required to have at least 12 months of eligibility with the managed 81

care organization, the vast majority of patients were eligible for the entire 48 months of the study (mean number of months is 43.7) and spent the majority of eligible months with their linked PCP (mean number of months was 36.3). The ratio of months with linked PCP over total months of eligibility ranged from 0.60 to 1.0 by study design however the mean ratio was 0.88 indicating the majority of time was spent with the linked PCP. However, only 37 percent of patients remained with their linked PCP for the entire period of the study whereas, 88 percent of patients received care at the same clinic or site over the study period. Medicaid eligibility was known only for year 2007 and 11 percent were identified as receiving Medicaid/Medicare benefits in that year. Only five percent were classified at the federal poverty level for household income (as determined by ZIP code proxy). Slightly less than half the patients were male (46 percent) and the mean and median ages were 74, with the oldest patient being 98 years old. Six percent left the organization in 2008 and 60 percent were the same sex as their PCP.

The overall risk (age, sex, and illness burden) for the patient population ranged from 0.37 to 12.10 with a mean overall risk of 2.7. The average number of distinct medication classes ranged from one to 40.5 with a mean of nine medication classes. The number of chronic illness ranged from two to 46 with a mean of 15. The proportion of the patient population with high cost disease (stage III or stage IV renal disease, CHF, COPD and/or emphysema) was very high at 57 percent. Twenty-seven percent had diagnoses of anxiety, depression, or substance use, nine percent had dementia and three percent were tobacco dependent. The mean baseline A1c and LDL values for the patient population were 7% and 106.4 mg/dL respectively.

82

Patients had an average of annual 6.5 visits to primary care and an average of 5.5 visits to specialists. Use of other services including physical and occupational therapy, pain management, wound care, acupuncture, chiropractic services, etc. by patients averaged 2.9 visits per year. At least 10% of patients visited an endocrinologist, four percent of patients had visits with a nephrologist, and seven percent of patients received health education services at least once over the study period. A complete summary of patient characteristics in the sample are reported in Table16.
Table 16. Patient Descriptive Statistics Variable
AgeS Group Scan TotalMos PhyMos CtPCP CtSite PCPs Sites MM Termed FPL200 Pov10 Pov25 Ptsex RskF TotRAF Smos AvgMeds MDT Sexconc avgPCVis AvgSpVis AnyEndo AnyNeph avgOther AnyBHlth AnyHE AnyUCare Hcostdz

Label
Age as of January 1, 2004 Patient in group model care SCAN insurance plan Total months of eligibility with group Months-majority PCP All time with same PCP All time at same site Total number of PCPs Total care sites Medicaid eligibility in 2007 Patient left the group in 2008 Median household income < $24,000 Median household income < $27,471 Median household income < $34,824 Male (male=1) HCC in 2007 Overall risk factor for 2007 Months-majority site Average # distinct medication classes Time-majority PCP/total time Patient Sex Matches Linked PCP Sex Avg Annual Visits to Primary Care Average Annual Visits to Specialists Any Visits to an Endocrinologist Any Visits with Nephrologist Average Claims for Other Services Any Visits with Behavioral Hlth Any Visits to Health Education Any Visits to Urgent Care End-stage Renal, CHF, COPD

N
5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5675 5673 5675 5675 5675 5675

Mean 74.22 0.86 0.25 43.67 36.31 0.37 0.88 1.77 1.15 0.11 0.06 0.05 0.10 0.25 0.46 1.99 2.68 42.08 9.26 0.86 0.60 6.53 5.50 0.10 0.04 2.89 0.03 0.07 0.01 0.57

Median 74.00 1.00 0.00 48.00 36.00 0.00 1.00 2.00 1.00 0.00 0.00 0.00 0.00 0.00 0.00 1.65 2.27 48.00 8.80 0.80 1.00 6.00 4.30 0.00 0.00 1.50 0.00 0.00 0.00 1.00

Min.
62.00 0.00 0.00 12.00 7.00 0.00 0.00

Max.
98.00 1.00 1.00 48.00 48.00 1.00 1.00 6.00 4.00 1.00 1.00 1.00 1.00 1.00 1.00 11.41 12.10 48.00 40.50 1.00 1.00 50.30 60.30 1.00 1.00 44.00 1.00 1.00 1.00 1.00

Std Dev
6.78 0.35 0.44 8.48 8.27 0.48 0.33 0.70 0.41 0.31 0.25 0.21 0.30 0.43 0.50 1.25 1.45 9.48 4.17 0.13 0.49 3.42 4.83 0.30 0.20 3.97 0.17 0.26 0.10 0.49

1.00
1.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.37 7.00 0.00 0.60 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

83

Variable
TotChrDz ADDA Dementia Tobdep Firsta1c Firstldl

Label
Chronic illnesses-observed claims Anxiety, Depression, Substance Abuse Dementia Tobacco Dependence Baseline A1c result Baseline LDL result

N
5675 5675 5675 5675 5269 4956

Mean 15.17 0.27 0.09 0.03 7.08 106.41

Median 14.00 0.00 0.00 0.00 6.80 103.00

Min.
2.00 0.00 0.00 0.00 4.20 21.00

Max.
46.00 1.00 1.00 1.00 18.30 286.00

Std Dev
6.56 0.44 0.29 0.17 1.39 34.03

Descriptive Results of Hierarchical Clustering


Another important aspect to the descriptive analysis in preparation for fitting hierarchical models is the assessment of data cluster size at each level of the proposed analysis. As the patient and PCP samples changed depending upon the respective dependent variable, the cluster sizes varied across models. A minimum of three patients per PCP was required initially for the final patient and PCP samples as fewer than three do not permit discovery of variation [90, 93]. PCP and patient sample sizes and unit clusters for the potential sample in the longitudinal analyses are reported in Tables 17 and 18. As this study was a retrospective longitudinal analysis of actual health care organization data optimal sample sizes for clusters were limited by the actual number of patients and physicians identified and existing relationships. The total number of potentially eligible patients at level two was 5675 and the number of PCPs at level three was 179. The majority of PCPs had at least 20 patients (57.5%) in the study and just over one fourth (26.2%) had fewer than ten patients. Additionally, the majority of PCPs (53.1%) were linked to patients with at least three full years of data in the four-year study period. Overall, 75 percent of all patients had all four years of measurement with the organization. 84

Table 17. Level 3 Model I PCP Sample with Linked Patients


Number patients linked to PCP Number of PCPs % of PCPs % of linked patients with all time points Number of PCPs % of PCPs

10 > 10 - < 20 20 - < 35 35 - <50 > 50

47 29 40 22 41 179

26.2 16.2 22.3 12.3 22.9 100.00

= > 90 = > 75 = > 50 < 50

95 53 19 12 179

53.1 29.6 10.6 6.7 100.00

Table 18. Level 2 Model I Patient Sample Size


Number of time points observed Number of patients % of patients Cumulative frequency of individuals

1 2 3 4

224 394 792 4265

4.0 7.0 14.0 75.1

224 618 1410 5675

Cross Sectional Analyses 2007 Survey Linked Data


Descriptive Statistics Patient Survey Responses
Although two waves of patient surveys were implemented the response rate was 19.6 percent. A total of 586 usable surveys were linked to claims data for subsequent analyses. The corrected patient survey denominator was 2990 following the removal of selected patients determined to be ineligible for the survey. These patients included the following: patients institutionalized or in SNF settings for greater than 90 days in 2007 (451 patients), patients terming the organization in 2008 (895 patients), patients receiving hospice during 2007 (83 patients), patients with a diagnosis of dementia (1103 patients), and patients with fewer than six months linked to their PCP in 2007. Losses among the 764 surveys returned and unable to be used included the following: patients without a

85

diagnosis of diabetes (spouse completing survey mailed to the home), patients with fewer than six months linked to their PCP in 2007, diagnosis of dementia (35 patients), and patients not completing the requisite number of questions within the RC and trust indices.

The vast majority of patients actually had the same sample selection probability with the exception of those patients coming from very high volume or very low volume sites (sampling weight for 64.6% = 1.33; 6.9%, 1.0 to 1.32; 7.2% = 1.44; 12.6%, 3.8 to 4.0). Selection probabilities were based upon sites not physicians therefore in models examining patient at level-one and PCP at level-two sampling weights were not used.

The patient responders were compared to the non-responders among the patient population who were selected to be surveyed to examine potential differences between the population groups. Unlike a common situation with surveys completed anonymously without knowledge about the comparability between responders and non-responders the entire eligible survey population could be successfully linked to their enrollment and claims data in assessing comparability between groups. Overall, few significant differences were observed between patients responding to the survey and those selected for the survey but not responding. None of the risk variables other than total medication classes dispensed in 2007 were significant. Patient responders had slightly higher incomes overall although no differences were observed in Medicaid benefits or the lowest level of poverty variables. Significant and marginally significant differences within enrollment and claims data are reported in Table 19.

86

Table 19. Patient Survey Responders as Compared with Non-responders


(Significant or marginally-significant differences only are shown)

Variable

Label

Responders

Nonreponders

P Value
P Value < .0001 .005 P Value** 0.04 0.04 0.008 0.0002 < .0001 0.01 0.10

MDT TotMeds ADDA AnyNeph CtPCP CtSite Group Pov25 Scan

Time-majority PCP/total time Tot number of med classes -07 Alcohol drugs depression or anxiety Any visits to a nephrologist All time with same PCP All time at same site Patient in group model care Median household income < $34,824 SCAN insurance plan (member)

n=586 n=2404 Mean [standard deviation] 0.97 [0.1] .095 [.01] 10.8 [5.0] 10.1 [5.0] % Patients (n) 0.1 (2) 1.3 (32) 9.2 (54) 88.6 (519) 87.6 (4968) 86.4 (506) 21.0 (123) 28.5 (167) 12.2 (293) 84.2 (2024) 90.0 (3920) 76.0 (1826) 25.9 (622) 25.2 (1445)

* Students t-test ** Chi-square test

Dependent Variable Descriptive Statistics


The patient was the unit of analysis and sample size fluctuated per model analysis as data was not available on all outcome measures for survey respondents. The acute utilization rates were similar to those found in the longitudinal patient sample representing a much larger patient population; any ACSC hospitalization was only 6 percent (7 percent in the longitudinal sample). Diabetes screens were higher at 84 percent versus 72 percent in the longitudinal patient sample. The diabetes composite screening measure for A1c, LDL and CR screens was much higher at 42 percent (versus 29 percent in the longitudinal patient sample) however the 2007 colorectal screen measure was constructed from claims data from the full four-year look back period. A1c control was 67 percent and similar at the more restrictive 2007 A1c criteria to the longitudinal patient sample (66 percent). A1c and LDL control were identical at 55 percent. Medication adherence was higher 87

among patient survey respondents at 89 percent for both oral diabetic and Ace/Arb mediations (medication adherence rates in the longitudinal patient sample were 83 and 84 percent respectively). The rates for all dependent variables among survey respondents are reported in Table 20.
Table 20. Patient Survey Linked Dependent Variables (Dichotomous Variables 0/1) Variable Label N Mean Std Dev AnyACSC Any Ambulatory Care Sensitive hosp 586 0.06 0.25 AcuteUt Any ACSC or ED visit for chronic ill 586 0.07 0.26 AnyRadmt Any Readmission within 30-days hosp 586 0.06 0.24 DMScreen Diabetes Screens A1c and LDL 2007 584 0.84 0.36 SMScUT Diabetes A1c & LDL without acute ut 584 0.79 0.4 DMC Diabetes A1c, LDL and CR screens 469 0.42 0.49 DMCUT Diabetes composite without acute ut 469 0.38 0.49 A1c_Cl HEDIS A1c Control 494 0.67 0.47 A1c_clUT HEDIS A1c Control without acute ut 494 0.63 0.48 A1cldlcl HEDIS A1c & LDL Cl 450 0.55 0.50 A1cldUT A1c & LDL Cl without acute ut 450 0.52 0.50 MedAdh Adherence to oral diab meds > 0.80 349 0.89 0.31 MedadUT DM meds adherence without acute ut 349 0.82 0.39 AceAdh Adherence to Ace/Arb > 0.80 398 0.89 0.31 AceadUT Ace Adherence without acute ut 398 0.82 0.39 HSat Satisfaction with PCP; often, very often 586 0.78 0.41 TotSat Satisfaction with PCP; very often 586 0.39 0.49 DiscEOL Discussed end of life wishes -PCP 586 0.25 0.43

Patient Descriptive Statistics


The patients responding to the survey had similar utilization patterns to the patients in the longitudinal sample with eight percent having any health education visits during the year, 12 percent having any visit with an endocrinologist, and the average total visits to primary care and to specialists were 6.8 each (slightly higher for visits to specialists than in the longitudinal patient sample). The average overall patient risk score was 2.61 (2.68 in the longitudinal patient sample), 12 percent had Medicaid benefits in addition to Medicare (11 percent in the longitudinal sample), and 28 percent were members of

88

SCAN health plan (25 percent in the longitudinal sample). The average number of years that survey respondents reported seeing their current PCP was 7.7 years and 89 percent had the same PCP for the entire 2007 year. The survey respondents represented the known diversity in the overall patient study population with 45 percent identified as a member of a minority group; seven percent self-identifying as African American, nine percent as Asian, and 29 percent as Latino. Forty seven percent of the respondents were male and the average age in 2007 was 77 years (the same average age found in the longitudinal patient sample).

The hypothesized potential barriers of language and lower levels of formal education were confirmed by survey respondents. Twenty-four percent reported less than a high school education with 11 percent reporting less than eight years of education and eight percent reporting difficulty reading health information provided to them. Nineteen percent denoted a preferred language other than English and nine percent reported that their doctor did not speak their same primary language. Fourteen percent reported living alone without adult children living nearby, 33 percent responded that their overall health was poor or fair, and 11 percent reported difficulty paying for medications.

Hypothesized protective factors were also conveyed by survey respondents. Fifty percent of respondents indicated that they walked or exercised often or very often and 79 percent reported being moderately to very religious or spiritual. Self-reported medication adherence was very high at 80 percent (consistent with actual pharmacy-claims generated adherence measures), and survey respondents denoted high levels of patient self-efficacy

89

for self-management of diabetes and their other chronic conditions (average self-efficacy on a ten-point scale was 7.9).

A key patient covariate, the patient component of the Relational Coordination index was skewed with a mean of 4.21 on a five-point scale.
Table 21. Patient Survey - Patient Descriptive Statistics

Variable
PtRC zPtRC HighPtRC AA ADDA Agee AnyEndo AnyHE AnyNeph Asian CM CtPCP CtSite DfPayMed DrKnow DrLangDc EdLt8 EdLtHS Group HighSE Hlpcsurv Latino LowHltLt LowMedAd LvAlone MDT MinPt MM NonEng PoorHlth Pov25 PtSex Relig RExerc SCAN

Label
Patient RC Score z-transformed Pt RC score Pt RC above average African American Patient Dx of anxiety, depression or drug abuse Age in 2007 Any endocrinology visits Any health ed visits Any nephrology visits Asian Patient Often or very often CM calls Same PCP all of 2007 Same site of care all of 2007 Has difficulty paying for meds PCP knows values/beliefs PCP does not speak pt lang. Education < 8 yrs Education < High School Pt assigned to group PCP Pt self-efficacy = > 7 Pt had help to fill out survey Latino Patient Difficulty reading info Morisky = < 2
Live alone without adult children

N
586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586 586

Mean
4.21 -0.00 0.54 0.07 0.00 76.96 0.12 0.08 0.09 0.09 0.10 0.89 0.97 0.11 0.60 0.09 0.11 0.24 0.86 0.51 0.10 0.29 0.08 0.20 0.14 0.97 0.45 0.12 0.19 0.33 0.21 0.47 0.79 0.50 0.28

StDev
0.62 1.00 0.50 0.26 0.06 6.44 0.32 0.27 0.28 0.28 0.31 0.32 0.16 0.31 0.49 0.29 0.31 0.43 0.34 0.50 0.30 0.45 0.28 0.40 0.35 0.09 0.50 0.32 0.39 0.47 0.41 0.50 0.41 0.50 0.45

Min
1.43 -4.46 0.00 0.00 0.00 64.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.50 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

Max
5.00 1.26 1.00 1.00 1.00 98.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00

Ratio of months linked to PCP Minority Patient Medicaid benefits Preferred lang other than Eng SF-12 overall hlth poor or fair Median household inc < $34,824 Patient male Moderately or very religious
Walks or exercises often/very often

Member of SCAN health plan

90

Variable
SE SiteT TotChrDz TotRAF TPCVIS TSpVIS YrsMD

Label
Self-efficacy, 0 to 11 Lorig Ratio of months linked to site Total number chronic conditions Risk Factor overall 2007 Total PC Visits in 2007 Total Specialist visits 2007 Years with PCP

N
586 586 586 586 586 586 586

Mean
7.89 0.99 2.65 2.61 6.80 6.77 7.68

StDev
1.82 0.04 3.06 1.34 3.72 5.93 6.71

Min
1.00 0.60 1.00 0.51 0.00 0.00 0.50

Max
10.00 1.00 19.00 8.05 27.00 40.00 45.00

PCP Descriptive Statistics Patient Survey Level-Two


Overall patient respondents gave their PCPs relatively high scores for key predictor variables, Relational Coordination (RC) and trust, with an average of 4.1 each and minimum scores of 2.9 and 3.1 respectively. Thirty-six percent of PCPs had above average scores on RC and trust scores. PCPs were rated lower on the care coordination domain from PACIC with an average score of 2.6. PACIC assesses PCP behavior versus the affective dimensions of RC and trust. Thirty-three percent of PCPs scored above average on RC, trust and PCC indices and thirty percent of PCPs scored above average on RC, trust, PCC and PACIC indices. Twenty-nine percent of PCPs assessed the risk for falls in more than half of linked survey respondents whereas only 17 percent assessed alcohol use and 11 percent screened for depression in more than half of linked respondents. The average score for PCP review of medications was 3.9 although the variation in these scores was the greatest at a minimum of 0.59 and a maximum of 5.0. The average score for PCP assessment of patient treatment preferences was 3.4 but also had a wide spread (0.62 to 4.67).

Sixty-five percent of the PCPs linked in the patient survey sample were board-certified in Internal Medicine, five percent had training in geriatrics and 34 percent were foreign-

91

trained. Sixty-two percent were male physicians and forty-four percent had practice medicine for more than twenty years. Eighteen of the PCPs were associated with an IPA. PCP descriptive statistics are reported in Table 22.

Table 22. PCP Descriptive Statistics - Patient Survey - Level Two Variable Label N Mean
MDRC zMDRC HighMDRC MDtrust zMDtrust HTrust HMDRCTrt MDPCC zmdpcc MDPACIC zmdpacic HighMD3 HighMD MDRCall zmdrcall HMDRCall MDAssFal MDassAlc MDassDp Medsrev zmedsrev PrefTx zpreftx DOPANP FTr geri IM IPA LT10yrs Panel07 PCPsex Ref2007 refrat Spec yrs10_20 YrsGE20 Relational Coordination (RC), avg score z-Transformed RC - PCP component Above average on MD RC Patient reported trust in PCP, avg score z-Transformed Trust in PCP scale Above average on Trust scale Above average on MD RC & Trust Primary Care Component (PCC), avg score z-Transformed PCC- care coordination Pt Assess of Chronic Illness Care, avg score z-Transformed PACIC - care coordination PCP score > average on RC, Trust, and PCC PCP score > av on RC, Trust, PACIC, and PCC Full RC Index aggregated to PCP, avg score z-Transformed RC - full index Above average on full RC index PCP assessed fall risk > half of patients PCP assessed alcohol use in > half of patients PCP assessed depression in > half of patients PCP reviews all medicines patient is taking z-transformed value for medication review PCP assesses pt preference about treatment z-transformed - pt preferences for treatment PCP is DO, PA or NP PCP medical training outside the U.S. PCP is geriatrician PCP board certified Internal Medicine PCP is affiliated with member IPA PCP has practiced medicine less than 10 years Patient caseload in 2007 Male PCP Total # of referrals issued by PCP in 2007 Ratio of PCP referrals to patient caseload 2007 Another specialty other than primary care PCP has practiced medicine for 10 to < 20 years PCP has practiced medicine => 20 years 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 103 4.12 0.00 0.51 4.07 0.00 0.45 0.36 3.99 -0.00 2.58 0.00 0.33 0.30 4.17 0.00 0.55 0.29 0.17 0.11 3.94 0.00 3.39 -0.00 0.07 0.34 0.05 0.65 0.18 0.20 1183.88 0.62 675.27 0.63 0.04 0.36 0.44

Std Dev
0.38 1.00 0.50 0.32 1.00 0.50 0.48 0.41 1.00 0.53 1.00 0.47 0.46 0.35 1.00 0.50 0.46 0.37 0.31 0.59 1.00 0.62 1.00 0.25 0.48 0.22 0.48 0.39 0.40 507.87 0.49 261.09 0.39 0.19 0.48 0.50

Min.
2.93 -3.10 0.00 3.07 -3.11 0.00 0.00 2.67 -3.22 1.30 -2.44 0.00 0.00 3.12 -3.05 0.00 0.00 0.00 0.00 1.67 -3.87 1.67 -2.76 0.00 0.00 0.00 0.00 0.00 0.00 116.00 0.00 82.00 0.30 0.00 0.00 0.00

Max.
4.93 2.11 1.00 4.80 2.27 1.00 1.00 4.94 2.32 4.13 2.95 1.00 1.00 4.86 1.98 1.00 1.00 1.00 1.00 5.00 1.79 4.67 2.05 1.00 1.00 1.00 1.00 1.00 1.00 3187.00 1.00 1422.00 4.30 1.00 1.00 1.00

92

Clinic Descriptive Statistics Patient Survey - Level three


Although patients were selected for participation in the survey from the 29 group clinics and eight IPA affiliates, survey respondents were linked to only 25 group clinics and eight IPA affiliates. Moreover, one of the clinics refused to participate in the in-person interviews to evaluate the clinics Chronic Care Model components for the year 2007. As previously mentioned IPA affiliates did not participate in the assessment of CCM components therefore CCM clinic data were available only for 24 sites (Table 23).

The distributions for the scores of the overall CCM component and individual CCM elements were left skewed with the majority of clinics scoring toward the top of the range. Consistent with prior research on the CCM however, part two of the model, linkages with existing community resources scored among the lowest of all components (mean = 5.1 on an 11-point scale). The recently added component of integration that assesses the degree to which all components of the CCM work together also scored poorly. The overall mean score for the last sub-component of part three, clinic information system, was 8.4, the highest mean score of any CCM element. The actual days of the availability of electronic health records ranged from 88 to 577 days indicating that some clinics did not have access to electronic health records for the entire year of 2007.

93

Table 23 Clinic Descriptive Statistics linked to Patient Survey Respondents Variable Label N Mean StdDev Min CCMF CCMF Composite overall score 24 6.68 1.01 4.27 CCMPt1 CCM Part 1- leadership 24 7.91 1.51 4.17 CCMPt2 CCM Part 2- community 24 5.08 1.11 3.33 linkages CCMPt3 CCM average provider practice 24 7.30 1.14 4.64 CCMPt3a CCM self-mgt resources 24 6.25 1.32 3.75 CCMPt3b CCM decision-support 24 6.92 1.79 3.25 CCMPt3c CCM delivery system design 24 7.65 1.44 4.50 CCMPt3d CCM clinic info system 24 8.38 1.10 5.80 CCMPt4 Integration of all CCM elements 24 4.59 1.24 1.50 AvgCCEHR Sum of Pts 3b,3c,&3d 24 22.94 3.60 14.55 AvgVis Total visits in 2007 24 3160.37 421.75 2332.73 MinSc Ratio of minority pts (0-10) 24 5.58 2.36 2.00 CMHE On-site HE or care mgt 24 0.29 0.46 0.00 EHRdays Days of electronic health records 24 361.96 151.17 88.00 zCCMF z-transformed overall CCM score 24 -0.00 1.00 -2.39 zCCMPt1 z-transformed Part 1 score 24 -0.00 1.00 -2.48 zCCMPt2 z-transformed Part 2 score 24 0.00 1.00` -1.57 zCCMPt3 z-transformed Part 3 score 24 0.00 1.00 -2.34 zCCMPt4 z-transformed Part 4 score 24 0.00 1.00 -2.50 zEHRdays z-transformed EHR days 24 0.00 1.00 -1.81

Max 8.66 9.67 7.00 10.00 9.25 10.00 10.33 10.40 7.67 30.73 3885.00 10.00 1.00 577.00 1.97 1.17 1.72 2.37 2.49 1.42

Descriptive Statistics PCP Survey Responses PCP Survey Descriptive Statistics


Key variables from the PCP survey were intended to be included in the patient survey analyses however the reduction in sample size resulting from matching PCPs to survey respondents prohibited this approach. Patients affiliated with IPAs were included in the sampling frame for the patient survey but IPA PCPs were excluded from the PCP survey contributing to the reduction in potential sample with both PCP and patient survey data. A total of 91 PCPs completed the PCP survey however only 83 PCPs linked to patients in the study population for an overall response rate of 74 percent. The PCP survey descriptive results are presented in Table 24.

94

The physicians reported wide variation in the key domains of interest care coordination activities, end-of-life discussions, and medication management. The mean score for the PCP care coordination domain was 3.7 (minimum of 2.29 and maximum of 5.57 on a sixpoint scale) that included eight questions assessing care coordination behaviors such as referral to health education or community resources, use of physician patient intervention reports, following up with specialty care provided on assigned patients, etc. PCPs reported the highest compliance with medication management with a mean of 4.7 on a 6point scale. Similar variation was found within the transfer of information domain, a frequently cited barrier to optimal care coordination within primary care.

Twenty-two percent responded that low health literacy was frequently a barrier to meeting patient goals and six percent identified language discordance as a barrier. However, only seven percent of physicians reported frequently or always assessing patient recall of the treatment as part of the visit.

The majority of PCPs were male (65 percent) and the over 80 percent were between the ages of 35 and 60 years of age; 12 percent had worked for the organization fewer than three years and six percent had practiced medicine for less than five years. Thirty-six percent of the PCPs were family practitioners and the average number of assigned patients that were identified as part of the study population was 46.
Table 24. PCP Descriptive Statistics PCP Survey

Variable
AdeTim AgeGT60 AgeLE35

Label
Adequate time for care coordination PCP age >60 years PCP age < 35 years

N
83 83 83

Mean
2.31 0.12 0.07

StDev
1.20 0.33 0.26

Min
1.00 0.00 0.00

Max
5.00 1.00 1.00

95

Variable
AssHbts AssLivSt CMHigh CMHthPn CarGivH Coord CMCC DMRD EOL FP HAssDep HAssRcll HCPNew HlthLitB HthPnDM HospCC InadTrn InfoTrns Lang DC MAssRcll MDLE5 Medsmgt MinPts PADO PCPSex Teamcpt Totpts zCoord zEOL zMedsMgt zInfTrns

Label
Assess all health habits diet, exercise, Assess pt living situation > 20% of patients in Care Mgt CM done by health plans is helpful Works with caregiver > 20% of time Avg of 8-question coordination domain Care mgt helps PCP care coordination He or RD provides feedback to PCP Avg of 2-question domain (end-of-life) Family practice PCP Assesses for depression frequently Assesses pt recall almost always Less than 3yrs with organization Health literacy frequently a barrier DM done by health plans is helpful Hospital impacts my ability to CC Inadequate training & support Avg of 5-question Info Transfer domain Language discordance often barrier Assesses patient recall frequently In-practice less than five years Avg of 4-question medication domain Majority of patients are minority grps Physician Assistant or DO Male physician Sees patients and self as a team Total patients in study sample z-transformed coordination score z-transformed end-of-life score z-transformed medication mgt score z-transformed info transfer score

N
83 83 83 82 83 83 82 82 83 83 83 83 83 83 80 83 83 83 83 83 83 83 83 83 83 83 83 83 83 83 83

Mean
4.47 4.08 0.13 3.04 0.18 3.70 3.87 3.63 3.91 0.36 0.18 0.07 0.12 0.22 3.01 3.76 2.78 4.16 0.08 0.39 0.06 4.69 0.23 0.12 0.65 0.35 46.45 0.14 0.97 0.45 0.12

StDev
0.86 0.91 0.34 0.91 0.39 0.72 0.73 1.07 0.83 0.48 0.39 0.26 0.33 0.41 0.68 1.12 1.12 0.61 0.28 0.49 0.24 0.65 0.42 0.33 0.48 0.48 33.72 0.35 0.09 0.50 0.32

Min
3.00 2.00 0.00 1.00 0.00 2.29 2.00 1.00 2.00 0.00 0.00 0.00 0.00 0.00 2.00 1.00 1.00 2.80 0.00 0.00 0.00 2.75 0.00 0.00 0.00 0.00 3.00 0.00 0.50 0.00 0.00

Max
6.00 6.00 1.00 5.00 1.00 5.57 5.00 5.00 5.33 1.00 1.00 1.00 1.00 1.00 5.00 5.00 5.00 5.60 1.00 1.00 1.00 6.00 1.00 1.00 1.00 1.00 210.00 1.00 1.00 1.00 1.00

96

Chapter 5: Bivariate Relationships Longitudinal and Cross-sectional Samples


Simple pooled sample bivariate statistics comparing level one, two and three covariates with each respective dichotomous dependent variable were conducted to identify potential covariates for inclusion in the Hierarchical Generalized Linear Models (HGLM) with the longitudinal data. Statistical significance may have been overestimated in these relationships as dependent variables, and patient and physician variables are not independent observations. Bivariate statistics comparing survey data responses at the clinic, PCP, and patient level with outcome variables were also conducted as an exploratory step prior to fitting the cross-sectional HGLMs. Only significant findings are reported for each quality or adherence measure.

Results of Bivariate Associations Longitudinal Sample


Two dependent patient outcomes across the fours years of the study period annual diabetes screens (A1c and LDL tests) and annual DM screens with A1c and LDL control and the absence of acute utilization (no ACSC hospitalizations and no ED visits for CHF, COPD, diabetes, asthma and hypertension) were compared with key PCP and patient predictors and covariates using Student t-test for continuous variables and Pearsons chisquare tests for categorical variables. Significant results as compared with either dependent variable are reported in Table 25.

97

Patient continuity with the PCP as a time-varying covariate at level-one (same PCP over the calendar year) was significant at the 0.05 level for the diabetes screening measure however PCP continuity over the entire study period was not significantly associated with either diabetes measure. The key PCP predictors of access, communication and coordination were significant at the 0.01 level as associated with the diabetes screen measure however, only coordination was significant at the 0.01 level in the diabetes screen measure combined with A1c and LDL control and the absence of acute utilization. PCP communication was significantly associated with this diabetes measure at the 0.05 level and PCP access was not significant. The patient level covariates for baseline A1c and LDL results and the average A1c and LDL results for PCP patient panels were all significantly associated with the diabetes screen with control and the absence of acute utilization measure.

Table 25. Bivariate Statistics: Diabetes Screens, and Diabetes Screens with A1c and LDL control and No Acute Utilization 2004 through 2007 (significant findings) DMScreens Overall Percent n=19,905 72.1 27.9 Yes No 1.7 [1.1] 6.8 [4.3]* 5.9 [6.0]* 0.5 [1.3]* 0.2 [0.9]
DMScreens/Control No Acute Utilization

n=16,999 42.1 Yes 57.9 No

Variable
Time TPCVis TSpVis TEndoVis THE

Label
Number of eligible years Total # PC visits per year Total # Specialist visits/yr Total # Endo visits/yr Total # health education/yr

Level 1-repeated measures Mean [standard deviation] 1.4 [1.1] 1.5 [1.1]* 1.7 [1.1] 5.9 [4.6] 6.8 [4.1]* 6.5 [4.6] 5.1 [5.6] 5.9 [5.7]* 5.6 [ 6.0] 0.1 [0.6] NS 0.1 [0.4] NS % Observations NS 93.0 92.1 Mean [standard deviation]

Pt linked to same PCP each yr PCPCS Level 2-pt variables

significant at p-value < .01

98

DMScreens Overall Percent n=19,905 72.1 27.9 Yes No 76.1 [6.9] 74.1 [6.4] 9.5 [4.1]* 8.3 [4.0] 6.7 [3.4]* 6.2 [3.5] 5.3 [4.7] 5.8 [5.0] NA NA 15.7 [6.5]* 15.2 [6.9] 37.6 [7.2]* 2.69 [1.4] 25.9 12.2 * 8.8 * 8.1 * 4.1 * 87.6 * 58.7 NS 9.3 * NS 46.7 NS 60.8 * 5.7 * 58.3 7.4 44.8 11.0

DMScreens/Control No Acute Utilization

Variable
Ages AvgMeds AvgPCVis AvgSpVis FirstA1c FirstLDL TotChrDz PhyMos TotRAF ADDA

Label
Age as of January 1, 2004
Avg total med classifications/yr

Avg primary care visits/yr Average visits with specialist/yr Baseline A1c result Baseline LDL result Chronic conditions-observable in claims Total months with linked PCP Overall risk factor for 2007

n=16,999 42.1 57.9 Yes No 74.0 [6.4]* 74.3 [6.7] 9.5 [4.0]* 9.2 [4.0] 6.7 [3.3]* 6.6 [3.3] 5.8 [4.8]* 5.5 [4.9] 6.9 [1.1]* 7.2 [1.5] 100.1 [28.6]* 114.5 [33.4] NS 37.1 [7.2] 2.7 [1.5] 27.6 10.0 7.7 9.2 5.1 86.7 NS 9.1 * 8.7 * 22.1 * 48.9 * 22.7 61.8 4.8 * 10.9 10.5 25.8 43.8 24.0 60.0 6.4

Alcohol drugs depression or anxiety AnyEndo Any visits to endocrinologist AnyHE Any visits to health education Dementia Dementia FPL200 Median household inc < $24,000 Group Patient in group model care HCostDz End-stage renal disease, CHF, COPD and/or emphysema MM Medicaid coverage in 2007 Pov10 Median household inc < $27,471 Pov25 Median household inc < $34,824 PtSex Male SCAN SCAN insurance plan SexConc Patient sex matches PCP sex Termed Patient left the group in 2008 Level 3-PCP variables PredACS PredCM PredCD
PredCMCD

38.0 [7.0] 37.6 [7.3]* 2.7 [1.5] 2.6 [1.4]* % Observations 29.1 25.2 * 4.2 3.8 5.7 84.9 57.0 11.8 * 8.5 7.6 * 3.3 * 87.8

AvgAge Avg1A1c Avg1LDL AvgPtMed


PCP access score PCP communication score PCP coordination score Sum of Comm & Coord Avg age of linked pts Avg A1c of linked pts Avg LDL of linked pts Avg # med classes among linked pts

Mean [standard deviation] 0.01 [0.7] NS 0.06 [0.7] -0.09 [0.7] -0.03 [0.7]* -0.03 [0.7] 0.06 [0.6]* 0.0 [.6] 0.07 [0.6]* -0.09 [1.3] 0.03 [1.3]* 0.03 [1.3]* NS 74.2 [1.8] 74.5 [1.7] NA 7.05 [0.4]* NA 107.0 [7.9]* 9.3 [1.2]* 9.1 [1.2] 9.4 [1.3]*

-.06 [0.7] 0.03 [0.7] -0.02 [1.3] 7.1 [0.4] 108.3 [7.9] 9.3 [1.2]

significant at p-value < .01 significant at p-value < .05

99

DMScreens Overall Percent n=19,905 72.1 27.9 Yes No NS NS 677.6 [232.8]* NS 0.27 [0.1]* 660.5 [244.9]

DMScreens/Control No Acute Utilization

Variable
AvgRsk AvgTRAF AvRef AdddaPts HAddaPts HCostPts Panel RatMM RatPov10 RatPov25 HPredACS HPredCM HPredCD
HPDCMCD

Label
Average HCC of linked patients Average overall risk linked pts Average annual referrals issued Ratio of dementia and pts below/study pts Ratio of pts - anxiety, dep. or subst abuse/study pts Ratio of high cost pts/study pts Average Pt Panel or Caseload Ratio of Medicaid pts/study pts #10% poverty pts/study pts #25% poverty pts/study pts PCP Access score > 0 PCP Commun. score > 0 PCP Coordination score > 0 Sum commun & coord >0 Geriatric training or geriatrician Foreign-trained PCP Internal Medicine Independent Practice Assoc In practice less than 10 years Male PCP PCP speaks Spanish Specialty beyond IM or FP In practice for ten to 20 years In practice greater than 20 years PCPSex*Lt10Yrs

n=16,999 42.1 57.9 Yes No 2.02 [0.3] 2.01 [0.3] 2.69 [0.4] 2.67 [0.3] 680.1 [233.4]* 0.355 [.1] 0.27 [0.1]* 670.0 [236.3] 0.351 [.1] 0.26 [0.1] 0.56 [0.2] 1210.0 [444.5] 0.11 [0.1] 10.2 [.2] 0.25 [0.3] 50.1 46.7 51.3 38.4 4.1 33.8 74.0 21.7 29.7 3.2 34.0 10.3

Geri FTr IM IPA LT10yrs PCPSex Sp Spec Yrs10_20 Yrs_GE20 PCPSexNw

0.58 [0.2]* 0.56 [0.1] 0.59 [0.1]* 1225.6 1190.1 1230.8 [440.3]* [456.1] [444.2]* 0.10 [0.1]* 0.11 [0.1] 0.10 [0.1]* NS .09 [.2] 0.24 [0.2]* 0.25 [0.3] 0.23 [0.2]* % Observations 48.8 51.7 51.3 49.0* 41.7 49.5 * 53.9* 49.6 55.0 * 40.5 * 33.9 40.8 * 4.5 3.4 3.8 35.4 * 32.3 36.4 * 75.3 69.8 * 71.2 11.9* 14.6 NS 22.9 * 19.3 23.7 * 64.8 * 69.4 NS NS 27.2 * 2.8 * 3.7 2.6 NS 32.3 44.3 * 48.3 NS 10.8 * 9.5 11.2

The first diabetes composite measure (DMC) included three components, evidence of an annual A1c and LDL test, and a colorectal screening in accordance with HEDIS specifications. The second diabetes composite measure (DMCUT) included five components, evidence of DMC and the absence of an ACSC hospitalization or an ED visit for chronic illness. Bivariate results of the significant relationships between the diabetes composite measures in the longitudinal data sample and key predictors and

100

covariates at level one, two, and three are depicted in Table 26. Again, only significant results for at least one of the dependent variables are reported.

Physician continuity as a time-varying covariate at level one was not found to be significant, however continuous patient time with the same PCP over the entire period of the study, a level-two key predictor was found to be significant at the 0.01 level in both models DMC and DMCUT. The key predictors of PCP communication and the sum of predicted score for communication and coordination as continuous variables were significant at the 0.01 level in both models, and coordination was significant at the 0.05 level in the composite measure that combined utilization (DMCUT). Similarly, the dichotomous variables for above average scores for communication and communication combined with coordination were both significant at the 0.01 level in both models. In all cases higher levels of physician communication and communication combined with coordination were associated with better outcomes on these diabetes composite measures.
Table 26. Bivariate Statistics: Diabetes Composite (A1c, LDL, CR screens) and No Acute Utilization 2004 through 2007 (significant findings) DMC/ No Acute Utilization DMC n=15,013 n=15,013 28.6 (4297) 26.6 (3994) Overall Percent Yes No Yes No Variable Label Level 1-repeated measures Mean [standard deviation] Time Number of eligible years 1.6 [1.1] 1.8 [1.1]* 1.6 [1.1] 1.8[1.1] TPCVis Total # PC visits per year 7.2 [4.7]* 6.3 [4.3] 6.9 [4.4]* 6.4 [4.3] TSpVis Total # Specialist visits/yr 6.4 [6.4]* 5.3 [5.7] 6.2 [6.2]* 5.4 [5.8] TEndoVis Total # Endocrinologist vis/yr 0.5 [1.4]* 0.4 [1.1] 0.5 [1.4]* 0.4 [1.2] Level 2-pt variables Mean [standard deviation] Ages Age as of January 1, 2004 72.7 [4.8] 71.7 [5.0]* 72.7 [4.9] 71.8 [4.9]
significant at p value < .01 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables)

101

Overall Percent

Variable
AvgMeds AvgPCVis AvgSpVis AvgOther TotChrDz TotalMos TotRAF ADDA AnyEndo AnyHE CtPCP CtSite Dementia FPL200

Label
Avg med classes over all yrs Avg annual primary care visits Avg annual visits with specialists Average annual other services Chronic conditions-observable in claims Total months in study Overall risk factor for 2007

DMC n=15,013 28.6 (4297) Yes No 9.8 [4.3]* 9.0 [4.1] 7.0 [3.6]* 6.4 [3.4] 6.3 [5.3]* 5.5 [4.8] 3.2 [4.2]* 15.9 [6.6]* 45.7 [5.3]* 2.59 [1.4]* 25.8 12.9 * 9.8 * 36.6 * 89.9 * 6.0 * 4.7 48.6 NS 24.1 * 63.0 * 4.3 * -0.01 [0.7]

DMC/ No Acute Utilization

n=15,013 26.6 (3994) Yes No 9.6 [4.2]* 9.1 [4.2] 6.9 [3.5]* 6.4 [3.5] 6.1 [4.8]* 5.5 [4.9] 3.1 [2.9] 15.5 [6.3] 2.9 [3.9] 15.3 [6.7] 46.4 [4.5] 2.7 [1.5] 27.4 10.6 7.5 32.0 87.2 7.6 4.0 46.7 23.1 21.8 60.0 5.8

2.9 [3.9] 15.1 [6.6]

Alcohol drugs dep or anxiety Any visits to endocrinologist Any visits to health education All time with same PCP All time at same clinic/site Dementia Median household income < $24,000 PtSex Male Pov25 Median household income < $34,824 SCAN SCAN insurance plan SexConc Patient sex matches PCP sex Termed Patient left the group in 2008 Level 3-PCP variables PredCM PredCD
PredCMCD

46.4 [4.4] 45.7 [5.3] 2.62 [1.5] 2.5 [1.4] % Observations 27.4 25.6 10.4 12.6 * 7.3 9.4 * 31.9 36.6 * 87.2 89.1 * 7.5 5.6 * 3.9 4.6 * 46.7 21.6 60.1 5.7 48.6 24.8 23.7 63.4 * 3.9 *

PCP communication score PCP coordination score Sum of Comm & Coord Average age of linked patients Avg HCC of linked patients Avg overall risk linked pts Avg # chronic illnesses among linked pts Avg # med classes, linked pts Average pt panel/caseload Ratio of Medicaid pts/study pts

Mean [standard deviation] -0.05 [0.7] 0.0 [0.7]* -0.05 [0.7] 0.07 [0.6] 0.06 [1.3]* 74.1 [1.8]* 2.0 [0.3]* 2.66 [0.4]* 15.3 [1.8]* 9.4 [1.4]* 1248 [444]* 0.1 [.1]* 0.05 [0.6] 0.00 [1.2] 74.2 [1.7] 2.02 [0.3] 2.69 [0.3] 15.2 [1.8] 9.2 [1.2] 1225.0 [435.7] 0.099 [.1]

AvgAge AvgRsk AvgTRAF AvgPtCdz AvgPtMed Panel RatMM

NS 0.06 [1.3]* 0.00 [1.2] 74.1 [1.8] * 74.2 [1.7] 2.00 [0.3]* 2.02 [0.3] 2.69 [0.4]* 2.66 [0.4] 15.3 [1.8]* 15.2 [1.8] 9.4 [1.4]* 9.2 [1.2] NS 0.1 [.1]* 0.099 [.1]

significant at p value < .01 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables) significant at p value < .01 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables)

102

Overall Percent

Variable
RefRat HPredCM
HPDCMCD

Label
Ratio of avreferrals/caseload PCP Commun. score > 0 Sum commun & coord >0 Geriatric training or geriatrician Foreign-trained PCP Internal Medicine PCP speaks Spanish

Geri FTr IM Sp

DMC/ No Acute Utilization DMC n=15,013 n=15,013 28.6 (4297) 26.6 (3994) Yes No Yes No 0.6 [0.1]* 0.6 [0.1] 0.6 [0.1]* 0.6 [0.1]* % Observations 50.3 * 47.1 50.5 * 47.1 59.2* 54.3 59.5 * 54.4 2.7 * 4.2 2.6 * 4.2

38.5 * 66.2 26.2 *

33.2 72.7 29.5

38.7 * 66.1 * 26.5 *

33.3 72.6 29.4

Few significant results were found in the bivariate results comparing the diabetes medication adherence measure in the longitudinal data sample with key predictors and covariates at level one, two, and three. As this finding was contrary to study hypotheses, an adherence composite outcome measure was created and examined that combined the medication adherence measure with the absence of ACSC hospitalization and ED visit for chronic illness. As the goal of medication adherence is to manage chronic illness and avoid adverse events necessitating hospitalization or an ED visit the adherence composite constituted the clinically meaningful outcome. Additionally, a second dataset was created that removed PCPs with fewer than 10 linked patients with the intention of improving the ability to find a significant relationship at the PCP-level if one existed; patient sample was reduced by 8.9 percent. Again, few significant associations were found in the bivariate relationships in the full or restricted patient/PCP sample. Physician continuity throughout the entire period of medication utilization was significantly associated with the medication adherence measure composite at the 0.01 level but only in the large data sample. PCP predicted coordination score as a dichotomous variable was

103

significant at the 0.05 level in the large data sample. Neither variable was significant in the restricted data sample. All significant associations are reported in Table 27.
Table 27. Bivariate Statistics: Medication Adherence to Oral Diabetes and No Acute Utilization 2004 through 2007 and Adherence only (significant findings) DMRX Adherence/No
acute utilization

DMRX Adherence
(at least 10 pts/PCP)

Overall Percent

Variable
Level 1-repeated measures

Label

n=12,159 76.8 23.3 Yes No 1.6 [1.1] 0.86 [0.3]* 9.9 [3.9]*

n=10,854 83.5 (9064) Yes No

Time Number of eligible years GenRat Ratio of generic/total meds/yr TotMeds Total med classes per year Level 2-pt variables Ages AvgMeds AvgPCVis AvgOther TotChrDz PhyMos TotRAF ADDA AnyBhlth AnyEndo CtPCP CtSite Dementia FPL200 HCostDz MM Pov25 SexConc Termed

Mean [standard deviation] 1.7 [1.1] 1.6 [1.1] 1.7 [1.1] .87 [0.3] 0.85 [.3]* 0.87 [.3] 10.5 [5.6] 9.7 [4.6]* 10.1 [5.4 74.1 [6.7] 9.8 [4.4] 2.8 [4.3] 15.8 [6.8] 36.5 [7.0] 3.0 [1.6] 27.7 3.7 10.7 30.4 NS 7.8 * 4.1 * 57.8 * 10.2 * 25.3 * 61.7 * 5.7* 11.1 5.8 61.4 13.0 28.0 58.0 8.5

Age as of January 1, 2004 Average annual med classes Average annual PCP visits Average use of other services Chronic conditions-observable in claims Total months with linked PCP Overall risk factor for 2007 Alcohol drugs depression or anxiety Any visits to behavioral health Any visits to endocrinologist All time with same PCP All time at the same clinic/site Dementia Median household inc < $24,000 End-stage renal disease, CHF, COPD and/or emphysema Medicaid coverage in 2007 Median household inc < $34,824 Patient sex matches PCP sex Patient left the group in 2008

Mean [standard deviation] 73.7 [6.5] 73.8 [6.5] 74.3 [6.7] 9.8 [3.9] 10.0 [4.5] 9.7 [3.9]* 6.6 [3.3]* 6.8 [3.4] NS 2.8 [3.8] 3.0 [4.3] 2.7 [3.8] 14.9 [6.2]* 16.2 [6.9] 14.8 [6.3]* 37.6 [7.3]* 2.6 [1.3]* 24.2 2.6 12.1 35.5* 88.0 8.0* 3.9 57.0* 10.0 * 24.5 * 61.8 5.8 * 36.9 [7.2] 37.5 [7.1]* 3.1 [1.7] 2.6 [1.3]* % Observations 27.5 24.4 3.4 10.6 32.1 86.4 11.5 5.4 61.2 12.7 28.0 59.2 8.6 2.7 12.1 34.8 *

significant at p-value < .01 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables) significant at p-value < .05 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables)

104

DMRX Adherence/No
acute utilization

DMRX Adherence
(at least 10 pts/PCP)

Overall Percent

Variable
Level 3-PCP variables HCostPts RefRat HPredCD IM Lt10Yrs Sp

Label
Ratio of high cost pts/study pts

n=12,159 76.8 23.3 Yes No

n=10,854 83.5 (9064) Yes No

Ratio of AvRef/pt panel PCP Coord score >0 Board-certified Internal Medicine In-practice less than 10 yrs PCP speaks Spanish

0.58 [0.2] * 0.57 [0.1]*

Mean [standard deviation] 0.57 [0.2] 0.58 [0.15]* 0.57 [0.2] 0.56 [0.1] 0.57 [0.12]* 0.56 [0.1] % Observations (n) 49.7 NS 52.1 * NS 73.5 76.0 21.8 24.0 NS 29.4* 31.5

The initial findings of bivariate relationships between Ace/Arb medication adherence measure and key predictors at level-one, two and three were similar to those found with the oral diabetes medication adherence measure. Again, a medication adherence composite measure was created and examined and a second dataset was created that retained only PCPs linked to at least nine patients (a larger reduction in the patient sample of 21.5 percent). Few significant findings of association were found. Continuity with the same PCP over the entire period of prescribing was significantly associated with Ace/Arb adherence composite measure in both samples of patients at the 0.01 level. PCP predicted communication score as a dichotomous variable was significant at the 0.05 level in the smaller patient sample only. All significant findings are reported in Table 28.

105

Table 28. Bivariate Statistics: Medication Adherence to Ace Inhibitors and Arb Medications and No Acute Utilization 2004 through 2007 (significant findings) Ace/ArbRX
adherence/No acute utilization

Ace/ArbRX Adherence
(at least 9 pts/PCP)

Overall Percent

Variable
Time TotMeds

Label
Number of eligible years Total med classes/year

n=12,846 77.0 23.0 Yes No

n=10,083 83.9 (8459) Yes No

Level 1-repeated measures Mean [standard deviation] 1.7 [1.1] 1.6 [1.7] 1.7 [1.1]* 1.5 [1.1] 10.1 9.5 [4.7] 10.0 [4.5] 10.5 [4.9] [4.9]* NS 0.76 [.4]* 0.80 [.4] Mean [standard deviation] 74.3 [6.5]* 74.8 [6.8] 74.4[6.5] 9.8 [4.0]* 10.3 [4.4] 9.9 [4.2]* 6.7 [3.3]* 7.0 [3.5] 6.8 [3.4] 5.7 [4.8]* 5.9 [5.1] 5.9 [4.9]* NS 15.5 [6.2]* 17.0 [7.1] 2.6 [1.4]* 37.6 [7.2]* 3.2 [1.6] 37.0 [7.3] 3.0 [4.1]* NS NS 36.9 [6.9] 29.0 28.1 NS 8.0 * 4.3 * 89.0 * 66.2 12.0 11.2 27.1 44.1 10.0 * 24.3 * NS NS NS 12.5 29.3 10.7 6.1 92.1 74.7 [6.8] 9.4 [4.0] 6.6 [3.3] 5.3 [4.6] 2.4 [4.0]

GenRat Ratio of generic/total meds/yr Level 2-pt variables Ages AvgMeds AvgPCVis AvgSpVis AvgOther TotChrDz TotRAF PhyMos ADDA CtPCP CtSite Dementia FPL200 Group HCostDz MM Pov10 Pov25 PtSex

Age as of January 1, 2004 Average med classifications Average annual PCP visits Average annual Specialist visits Average use of other services Chronic conditions-observable in claims Overall risk factor for 2007 Total months with linked PCP Alcohol drugs depression or anxiety All time with same PCP All time at same clinic/site Dementia Median household inc < $24,000 Patient is assigned to group PCP End-stage renal disease, CHF, COPD and/or emphysema Medicaid coverage in 2007 Median household inc < $24,471 Median household inc < $34,824 Male patient

37.4 [7.1]* % Observations (n) NS 26.5 31.6 86.1 11.8 5.4 NS 33.9 *

35.0 * 87.8 7.4 * 4.2 *

61.0 * 10.1 * 9.7 24.6 * 46.3

significant at p-value < .01 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables) significant at p-value < .05 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables)

106

Ace/ArbRX
adherence/No acute utilization

Ace/ArbRX Adherence
(at least 9 pts/PCP)

Overall Percent

Variable

Label

SCAN SCAN Insurance Plan Termed Patient left the group in 2008 TobDep Claims identified tobacco dep Level 3-PCP variables AvgRsk AvgTRAF ADDDApts Panel RefRat HPredCM IPA Average HCC linked pts Average overall risk linked pts Ratio of Mental Hth pts/study
pts

n=12,846 77.0 23.0 Yes No 24.0 * 26.7 5.1 * 8.5 2.5 * 3.6

n=10,083 83.9 (8459) Yes No NS NS 2.7 3.6

Mean [standard deviation] 2.1 [0.3] 2.0 [0.3]* 2.06 [.4] 2.08 [.4] 2.67 [0.4] 2.69 NS 0.35 [0.1]* 0.36 [0.1] 0.35 [0.1]* 0.36 [0.1] NS NS 1253.5 [421.9]* 0.57 [0.1] 1285.8 0.56 [0.1] 44.2 7.9

Average Pt panel/caseload Ratio of avReferrals/pt caseload PCP comm score > 0

% Observations (n) NS 47.2 NS 10.5

Two key Chronic Care Model components (the on-site presence of health education and/or care management services and the days of electronic medical records) were examined in the group-only patient/PCP sub-population of the longitudinal study sample. Bivariate results of the significant relationships between two diabetes outcome measures (annual DM screens of A1c and LDL, and annual DM screens of A1c and LDL with A1c control and without a ACSC hospitalization or an ED visit for chronic illness) and key predictors and covariates at level one, two, and three are reported in Table 29.

The presence of health education and/or care management services was significantly associated with the diabetes screen measure at the 0.01 level. There was no association found with the diabetes screen combined with A1c control and the absence of acute

significant at p-value < .05 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables) significant at p-value < .01 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables)

107

utilization measure. The days of electronic medical records days were significantly associated with both diabetes (DM) measures at the 0.01 level. Continuity with the PCP as a time-varying annual covariate was significantly associated with the DM screen measure at the 0.05 level, however continuity with the same PCP over the entire study period was not found to be significantly associated with either DM measure. The key predictors of PCP access, communication and coordination were significantly associated with the DM screen measure at the 0.01 level however, only coordination and the sum of coordination and communication were significantly associated with the DM screen measure at the 0.01 level.
Table 29. Bivariate Statistics: Diabetes Screens and DM Screens with A1c control and No Acute Utilization 2004 through 2007 Group Model Only (significant findings) Grp-DMScreens n=17,157 72.7 (12,474) Yes No
Grp-DMScreens/A1c Control No Acute Utilization

Overall Percent

Variable
Level 1-repeated measures Time TPCVis TSpVis TEndoVis THE

Label
Number of eligible years Total # PC visits per year Total # Specialist visits/ yr Total # Endo visits/yr Total # health education/yr

n=13,544 68.2 (9230) Yes No

PCPCS Pt linked to same PCP/year Level 2-pt variables Ages AvgMeds AvgPCVis AvgSpVis Age as of January 1, 2004 Avg # med classes annually Average annual primary care visits Average annual visits with specialists

Mean [standard deviation] 1.4 [1.1] 1.5 [1.1]* 1.8 [1.2] 1.7 [1.1] 6.7 [4.3]* 5.7 [4.6] 6.6 [4.1]* 6.9 [5.0] 5.8 [5.9]* 5.0 [5.5] 5.5 [5.5]* 6.2 [6.5] 0.5 [1.4]* 0.1 [0.6] 0.4 [1.2]* 0.6 [1.6] 0.04 [0.3] 0.12 [0.7]* 0.16 [0.9] 0.13 [0.8] % Observations (n) NS 93.0 92.0 Mean [standard deviation] 76.0 [6.9] NS 74.0 [6.5] 9.4 [4.1]* 8.1 [4.0] 9.0 [3.8]* 9.8 [4.3] 6.6 [3.4]* 61. [3.4] 6.5 [3.3] 6.6 [3.6] 5.8 [4.9]* 5.1 [4.5] 5.5 [4.7]* 5.9 [5.3]

significant at p value < .01 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables) significant at p value < .05 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables)

108

Grp-DMScreens n=17,157 72.7 (12,474) Yes No NA 14.9 [6.6] 15.4 [6.5] 2.68 [1.4] 26.3 13.1 * 9.2 * 8.0 * 4.6 * NS NS 10.4 * 25.4 * NS 61.4 * 5.5 * 3.0 58.7 7.6 3.7 12.5 27.6

Grp-DMScreens/A1c Control No Acute Utilization

Overall Percent

Variable
FirstA1c TotChrDz TotRAF ADDA AnyEndo AnyHE Dementia FPL200 HCostDz MM Pov10

Label
Baseline A1c result Chronic conditionsobservable in claims Overall risk factor for 2007 Alcohol drugs depression or anxiety Any visits to endocrinologist Any visits to health education Dementia Median household inc < $24,000
End-stage renal disease, CHF, COPD and/or emphysema

n=13,544 68.2 (9230) Yes No 6.9 [1.1]* 7.7 [1.7] 15.1 [6.2]* 15.7 [7.0]

2.74 [1.5] 2.6 [1.3]* 3.0 [1.6] % Observations 29.5 NS 4.6 4.1 12.0 6.4 11.3 * 8.4* 7.7 * 4.1 * 58.7 * 9.8 * 10.1* 24.6 * 21.4* 62.2* 4.5* NS 16.3 10.5 10.0 6.2 63.0 12.4 11.6 28.0 23.8 59.0 8.0

Medicaid coverage in 2007 Median household inc < $27,471 Pov25 Median household inc < $34,824 SCAN SCAN insurance plan SexConc Patient sex matches PCP sex Termed Patient left the group in 2008 TobDep Tobacco dependence - claims Level 3-PCP variables PredACS PredCM PredCD PredCMCD AvgAge AvgRsk AvgTRAF AvgPtMed EHRDays HCostPts RatMM

PCP Access score PCP Communication score PCP Coordination score


Sum Comm & Coord

0.06 [0.7] -0.1 [0.8]* 0.01 [0.7]* -0.1 [1.4]* 74.2 [1.7]*

Mean [standard deviation] -0.04 [.7] NS -0.2 [0.8] -0.12 [0.8] -0.17 [.8]

Avg age of linked patients Avg HCC of linked patients Avg overall risk linked pts
Average # med classes linked pts

-0.10 [0.7] -0.29 [1.4] 74.5 [1.6] NS 2.7 [0.3] 9.2 [1.1] 376.7 [130.5] 0.58 [0.1] 0.11 [0.1]

0.02 [0.7]* -0.1 [1.4]* 74.2 [1.7]* NS NS NS 388.3 [134.8]* 0.60 [0.1]* 0.11 [0.1]*

-0.05 [0.7] -0.2 [1.4] 74.3 [1.6]

2.69 [0.4]* 9.3 [1.1]* 386.4 [134.2]* 0.60 [0.1]* 0.11 [0.1]*

Days EHR available to PCP Ratio of high cost pts/study pts Ratio of Medicaid pts/study pts

376.5 [134.4] 0.58 [0.1] 0.12 [0.1]

significant at p value < .01


significant at p-value < .05 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables)

109

Grp-DMScreens n=17,157 72.7 (12,474) Yes No 0.25 [0.2]* 0.27 [0.2] 0.32 [0.1] 0.2 [0.9]* 39.6 * 48.1 45.6 * 49.7 * 34.5 * 4.0 * 36.4 * 74.4 20.2 * 65.3 * 45.1 * 0.31 [0.1]

Grp-DMScreens/A1c Control No Acute Utilization

Overall Percent

Variable
RatPov25 ProMHPts ZEHRDays CM/HE HPredACS HPredCM HPredCD HPDCMCD Geri FTr IM LT10yrs PCPSex Yrs_GE20

Label
Ratio of pts at 25% pov/study pts Ratio of mental health pts/study pts

n=13,544 68.2 (9230) Yes No 0.25 [0.2]* 0.27 [0.3] 0.32 [0.1]* 0.31 [0.1]

Z-transformed EHR Days HE or Care mgt on-site Access score > 0 Commun. score > 0 Coordination score > 0 Sum Commun & Coord >0 Geriatric training or geriatrician Foreign-trained PCP Internal Medicine In practice less than 10 years Male PCP In practice > 20 years

0.1 [0.9] 0.20 [0.9]* 0.14 [0.9] % Observations 42.8 40.0 * 42.3 43.1 NS 39.0 45.3 * 42.0 43.8 49.9 * 45.7 28.8 34.0 32.0 5.2 NS 31.5 80.1 17.3 70.6 49.3 36.8 * 73.7 * NS 64.6 * 44.8 67.8 46.8 31.6 76.3

Results of Bivariate Associations Cross-sectional Sample


The key variables of PCP Relational Coordination and Trust were compared with the same diabetes quality and adherence measures as done for the longitudinal data sample. In the first two diabetes (DM) screening measures RC was not significant; PCP trust was marginally significant in the DM screening measure and significant at the 0.05 level in the DM screening with the absence of acute utilization measure. Wagners care coordination index within the Patient Assessment of Chronic Illness Care (PACIC) was significantly associated with both measures at the 0.05 level. Above average scores on all four PCP care coordination indices RC, trust, PCC and PACIC were positively

significant at p value < .01

110

associated with both measures at the 0.05 level. Additionally, the PCP mediation management domain score was significantly associated with both diabetes (DM) measures at the 0.01 level.

Although sixty variables were gathered on the patient survey few were found to be associated with the DM screens measures. The care manager variable indicating frequent contacts from care managers was negatively associated with both measures at the 0.01 level indicative of the high risk status of this patient subset. As few significant associations were found marginally significant relationships at the 0.10 level are also reported (Table 30).
Table 30. Bivariate Statistics: Diabetes Screens and DM Screens and No Acute Utilization Patient Survey Respondents (significant findings) DMScreens Overall Percent n=584 84.4
DMScreens/No Acute Utilization

n=584 78.9 No Yes No

Variable
Level 1-patient Hplcsurv Relig CM Pov 25

Label

Yes

Pt had help filling out survey Moderate or very religious Receives frequent CM calls Median household income < $34,824 Level 2-PCP variables zMDTrust zMDPCC zMDPACIC zmedsrev HighMD HighMD3 IM

9.3 + 77.7 + 9.1 NS

% Observations 15.4 9.1+ 85.7 77.0 17.6 8.2 19.5 +

14.6 86.2 18.7 26.5

z-transformed Trust score z-transformed PCC score z-transformed PACIC score z-transformed Meds Review > avg RC, Trust, PCC, PACIC > avg RC, Trust and PCC Board-certified Internal
Medicine

Mean [standard deviation] 0.04 [0.9]+ -0.4 [1.0] 0.04 [0.9] 0.04 [0.9] -0.3 [0.9] 0.04 [0.9] -0.02 [0.9]+ -0.4[0.9] -0.01 0.9] 0.02 [0.9]* -0.5 [1.0] 0.02 [0.8]* % Observations 18.7 30.8 30.6 23.1 35.1 + 35.1 72.6 + 81.3 71.8

-0.3 [1.0] -0.3 [0.9] -0.3 [.09] -0.3 [1.0] 21.1 26.0 82.1

significant at p-value < .05 (Students t-test for mean comparisons, Pearson chi-square test for categorical variables)

111

significant at the p < 0.01; students t-test for continuous variables and Chi-square test for categorical variables significant at the p < 0.05; students t-test for continuous variables and Chi-square test for categorical variables + significant at the p < 0.10; students t-test for continuous variables and Chi-square test for categorical variables

PCP RC was found to be significantly associated with the diabetes (DM) composite measures at the 0.05 level and at the 0.01 level with the DM composite and the absence of acute utilization measure; trust was not significant. None of the other care coordination indices were significantly associated with the DM composite measures. Education less than eight years was negatively associated with both measures at the 0.01 level. Regular exercise was associated with the DM composite measure and absence of acute utilization at the 0.01 level, likely indicative of lower risk burden and better health as the total risk burden score assigned by the organization is also associated with this measure at the 0.05 level. Significant and marginally significant (p < 0.10) are reported in Table 31.

Table 31. Bivariate Statistics: Diabetes Screening Composite and Composite with No Acute Utilization Patient Survey Respondents (significant findings)

DMC Overall Percent Variable Label Level 1-patient


Agee YrsMD TotRAF EdLt8 EdLtHS Hlpcsurv LowHthLt PoorHlth Pov25 Patient age in 2007 Years seeing current PCP Overall Patient Risk Burden Education < 8 years Education < high school Had help filling out survey Low health literacy SF-12 poor or fair health Median household income < $34,824

n=469 42.4 Yes


74.6 [5.0]+ 7.6 [7.2] NS 6.5 20.1 + 6.5 5.5 + NS 16.1

No
75.6 [5.1] 6.8 [6.3]

DMC/No Acute Utilization n=469 38.4 Yes No


74.5 [5.0] 7.7 [7.2]+ 2.3 [1.2]+ 75.7 [5.1] 6.9 [6.4] 2.5 [1.4] 13.8 27.0 12.1 NS 36.3 25.3

Mean [standard deviation]

% Observations 14.1 6.1 26.7 18.9 12.6 6.7 + 9.6 25.6 25.6 15.6

112

DMC Overall Percent Variable Label


Rexerc Walks or exercises often Level 2-PCP variables zMDRC zMDRCall IM z-transformed MD RC z-transformed RC score Board-certified Internal Medicine

n=469 42.4 Yes


55.8 +

No
47.0

DMC/No Acute Utilization n=469 38.4 Yes No


59.4 45.3

0.05 [0.8] 0.06 [0.8] 67.3

Mean [standard deviation] -0.28 [1.0] 0.05 [0.8] -0.28 [1.0] -0.24 [1.0] 0.08 [0.8] -0.24 [1.0] % Observations 77.4 77.9 65.6

significant at the p < 0.01; students t-test for continuous variables and Chi-square test for categorical variables significant at the p < 0.05; students t-test for continuous variables and Chi-square test for categorical variables + significant at the p < 0.10; students t-test for continuous variables and Chi-square test for categorical variables

PCP RC and trust were not independently associated with the diabetes (DM) screening measure that included A1 control and the absence of acute utilization. The dichotomous variable indicating above average score in RC and trust was significantly associated with this measure at the 0.05. Significant relationships found at the patient level indicated risk burden (total number of chronic conditions, overall patient risk factor, frequent care manager support, had help completing the survey) or socioeconomic barriers (education less than eight years and less then high school, 25th percentile for median household income), and were negatively associated with the measure. Variables found to be significantly associated or marginally significant (p < 0.10) are reported in Table 32.
Table 32. Bivariate Statistics: Diabetes Screens with A1c control and No Acute Utilization Patient Survey Respondents (significant findings) DM Screens/A1c Control no UT N =493 60.2 Yes No Mean [standard deviation] 74.6 [6.2] 76.8 [6.1] 2.4 [2.9] 2.2 [2.6] 2.3 [1.5] 2.6 [1.2]

Overall Percent Variable Label Level 1- patient Agee TotChrDz TotRAF Age in 2007 # Chronic conditions - claims Overall Patient Risk Burden

113

Overall Percent Variable Label Level 1-patient EdLt8 Less than 8 years education EdLtHS Less than HS education Hplcsurv Pt had help filling out survey MinPt Pt is in minority group CM Receives frequent CM calls PoorHlth SF-12 poor or fair - health Pov 25 Median household inc < $34,824 Level 2-PCP variables HMDRCTrt IM >Avg RC and Trust Board-certified Internal Medicine

DM Screens/A1c Control no UT N =493 60.2 Yes No % Observations 8.4 15.3 33.3 18.3 7.4 + 12.6 39.9 50.8 7.4 13.1 31.0 39.9 17.5 26.8 % Observations (n) 39.2 + 31.7 68.5 + 77.1

significant at the p < 0.01; students t-test for continuous variables and Chi-square test for categorical variables significant at the p < 0.05; students t-test for continuous variables and Chi-square test for categorical variables + significant at the p < 0.10; students t-test for continuous variables and Chi-square test for categorical variables

Although it was strongly hypothesized that RC and trust would positively influence medication adherence statistically significant bivariate relationships were not found. In examining the association between the patient and PCP key predictors and covariates, and the medication adherence measures only one significant relationship was identified. Self-reported medication adherence was negatively associated with the pharmacy-claims medication adherence measure for oral diabetic medications. Patients reporting poor medication adherence were significantly less likely be adherent with the pharmacy claims generated medication adherence measure (p = 0.0002) providing further validation of the Morisky-medication scale [173].

The two additional dependent variables constructed from the patient survey data, discussions with PCP on end of life preferences and overall satisfaction with their PCP, were examined relative to the key predictors of RC and trust. RC, trust, PCC, and

114

PACIC were all found to be significantly associated with both measures (p < 0.01). The patient component of the overall RC index and patient self-efficacy were also associated with both measures. Poor health was negatively associated with PCP satisfaction at the 0.01 level. The care manager contact variable was negatively associated with PCP satisfaction and positively associated with discussions about end of life preferences. Living alone without adult children nearby was also positively associated with end of life discussions. This is the only model in which patient self-efficacy was associated with the outcomes, at the 0.01 level in the discussions about end of life care and at the 0.05 level in the total satisfaction with their PCP. Significant variables are reported in Table 33.
Table 33. Bivariate Statistics: End of Life and Total PCP Satisfaction Patient Survey Respondents (significant findings) DiscEOL n=586 25.3 (148) Yes No TotSat n=586 38.7 (227) Yes No

Overall Percent

Variable
Level 1-patient zPtRC SE

Label
z-transformed Pt RC Patient self-efficacy

0.5 [0.7] 8.2 [1.8] 12.8 24.3

Mean [standard deviation] -0.2 [1.0] 0.7 [0.6] -0.4 [1.0] 7.8 [1.8] 8.3 [1.9] 7.6 [1.7] % Observations (n) 7.3 5.7 NS 12.6 25.8 31.1 12.3 NS 22.5 NS 18.1 91.6 + NS NS NS NS 40.4 5.6 86.6

Asian CM CtPCP EdLt8 EdLtHS Latino LvAlone

Pt self-identifies as Asian Frequent calls from care mgt Same PCP all of 2007 Education < 8 years Education < high school Pt self-identifies as Latino
Lives alone no adult children close

6.1 18.2 + 22.3 20.3

PoorHlth SF-12 poor or fair health Level 2-PCP variables zMDRC zMDtrust zMDPCC z-transformed MD RC z-transformed MD trust z-transformed MD PCC 0.1 [0.9] 0.2 [0.9] 0.2 [0.2]

Mean [standard deviation] -0.2 [0.9] 0.2 [0.9] -0.3 [0.8] -0.1 [0.9] 0.3 [0.9] -0.2 [0.9] -0.1 [0.9] 0.3 [0.2] -0.2 [0.9]

115

Overall Percent

Variable
zMDRCall zMDPacic HMDRCTrt HighMDRC HighMD3 HighMD MDassfal MDassdep IM

Label
z-transformed full RC index z-transformed CCM Pacic index High MD RC & Trust MD RC > average > avg RC, trust, PCC > avg RC, trust, PCC, Pacic MD assesses fall risk > 50% pts MD assesses dep > 50% pts Board-certified Internal
Medicine

DiscEOL n=586 25.3 (148) Yes No 0.1 [0.9] -0.1 [0.9] 0.2 [0.9] -0.1 [0.8] 46.0 57.4 45.3 38.5

TotSat n=586 38.7 (227) Yes No 0.2 [0.9] -0.3 [0.9] 0.3 [0.9] -0.2 [0.8] 26.5 38.4 24.2 19.8 25.4 9.2 76.6

% Observations (n) 32.0 49.8 42.9 59.5 29.0 47.1 25.3 42.7 NS 33.9 NS NS 13.7 + 69.6 +

Significant at the p < 0.01; Students t-test for continuous variables and Chi-square test for categorical variables Significant at the p < 0.05; Students t-test for continuous variables and Chi-square test for categorical variables + significant at the p < 0.10; students t-test for continuous variables and Chi-square test for categorical variables

116

Other Bivariate Relationships Explored


As ethnicity/race was not observable in the longitudinal data sample it was not able to be included as a covariate or part of an interaction term in examining the potential differential impact of the key PCP domains of communication and coordination on quality and adherence measures by ethnicity/race. As it was collected on the patient survey self-reported minority status was compared with the key domains of PCP Relational Coordination and Trust and the other care coordination indices. Among the total patient survey sample of 586, almost half of the respondents described themselves as members of minority groups (45 percent). Minority status was negatively associated with lower PCP scores on all care coordination indices RC, trust, PCC and PACIC; negative association with the RC score was significant at the p < 0.0001 level (Students t-test). This association was not found for patient reported self-efficacy. Moreover, these negative bivariate associations between key PCP care coordination domains and AfricanAmerican and Asian patients as distinct groups were not found. The significant negative relationship between PCP RC, trust, PCC and PACIC were found with patients selfidentifying as Latino. The relationships between Latino patients and RC and trust scores were significant and negatively associated for both domains beyond the 0.001 level.

A similar pattern was observed for the bivariate relationship between patients identified as having low health literacy and the PCP key domains of RC, trust, PCC, and PACIC. However, these patients also reported their self-efficacy significantly lower than the rest of the patient survey respondents.

117

Lastly, significant correlations were found between the measured indices of RC and trust in the patient survey sample and the proxy variables used as the key PCP predictor in the longitudinal analyses communication and coordination. As previously reported there was high correlation between the individual variables of RC and trust, and between communication and coordination. A total of 84 PCPs were identified in both the longitudinal analyses and the cross-sectional analyses. An examination of the Pearson and Spearman correlation coefficients was done between the measured RC and trust domain scores and the predicted PCP scores for communication and coordination. All correlations were significant at the 0.01 level. The strongest correlations were found between PCP communication and trust, Pearson correlation coefficient of 0.36 (p = 0.001) and a Spearman correlation coefficient of 0.31 (p = 0.01) and between PCP communication and RC, Pearson correlation coefficient of 0.35 (p = 0.002) and a Spearman correlation coefficient of 0.33 (p = 0.003). Correlations between PCP coordination and trust and PCP coordination and RC were the following; Pearson correlation coefficients of 0.30 (p = 0.007) and 0.27 (p = 0.02), and Spearman correlation coefficients of 0.30 (p = 0.007) and 0.26 (p = 0.02) respectively. Most importantly the combined variable of PCP communication and coordination as a single predictor variable employed in the majority of final models was significantly correlated with RC and trust. The correlation matrix for these relationships is presented in Appendix I.

118

Chapter 6: Hierarchical Generalized Linear Model Regression Results

Tested Research Questions and Hypotheses


The central hypotheses in this study were: 1) four-year longitudinal analyses: PCP care coordination and communication, electronic health records, and care management/health education predict diabetes quality composite measures among seniors with multimorbidity; and 2) 2007 cross-sectional analyses: PCP relational coordination and trust and CCM components predict diabetes quality composite measures and adherence among seniors with multimorbidity.

Hypotheses tested in the 4-year longitudinal analyses include the following: H1-1: Clinic Level clinics with on-site DM/CM and more months of EMR improve the diabetes quality composite measure H1-2: Physician Level - Physician care coordination and communication improve the diabetes quality composite measure and medication adherence measure H1-3: Patient Level - Patient continuity with same PCP improves the diabetes quality composite measure and medication adherence Hypotheses tested in 2007 cross-sectional analyses include the following: H2-1: Clinic Level - Chronic Care Model (overall score) improves the diabetes quality composite measure H2-2: Clinic Level Chronic Care Model moderates physician relational coordination H2-3: Physician Level - PCP relational coordination and trust improve the diabetes quality composite measure, medication adherence measure, end of life discussion, and overall PCP visit satisfaction

119

H2-4: Longitudinal analyses predictors of PCP coordination and communication as proxy variables for the key variables of interest in cross-sectional analyses correlate with cross-sectional analyses predictors of PCP relational coordination and trust.

Longitudinal Analyses
The incorporation of time into HLM models produces additional challenges and opportunities [174]. The researcher must determine the appropriate transformation of time in the model that best answers the research question and reflects the underlying data structure. In a multilevel model with repeated measures at level-one the researcher fits a growth curve model however with a generalized hierarchical linear model (HGLM) with a dichotomous outcome the model being fit is the change in the gap or the probability of the outcome over time. Untransformed time in the model correlates with an interpretation of the intercept as the overall mean at baseline conditional on predictors and random effects. By subtracting the final time period from time, the interpretation of the intercept becomes the overall outcome at the end of a period of study conditional on predictors and random effects. Time transformed to the midpoint is the recommended time transformation for looking at the impact of duration on the intercept [174]. A time transformation to the midpoint correlates with an interpretation of the intercept as the overall outcome accounting for the duration of time conditional on predictors, etc. As the objective of this study was to examine multilevel relational predictors of quality outcomes and adherence, a transformation of time to denote duration was appropriate. Additionally, the choice of time transformation does not alter estimation of variance components or fixed effects parameter estimates; any time transformation merely changes the interpretation of the intercept coefficient [174].

120

In addition to determining the appropriate transformation of time in multilevel models with repeated observations it is important to determine if time has a random variance component in addition to its fixed effect estimate. This model with the inclusion of time with its fixed and/or random effects is referred to as the unconditional model and is equivalent to an empty model in multilevel models without time [90]. This unconditional model was the baseline model for comparison of conditional models containing levelspecific covariates in determining the best fit for the data and the significant predictors of the outcome.

HGLM Diabetes Screening Measure Outcome


In the three-level HGLM examining the log-odds or probability of patient having an A1c and LDL screen for years 2004 through 2007, the results for each level of the analysis are presented as an example of the model building process used subsequently to examine predictors for each dependent binary outcome measure.

Time was found to be a significant positive predictor of the DM screening measure. The predicted values for diabetes screens increased linearly with time however the gap in the probability or log odds from time point to point was most pronounced using time transformed to a midpoint measure as depicted in Figure 5. Therefore, time at the midpoint was used in the final DM screen longitudinal HGLM to interpret the model intercept as the probability or log odds of a patient having an A1c and LDL screen over the duration of their time measurements.

121

Figure 5. Predicted DM Screens as function of time duration

Predicted Annual DM Screens Over Time

This model included a total of 5661 patients linked to 179 PCPs; the majority of PCPs were linked to at least 20 patients (57.4%) in the study sample and 70 percent of patients had information for all four years. PCP and patient sample sizes and cluster size distribution for the Diabetes Screen model are presented in Tables 34 and 35.
Table 34. PCP Sample with Linked Patients Diabetes A1c and LDL Screens HGLM Model
Number patients linked to PCP Number of PCPs % of PCPs % of linked patients with all time points Number of PCPs % of PCPs

10 > 10 - < 20 20 - < 35 35 - <50 > 50

47 29 40 22 41 179

26.3 16.2 22.3 12.2 22.9 100.00

= > 90 = > 75 = > 50 < 50

95 53 19 12 179

53.1 29.6 10.6 6.7 100.00

122

Table 35. Patient Sample Size Repeated Measure Diabetes A1c and LDL Screens HGLM Model
Number of time points observed Number of patients % of patients Cumulative frequency of individuals

1 2 3 4

339 380 957 3985

06.0 6.7 16.9 70.4

339 719 1676 5661

The descriptive statistics for variables examined at level-one, level-two, and level-three for this model are reported in Table 36. This patient sample has been described previously. Ninety-three percent of patients were linked to the same PCP for each 12month period, although only 37 percent were linked to the same PCP for the entire study period. The average number of primary visits for this population was 6.6 per year and visits to endocrinologists averaged less than 0.5 per year. As indicated in the patient covariates reported at level-two the population had significant morbidity.
Table 36. Diabetes Screen Multilevel Descriptive Statistics Descriptive Data for the Variables in the Three-Level Model - DM Screens 2004 through 2007 - with Repeated Observations within Patients nested within PCPs Variable TimeM TPCVis TEndoVis PCPCS DMScreen N Mean Level one Std Min Max

19905 19905 19905 19905 19905 5661 5661 5661 5661 5661 5661 5661 5661

0.12 6.58 0.37 0.93 0.72


Level two

1.10 4.41 1.19 0.26 0.45 0.44 6.78 0.30 0.26 4.17 3.97 3.42 0.48

-1.50 0.00 0.00 0.00 0.00 0.00 62.00 0.00 0.00 0.00 0.00 1.00 0.00

1.50 63.00 14.00 1.00 1.00 1.00 98.00 1.00 1.00 40.50 1.00 50.30 1.00

ADDA Ages AnyEndo AnyHE AvgMeds AvgOther AvgPCVis CTPCP

0.27 74.22 0.10 0.07 9.26 2.89 6.53 0.37

123

Descriptive Data for the Variables in the Three-Level Model - DM Screens 2004 through 2007 - with Repeated Observations within Patients nested within PCPs

Dementia SexConc Termed TotChrDz AvgAge FTR HCostPts IPA PCPSex PCPSexNw HPDCMCD PredCMCD

5661 5661 5661 5661 179 179 179 179 179 179 179 179

0.09 0.60 0.06 15.18


Level three

0.29 0.49 0.24 6.55 2.84 0.47 0.22 0.46 0.49 0.31 0.49 1.62

0.00 0.00 0.00 2.00 63.00 0.00 0.00 0.00 0.00 0.00 0.00 -8.86

1.00 1.00 1.00 46.00 87.00 1.00 1.00 1.00 1.00 1.00 1.00 3.26

73.86 0.34 0.53 0.29 0.60 0.11 0.38 -0.15

In modeling the log-odds of the DM screening measure as a function of time at the midpoint and random effects of time at level-two and level-three, a non-significant variance component for time was identified at the patient level. A significant variance component was found at the PCP level (level-three) therefore a random effect for time was included only at level three. Time was modeled at level-two as fixed. Using HLM software notation this produced the unconditional model of: Level-1 Model Prob DM Screens (Y=1|B) = P log[P/(1-P)] = P0 + P1*(TIMEM) Level-2 Model P0 = B00 + R0 P1 = B10 Level-3 Model B00 = G000 + U00 B10 = G100 + U10 Level-1 variance = 1/[P(1-P)]

124

Laplace model estimates are reported as this estimation method produces the least biased variance components and explaining variation at the PCP level was a primary research interest. The Intraclass Correlation Coefficient (ICC) for this model was 7.7% using the Snijders formula that includes the fixed variance estimate at level-one for the logistic distribution of 3.29.

The Laplace estimates for the unconditional model results are reported in Table 37. Tau Beta had a moderate reliability estimate of 0.631 indicating that level-three units were reliably discriminated and provided reasonable estimates of the true sample mean [90, 93]. The variance components for the PCP level and time were significant. The fixed effect estimate for the intercept modeling the probability of obtaining an annual A1c and LDL screen was positive and significant. The coefficient for the fixed effect of time was also positive. When variation between physicians and time was taken into account without predictors the overall DM screen measure rose to 80.98 percent as compared to only 72 percent in the overall baseline descriptive result for this measure (e1.449/(1+ e1.449).
Table 37. Diabetes Screen HGLM Unconditional Model

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screens 2004 through 2007
(Unconditional Model) Fixed Effect Average DMScreens midpoint (G000) Change in DMScreens over duration (G100) Random Effect Level 1 and 2 Variation-Pt (R0) Level 3 Variance Comp. Coefficient Se t Ratio

1.449 0.393 df 5481 X2

0.069 0.027

20.970 14.620 p Value 0.000

1.513

9999.785

125

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screens 2004 through 2007
(Unconditional Model) Variation-PCP (U00) 0.399 177 732.892 Variation-Time(U10) 0.023 177 242.646 Level 1 % Variance Coefficient Between PCPs Prob DMScreens (P0) 7.68 Change in gap time (P1) .47 Variance-Covariance Components and Correlations Among the Level-2 And Level-3 Random Effects INTRCPT1 TIMEM INTRCPT2,B00 INTRCPT2,B10 Tau(Beta) 0.55556 -0.00864

0.000 0.001

-0.00864
Tau(Beta)

0.04448 -0.055 1.000


Deviance Statistic # Estimated Parameters

INTRCPT1/INTRCPT2,B00 TIMEM/INTRCPT2,B10

1.000 -0.055
Statistic for covariance components model Random level-2 coefficient

56904.947
Reliability Estimate*

Intrcpt1/intrcpt2, B00 0.631 TimeM/intrcpt2, B10 0.358 * - reliability estimates reported are based on only 173 of the 179 PCPs that had sufficient data for computation. Fixed effects and variance components are based on all the data.

Using the Raudenbush step-up approach (stepwise modeling) conditional covariates were added at each level and key contextual variables were entered last at the highest level of analysis. The time-varying covariates of total PCP visits and total endocrinologist visits yearly were added at level-one with fixed effects at level-two and three (Table 38). Hierarchical generalized linear modeling techniques limit the number of random components that can be effectively modeled. Both covariates were examined for a significant random component; total PCP visits had a very small random component that was marginally significant and accompanied by low reliability estimates and therefore not included in the model. The total number of yearly visits to specialists and a key 126

covariate of continuity with the PCP each year (dichotomous variable of one if the entire year is with the same PCP and zero otherwise) were not significant. The total number of visits to health education yearly was significant however to limit time-varying covariates to fewer than the total number of years available to model [174] the variable was not modeled at this level as it had the smallest z-statistic in the exploratory SAS proc genmod procedure. This model would not converge with Laplace estimates therefore unit-specific coefficient estimates and variance component estimates from Full Penalized-Quasi Likelihood estimation are reported. The ICC actually increased slightly to 8.1% not an unexpected finding with the addition of time-varying covariates.
Table 38. Diabetes Screen HGLM Level-One

Table Summarizing Full PQL Parameter Estimates for Three-Level Model DM Screens 2004 through 2007
(Conditional Model with Level-One Time-Varying Covariates) Fixed Effect For INTRCPT1, P0 For INTRCPT2, B00 INTRCPT3, G000 For TPCVIS slope, P1 For INTRCPT2, B10 INTRCPT3, G100
For TENDOVIS slope, P2

Coefficient

Se

t Ratio

1.128 0.066 0.400 0.302

0.060 0.008 0.029 0.022

18.888 7.917 11.554 13.627

For INTRCPT2, B20 INTRCPT3, G200 For TIMEM slope, P3 For INTRCPT2, B30 INTRCPT3, G300 Random Effect Level 1 and 2 Variation-Pt (R0) Level 3 Variation-PCP (U00) Variation-Time(U30) Level 1 Coefficient Prob DMScreens (P0) Change in gap time (P1) Variance Comp.

df 5474 174 174

X2 9654.340 762.703 242.573

p Value 0.000 0.000 0.001

1.448 0.415 0.022


% Variance Between PCPs

8.10 0.46 127

Table Summarizing Full PQL Parameter Estimates for Three-Level Model DM Screens 2004 through 2007
(Conditional Model with Level-One Time-Varying Covariates) Variance-Covariance Components and Correlations Among the Level-2 And Level-3 Random Effects INTRCPT1 TIMEM INTRCPT2,B00 INTRCPT2,B10 Tau(Beta) 0.41501 0.00772

0.00772
Tau(Beta)

0.02193

INTRCPT1/INTRCPT2,B00 TIMEM/INTRCPT2,B10

1.000 0.081 0.081 1.000


Deviance Statistic Statistic for covariance components model Random level-2 coefficient Intrcpt1/intrcpt2, B00 TimeM/intrcpt2, B10 # Estimated Parameters

unavailable
Reliability Estimate*

0.668 0.260 * - reliability estimates reported are based on only 175 of the 179 PCPs that had sufficient data for computation. Fixed effects and variance components are based on all the data. At level-two the patient level covariates that were significant in the bivariate results were entered and those significant at the 0.1 level were retained. A key covariate at the patient level, continuity with the PCP over the patients eligible time with the organization, (CtPCP) was not significant in the bivariate results and was not significant when entered into this conditional model. Continuous variables were grand-mean centered for interpretation of the intercept and to facilitate model convergence [90, 91, 93, 174, 175]. When variables are grand-mean centered the interpretation of the intercept represents the overall group mean for an individual patient with a grand average on every predictor. Variables included the following: age at the start of the study (Ages); termed the organization in 2008 (termed); average number of medication classes prescribed annually (avgmeds); patient gender is equal to PCP gender (sexconc); average annual visits to

128

primary care (avgPCVis); any visit to an endocrinologist (AnyEndo); average other services annually such as PT/OT, pain management, wound care, etc. (AvgOther); any visits to health education (AnyHE), outpatient or inpatient claims indicating diagnosis of anxiety, depression or drug abuse (ADDA); outpatient or inpatient claims indicating diagnosis of dementia (Dementia); and two modifying variables of time, Average Medication classes*TimeM and Termed*TimeM.

With 21 parameters estimated at level-one and level-two the model converged easily with Laplace estimation and the deviance statistic fell to 56038.826. All conditional covariates significant at the intercept were examined as modifiers of time however only the significant covariates of termed and average medication classifications were retained. All poverty variables were examined including Medicaid benefits and none were found to be even close to significance in this model. With the increase in level-two patient covariates the ICC at the PCP level increased further to 9.9 percent indicating differences in the clustering of patient characteristics within assigned PCPs. The probability of adherence to the diabetes screening measure (annual evidence of an A1c and LDL screen) was statistically reduced for the following patient covariates conditional on the other model covariates and the random overall patient intercept: age above the population mean; termed in 2008; average annual PC visits greater than the population mean; average annual other services greater the population mean; total chronic diseases greater than the population mean; a diagnosis of anxiety, depression and/or drug abuse; diagnosis of dementia. Additionally as interaction terms the covariates of termed and average medication classification above the population mean

129

negatively modified the positive impact of time duration on the probability of obtaining the diabetes annual A1c and LDL screens. The probability of obtaining these screens was statistically increased for the following patient covariates conditional on the model parameters: average medication classifications above the population mean; sex concordance; any visits to an endocrinologist; any visits to health education; total yearly PCP visits; total yearly endocrinologist visits; and time at the midpoint. Model estimates conditional on level-one and level-two covariates are reported in Table 39.
Table 39. Diabetes Screen HGLM Level-One and Level-Two

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screens 2004 through 2007
(Conditional Model with Level-One and Level-Two Covariates) Fixed Effect For INTRCPT1, P0 For INTRCPT2, B00 INTRCPT3, G000 For AGES, B01 INTRCPT3, G010 For TERMED, B02 INTRCPT3, G020 For AVGMEDS, B03 INTRCPT3, G030 For SEXCONC, B04 INTRCPT3, G040 For AVGPCVIS, B05 INTRCPT3, G050 For ANYENDO, B06 INTRCPT3, G060 For AVGOTHER, B07 INTRCPT3, G070 For ANYHE, B08 INTRCPT3, G080 For TOTCHRDZ, B09 INTRCPT3, G090 For ADDA, B010 INTRCPT3, G0100 For DEMENTIA, B011 INTRCPT3, G0110 For TPCVIS slope, P1 For INTRCPT2, B10 INTRCPT3, G100 Coefficient Se t Ratio

1.485 -0.046 -0.282 0.116 0.139 -0.041 0.447 -0.042 0.963 -0.025 -0.476 -0.363 0.098

0.086 0.005 0.143 0.010 0.057 0.016 0.153 0.009 0.168 0.006 0.081 0.109 0.008

17.181 -8.047 -1.968 11.533 2.462 -2.498 2.915 -4.798 5.733 -4.003 -5.860 -3.344 11.652

130

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screens 2004 through 2007
(Conditional Model with Level-One and Level-Two Covariates)
For TENDOVIS slope, P2

For INTRCPT2, B20 INTRCPT3, G200 For TIMEM slope, P3 For INTRCPT2, B30 INTRCPT3, G300 For TERMED, B31 INTRCPT3, G310 For AVGMEDS, B32 INTRCPT3, G320

0.357 0.387 -0.216 -0.022

0.038 0.029 0.086 0.005

9.311 13.147 -2.524 -4.294

Random Effect Variance Comp. df X2 p Value Level 1 and 2 Variation-Pt (R0) 1.331 5465 9075.822 0.000 Level 3 Variation-PCP (U00) 0.407 175 764.406 0.000 Variation-Time(U10) 0.022 175 237.145 0.001 Level 1 % Variance Coefficient Between PCPs Prob DMScreens (P0) 8.10 Change in gap time (P1) 0.47 Variance-Covariance Components and Correlations Among the Level-2 And Level-3 Random Effects INTRCPT1 TIMEM INTRCPT2,B00 INTRCPT2,B30 Tau(Beta) 0.61147 0.00369

0.00369
Tau(Beta)

0.04218 0.023 1.000


Deviance Statistic # Estimated Parameters

INTRCPT1/INTRCPT2,B00 TIMEM/INTRCPT2,B30

1.00 0.023
Statistic for covariance components model Random level-2 coefficient

56038.826
Reliability Estimate*

21

Intrcpt1/intrcpt2, B00 0.657 TimeM/intrcpt2, B30 0.341 * - reliability estimates reported are based on only 173 of the 179 PCPs that had sufficient data for computation. Fixed effects and variance components are based on all the data.

At level three many of the variables significant in the bivariate analysis were not significant geriatrician, board-certified internal medicine, specialist other than internal

131

medicine or family practice, years in practice, average annual referrals and average patient caseload. Predicted score for the PCP domain of access and the dichotomous variable for above average predicted score for PCP access (HPredACS) were not significant. As co-linearity was found between predicted scores for the PCP domains of communication and coordination two combined variables were examined HPDCMCD a dichotomous variable indicating the PCP was above average on both communication and coordination scores and PredCMCD a continuous variable as the sum of the predicted scores for communication and coordination. Both variables were significant in the bivariate analysis however model convergence with Laplace estimates resulted only with the inclusion of the dichotomous variable, HPDCMCD. The other final covariates at level-three included the following: Foreign-trained physician (FTr); a non-employee independent physician practice association PCP (IPA); male PCP (PCPSex); ratio of high cost patients (patients with HF, COPD, emphysema, and/or stage III or stage IV renal disease) to linked study patients (HCostPts); and male physician in practice less than ten years (PCPSexNw). Several other contextual PCP variables were examined such as the ratio of mental health patients to total linked patients, average patient age, average patient disease burden, average number of chronic conditions among linked patients; none were significant in the model. The three-level equation for the final model was the following.
Level-1 Model Prob DMScreens (Y=1|B) = P log[P/(1-P)] = P0 + P1*(TPCVIS) + P2*(TENDOVIS) + P3*(TIMEM) Level-2 Model P0 = B00 + B01*(AGES) + B02*(TERMED) + B03*(AVGMEDS) + B04*(SEXCONC) + B05*(AVGPCVIS) + B06*(ANYENDO) + B07*(AVGOTHER) + B08*(ANYHE) + B09*(TOTCHRDZ) + B010*(ADDA) + B011*(DEMENTIA) + R0 P1 = B10 P2 = B20 P3 = B30 + B31*(TERMED) + B32*(AVGMEDS)

132

Level-3 Model B00 = G000 + G001(FTR) + G002(PCPSEX) + G003(IPA) + G004(HPDCMCD) G005(HCOSTPTS) + G006(PCPSEXNW) + U00 B01 = G010 B02 = G020 B03 = G030 B04 = G040 B05 = G050 B06 = G060 B07 = G070 B08 = G080 B09 = G090 B010 = G0100 B011 = G0110 B10 = G100 + G101(IPA) B20 = G200 B30 = G300 + G301(FTR) + G302(HCOSTPTS) + U30 B31 = G310 B32 = G320

The model is also represented by the following equation for the combined model: Nijk = Y000 + Y001*FTRk + Y002*PCPSexk + Y003*IPAk + Y004*HPDCMCDk + Y005*(HCostPtsk HcostPts.) + Y006*PCPSexNwk + Y010*(Agesjk - Ages..) + Y020*termedjk + Y030*(AvgMedsjk AvgMeds..) + Y040*SexConcjk + Y050*(AvgPCVisjk AvgPCVis..) + Y060*AnyEndojk + Y070*(AvgOtherjk AvgOther..) + Y080*AnyHEjk + Y090*(TotChrDzjk TotChrDz..) + Y0100*ADDAjk + Y0110*Dementiajk + Y100*(TPCVisijk TPCVis) + Y200*(TEndoVisijk TendoVis) + Y300*TimeMijk + Y301*FTRk*TimeMijk + Y302*(HCostPtsk HcostPts.)*TimeMijk + Y101*(TPCVisijk TPCVis)*IPAk + Y310*Termedjk*TimeMijk + Y320*(AvgMedsjk AvgMeds..)*Timeijk + R0jk + U00k + U30k*TimeMijk The final fully conditional model results are reported in Table 40. A key PCP predictor, predicted communication and coordination scores that are above average was highly significant with a t-ratio of 3.897 and a p-value beyond 0 .001. Foreign trained physicians and PCPs with higher than average ratio of high cost patients to linked study patients had a marginally higher probability of patients with yearly A1c and LDL screens (p-values 0.055 and 0.068 respectively). PCPs affiliated with IPAs and male PCPs have a lower probability of patients with yearly A1c and LDL screens (p-values 0.018 and 0.017 respectively). All of the patient level covariates that were significant in the conditional one and two-level model remained significant with the exception of termed;

133

termed fell to marginally significant in the same negative direction with a p-value of 0.054. The interaction covariate with termed modifying time remained negative and significant at p-value of 0.013. While IPA overall predicted a lower probability of having patients with yearly A1c and LDL screens, IPA as a modifier to the time-varying covariate of total yearly PCP visits was positive and significant at a p-value of 0.037. Foreign-trained PCPs had a positive influence beyond time duration on the probability of patients with yearly A1c and LDL screens (p-value of 0.021) and a higher than average high cost patient ratio had a marginally significant negative influence on time duration (p-value of 0.063).

Importantly, receipt of health education services a key component in the Chronic Care Model was a significant predictor of the probability for obtaining yearly A1c and LDL screens. While health education resources were not modeled at the site/clinic level this finding supports the relationship between self-management support and quality outcomes. The deviance statistic for this final model was 55988.964 estimating thirty parameters; the deviance statistic for unconditional model was 56904.94703. The estimate of the level-three variance component was 0.303 and the estimate of the time variance component was 0.017; estimates of the initial variance components from the unconditional model were 0.399 and 0.023 respectively. Model predictors reduced levelthree variation however the precise amount of reduction in level-three variation was not calculated because time-varying predictors were in the model [174]. Although tempting to report a significant reduction in the level-three variance component the inclusion of time-varying predictors precludes such a calculation. Time-varying predictors reduce

134

level-two as well as level-three variance and it is impossible to distinguish the appropriate attribution to each level [174]. In this model the estimate for the level-one and level-two variance component of the unconditional model was 1.513 and 1.339 in the final model. While a large percentage of the explained variation at the patient level was the result of the fixed effects level-two parameters, the time-varying covariates at levelone also impacted level-two explained variance.
Table 40. Diabetes Screen HGLM Final Model

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screens 2004 through 2007
(Final Conditional Model) Fixed Effect For INTRCPT1, P0 For INTRCPT2, B00 INTRCPT3, G000 FTR, G001 PCPSEX, G002 IPA, G003 HPDCMCD, G004 HCOSTPTS, G005 For AGES, B01 INTRCPT3, G010 For TERMED, B02 INTRCPT3, G020 For AVGMEDS, B03 INTRCPT3, G030 For SEXCONC, B04 INTRCPT3, G040 For AVGPCVIS, B05 INTRCPT3, G050 For ANYENDO, B06 INTRCPT3, G060 For AVGOTHER, B07 INTRCPT3, G070 For ANYHE, B08 INTRCPT3, G080 For TOTCHRDZ, B09 INTRCPT3, G090 For ADDA, B010 INTRCPT3, G0100 For DEMENTIA, B011 INTRCPT3, G0110 Coefficient Se t Ratio

1.449 0.297 -0.379 -0.405 0.546 0.740 -0.046 -0.283 0.118 0.126 -0.042 0.438 -0.042 0.943 -0.026 -0.489 -0.350

0.154 0.153 0.156 0.170 0.140 0.404 0.006 0.147 0.010 0.059 0.016 0.161 0.009 0.172 0.006 0.084 0.112

9.401 1.931 -2.423 -2.384 3.897 1.833 -7.730 -1.922 11.481 2.130 -2.534 2.726 -4.706 5.488 -4.121 -5.803 -3.121

135

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screens 2004 through 2007
(Final Conditional Model) For TPCVIS slope, P1 For INTRCPT2, B10 INTRCPT3, G100 IPA, G101
For TENDOVIS slope, P2

0.094 0.032 0.355 0.357 0.148 -0.378 -0.219 -0.022

0.008 0.015 0.039 0.036 0.063 0.202 0.087 0.005

11.071 2.079 9.067 9.990 2.332 -1.869 -2.499 -4.188

For INTRCPT2, B20 INTRCPT3, G200 For TIMEM slope, P3 For INTRCPT2, B30 INTRCPT3, G300 FTR, G301 HCOSTPTS, G302 For TERMED, B31 INTRCPT3, G310 For AVGMEDS, B32 INTRCPT3, G320

Random Effect Variance Comp. df X2 p Value Level 1 and 2 Variation-Pt (R0) 1.338 5465 9077.165 0.000 Level 3 Variation-PCP (U00) 0.303 168 635.319 0.000 Variation-Time(U30) 0.017 172 217.969 0.010 Level 1 % Variance Coefficient Between PCPs Prob DMScreens (P0) 6.20 Change in gap time (P1) 0.37 Variance-Covariance Components and Correlations Among the Level-2 And Level-3 Random Effects INTRCPT1 TIMEM INTRCPT2,B00 INTRCPT2,B30 Tau(Beta) 0.43458 0.01963

0.01963
Tau(Beta)

0.03346 0.163 1.000


Deviance Statistic # Estimated Parameters

INTRCPT1/INTRCPT2,B00 TIMEM/INTRCPT2,B30

1.000 0.163
Statistic for covariance components model Random level-2 coefficient Intrcpt1/intrcpt2, B00 TimeM/intrcpt2, B30

55988.964
Reliability Estimate*

30

0.594 0.299

136

* - reliability estimates reported are based on only 173 of the 179 PCPs that had sufficient data for computation. Fixed effects and variance components are based on all the data.

HGLM Diabetes Screening with Control Measure


In modeling the probability of the DM screening with control measure (annual A1c and LDL screens, A1c and LDL control and the absence of an ACSC hospitalization and ED visits for diabetes, CHF, COPD, asthma and hypertension) as a function of time the best fit for the data incorporated time at the midpoint and time (midpoint) squared as fixed effects at level-one, level-two, and level-three. The use of TimeM and TimeM2 reflected the trend in this dependent variable over time as the overall rate increased between year one and two, decreased slightly in year three and decreased more in year four (Figure 1). The variance component for time at level two is not different than zero and at level-three the variance component is very small and marginally significant but accompanied by very low reliability estimates. Time-varying covariates for visits with primary care, specialists, etc. were not significant therefore the aggregated variables (average annual visits to primary care and specialty care) were used as level-two patient covariates. The ICC for this model was 4.2 percent indicating less than five percent of the total variation in this measure was attributed to the physician level.

The sample size for modeling the log-odds of this diabetes screening and control measure was reduced as lab values were not available for all patients in the study sample. The results for patient and PCP sample size and unit-size clusters for this diabetes composite dependent variable model are reported in Tables 41 and Table 42. A total of 5081

137

patients link to 178 PCPs; the majority of PCPs link to at least 20 patients (54.5%) and 50 percent of patients have information for all four years.

Table 41. PCP Sample with Linked Patients Diabetes A1c and LDL Control & No Acute Utilization HGLM Model
Number patients linked to PCP Number of PCPs % of PCPs % of linked patients with all time points Number of PCPs % of PCPs

10 > 10 - < 20 20 - < 35 35 - < 50 = - > 50

53 28 39 23 35 178

29.8 15.7 21.9 12.9 19.7 100.00

= > 90 = > 75 = > 50 < 50

82 57 26 13 178

46.1 32.0 14.6 7.3 100.00

Table 42. Patient Sample Size Repeated Measure Diabetes A1c and LDL Control & No Acute Utilization HGLM Model
Number of time points observed Number of patients % of patients Cumulative frequency of individuals

1 2 3 4

166 600 1771 2543

03.3 11.8 34.9 50.1

166 766 2537 5080

Level-two and level-three covariates were added as a step-up approach maintaining only those covariates with a p-value of at least 0.10. Descriptive statistics for this model are reported in Table 43. Again the dichotomous variable for the key variable of interest was used HPDCMCD indicating that the PCP predicated scores over time for communication and coordination are above average for PCP domains of communication and coordination.

138

Table 43. Diabetes Screen & Control Multilevel Descriptive Statistics

Descriptive Data for the Variables in the Three-Level Model - DM Screens with A1c and LDL Control and No Acute Utilization 2004 through 2007 - with Repeated Observations within Patients nested within PCPs
Variable TimeM TimeM2 N Mean Level one Std Min Max

16629 16629 5080 5080 5080 5080 5080 5080 5080 5080 5080 5080 5080 5080 5080 5080 178 178 178 178 178 178 178 178

0.14 1.27
Level two

1.12 1.00 0.44 6.68 0.26 0.20 4.16 3.39 4.81 0.28 1.38 32.27 0.43 0.50 0.24 1.43 0.54 0.54 1.51 0.21 0.47 0.49 0.50 0.49

-1.50 0.25 0.00 62.00 0.00 0.00 0.00 1.00 0.00 0.00 4.30 60.00 0.00 0.00 0.00 0.37 0.88 5.80 4.55 0.00 0.00 0.00 0.00 0.00

1.50 2.25 1.00 98.00 1.00 1.00 40.50 50.30 47.00 1.00 18.30 286.00 1.00 1.00 1.00 10.46 5.04 9.70 14.43 1.00 1.00 1.00 1.00 1.00

ADDA Ages AnyHE AnyNeph AvgMeds AvgPCVis AvgSpVis Dementia FirstA1c FirstLDL Pov25 PtSex Termed TotRAF AvgTRAF AvgA1c AvgPtMed HCostPts HMM HPDCMCD HPREDCD PCPSEX

0.26 74.00 0.08 0.04 9.36 6.57 5.51 0.09 7.09 107.78 0.25 0.46 0.06 2.66
Level three

2.57 7.05 9.74 0.53 0.33 0.38 0.48 0.60

Significant predictors of the probability or log-odds of patients in this sample with evidence of an annual A1c and LDL screen, A1c and LDL control and the absence of acute utilization are reported in Table 44. HPDCMCD was almost significant at the 0.05 level (p-value of 0.058). Contextual variables at the PCP level constructed by aggregating linked patient risk factors were largely responsible for the reduction in level139

three variation. As an example, PCPs managing a higher percentage of study patients with HF, COPD, emphysema or stage III or IV renal disease had higher log-odds of having patients with evidence of annual A1c and LDL, A1c and LDL control and the absence of potentially avoidable acute care utilization. Patient use of health education services was again a significant positive predictor for this composite measure (p = 0.006). Patients with higher than average number of annual medication classes prescribed, male patients, and patients with higher than average number of specialist visits had higher log odds for this diabetes composite.

The deviance statistic for this final model was 51049.785 estimating 28 parameters; deviance statistic for unconditional model was 52904.144. The final level-three variance component estimate was 0.11715; the initial variance component estimate from the unconditional model was 0.18151. The reduction in the level-three variance component was significant at 35.5 percent. The final mixed equation for this model is presented below: Nijk = Y000 + Y001*PCPSex + Y002*AvgTRAFk + Y003*HPDCMCDk + Y004*AvgA1ck + Y005*HCostPtsk + Y006*AvgPtMedk + Y007*(HMM) + Y010*(Agesjk - Ages..) + Y020*termedjk + Y030*Pov25 + Y040*PtSexjk + Y050*(ToTRaf-ToTRaf)jk + Y060*(AvgMedsjk AvgMeds..) + Y070*(AvgPCVisjk AvgPCVis..) + + Y080*(AvgSpVisjk AvgSpVis..) + Y090*AnyNephjk + Y100*AnyHEjk + Y0110*ADDAjk + Y0120*Dementiajk + Y130*(FirstA1cjk FirstA1c...) + Y140*(FirstLDLjk FirstLDL...) + Y100*TimeMijk + Y110*(TotRafjk TotRaf..)* TimeMijk + Y200*TimeM2ijk + Y210*(TotRafjk TotRaf..)* TimeM2ijk + R0jk + U00k

140

Table 44. HGLM: DM Screen & Control

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screens with A1c and LDL Control and No ACSC 2004 through 2007
(Final Conditional Model) Fixed Effect For INTRCPT1, P0 For INTRCPT2, B00 INTRCPT3, G000 PCPSEX, G001 AVGTRAF, G002 HPDCMCD, G003 AVG1A1C, G004 HCOSTPTS, G005 AVGPTMED,G006 HMM,G007 For AGES, B01 INTRCPT3, G010 For TERMED, B02 INTRCPT3, G020 For POV25, B03 INTRCPT3, G030 For PTSEX, B04 INTRCPT3, G040 For TOTRAF, B05 INTRCPT3, G050 For AVGMEDS, B06 INTRCPT3, G060 For AVGPCVIS, B07 INTRCPT3, G070 For AVGSPVIS, B08 INTRCPT3, G080 For ANYNEPH, B09 INTRCPT3, G090 For ANYHE, B010 INTRCPT3, G0100 For ADDA, B011 INTRCPT3, G0110 For DEMENTIA, B012 INTRCPT3, G0120 For FIRSTA1C, B013 INTRCPT3, G0130 For FIRSTLDL, B014 INTRCPT3, G0140 For TIMEM slope, P1 For INTRCPT2, B10 INTRCPT3, G100 For TOTRAF, B11 INTRCPT3, G110 Coefficient Se t Ratio

-0.740 -0.191 0.081 0.185 -0.116 0.950 0.031 -0.026 -0.007 -0.167 -0.069 0.174 -0.089 0.022 0.012 0.014 -0.359 0.288 -0.193 -0.173 -0.223 -0.018 -0.188 -0.039

1.049 0.099 0.137 0.097 0.115 0.348 0.032 0.096 0.004 0.117 0.078 0.060 0.029 0.008 0.009 0.007 0.134 0.103 0.063 0.094 0.022 0.001 0.0148 0.014

-0.705 -1.923 0.594 1.908 -1.015 2.731 0.967 -0.276 -1.666 -1.428 -0.878 2.883 -3.055 2.629 1.286 2.035 -2.680 2.787 -3.073 -1.834 -9.979 -20.646 -12.707 -2.702

141

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screens with A1c and LDL Control and No ACSC 2004 through 2007
(Final Conditional Model)

For TIMEM2 slope,


For TOTRAF, B21 INTRCPT3, G210 For INTRCPT2, B30 INTRCPT3, G300 Random Effect Level 1 and 2 Variation-Pt (R0) Level 3 Variation-PCP (U00)

P2

0.091 -0.030

0.019 0.015

4.797 -2.029

Variance Comp.

df 4888 170
Deviance Statistic

X2 6863.310 473.912
# Estimated Parameters

p Value 0.000 0.000

0.603 0.117

Statistic for covariance components model Random level-2 coefficient Intrcpt1/intrcpt2, B00

51049.785
Reliability Estimate*

28

0.511

HGLM Diabetes Screening Composite Measure


The log-odds of the diabetes screening composite (A1c, LDL, and colorectal screens) were modeled as a function of time-varying level-one covariates, time-invariant patient covariates at level-two, key PCP predictors and covariates at level-three, a random patient-specific and PCP-specific intercept, and a random time component at level-three. The Intraclass Correlation Coefficient (ICC) for this model was 7.3%. Time had a significant variance component at level-three and was modeled as a fixed effect transformed to the midpoint at level-one and level-two and as a fixed effect with a random effect at level-three. The Tau (beta) correlation for the unconditional model was

142

-0.482 and the reliability estimates for the intercept and time were moderate at 0.571 and 0.650 respectively.

The patient/PCP sample was reduced for this model to 4286 patients and 171 PCPs. The reduction in this model resulted from the imposed age ceiling for inclusion on an individual measure component within the composite, the colorectal screening measure. As presented previously the age ceiling of 80 years was employed as a result of the lack of medical consensus for colorectal screens in persons above 80 years of age. The results of patient clustering and patient repeated measures are reported in Tables 45 and 46. The majority of PCPs link to at least 20 patients (56.7%) and 70 percent of patients have information for all four years.

Table 45. PCP Sample with Linked Patients Diabetes Composite HGLM Model
Number patients linked to PCP Number of PCPs % of PCPs % of linked patients with all time points Number of PCPs % of PCPs

10 > 10 - < 20 20 - < 35 35 - <50 > 50

52 26 46 27 20 171

30.0 15.2 26.9 15.8 14.0 100.00

= > 90 = > 75 = > 50 < 50

53 63 26 25 171

31.0 36.8 15.2 17.0 100.00

Table 46. Patient Sample Size Repeated Measure Diabetes Composite HGLM Model
Number of time points observed Number of patients % of patients Cumulative frequency of individuals

1 2 3 4

247 648 408 2983

05.8 15.1 09.5 69.6

247 895 1301 4286

Descriptive statistics for this model are presented in Table 47. The descriptive statistics do not differ greatly from other longitudinal models except that a higher percentage of 143

PCPs overall scored above average on communication and coordination domain scores (51 percent). Sex concordance between PCPs and linked patients in this model was 61 percent. Slightly fewer patients have continuity with the same PCP over the entire study period than in previous models (36 percent).
Table 47. Diabetes Screening Composite Multilevel Descriptive Statistics

Descriptive Data for the Variables in the Three-Level Model - DM Screening Composite (A1c, LDL and Colorectal screens 2004 through 2007)
Variable Time TimeM DMC TPCVis TSPVis N Mean Level one Std Min Max

15013 15013 15013 15013 15013 4286 4286 4286 4286 4286 4286 4286 171 171 171 171 171 171 171 171

1.65 0.15 0.29 6.54 5.59


Level two

1.10 1.10 0.45 4.43 5.95 5.06 0.27 4.23 0.48 0.28 0.23 0.49 2.41 0.52 1.63 2.67 0.20 0.47 0.50 0.09

0.00 -1.50 0.00 0.00 0.00 62.00 0.00 1.00 0.00 0.00 0.00 0.00 66.50 0.88 4.55 5.75 0.00 0.00 0.00 0.00

3.00 1.50 1.00 56.00 114.00 80.00 1.00 40.50 1.00 1.00 1.00 1.00 81.33 5.04 15.26 27.00 1.00 1.00 1.00 1.00

Ages AnyHE AvgMeds CtPCP Dementia Termed SexConc AvgAge AvgTRAF AvgPtMed AvgPtCDz Geri FTr HPDCMCD RATMM

7218 0.08 9.30 0.36 0.09 0.05 0.61


Level three

73.81 2.60 9.58 15.17 0.04 0.33 0.51 0.08

The equation for the final fitted conditional HGLM model was: Nijk = Y000 + Y001*Gerik + Y002*FTrk + Y003*(AvgAgek AvgAge..) + Y004*(AvgRskk AvgRsk..) + Y005*HPDCMCDk + Y006*(AvgPtMedk AvgPtMed.) + Y007*(AvgPtCDz k AvgPtCDz.) + Y008*(RatMM k RatMM.) + Y010*(Agesjk Ages..) + Y020*CtPCPjk + Y030*termedjk + Y040*(AvgMedsjk AvgMeds..) + 144

Y050*SexConcjk + Y060*AnyHEjk + Y070*Dementiajk + Y100*(TPCVisijk TPCVis) + Y200*(TSpVisijk TSpVis) + Y300*TimeMijk + Y301*FTRk*TimeMijk + Y302*(HPDCMCDk)*TimeMijk + Y303*(AvgPtMedk AvgPtMed)*TimeMijk + Y101*(TPCVisijk TPCVis)*AvgMedsk + Y310*Termedjk*TimeMijk + R0jk + U00k + U30k*TimeMijk Significant predictors for the fitted hierarchical generalized linear model examining the composite diabetes screening measure are reported in Table 48. The key PCP predictor of HPDCMCD was marginally significant (p = 0.07) in the positive direction at the PCP intercept and was found to be a significant modifier to the effect of time indicating an additive impact on time with the diabetes composite measure. PCPs linked to study patients with a higher than average number of chronic conditions also demonstrated a modifying impact on time. Continuity of care with the same PCP over the entire time period was a significant positive predictor in this model (p < 0.001). The time-varying covariates of total PCP visits, total specialist visits were positive predictors for the logodds of adherence with this measure. At level-two the patient characteristics of sex concordance with their PCP and the average number of medication classifications were also positive predictors. Additional contextual variables at the PCP level found to be significant predictors of adherence to this measure were average number of medication classifications prescribed to linked patients and the ratio of patients receiving Medicaid benefits controlling for the overall average risk burden and age of linked patients.

The deviance statistic for this final model was 42839.573 estimating 27 parameters; deviance statistic for unconditional model was 43097.93. The level-three variance component estimates in the final model were 0.28284 for the level-three intercept (U00) and 0.14242 for time (U10); the initial estimates for the variance components from the

145

unconditional model were 0.37737 and 0.15333 respectively. This HGLM model reduced level-three variation by 25 percent and reduced variation in the random time component by seven percent. All the reduction in variation cannot be attributed to the level-three predictors, however, because time-varying covariates were included in the model.
Table 48. HGLM: Diabetes Screening Composite

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screen Composite Measure 2004 through 2007
(Final Conditional Model) Fixed Effect For INTRCPT1, P0 For INTRCPT2, B00 INTRCPT3, G000 GERI, G001 FTR, G002 AVGAGE, G003 AVGRSK, G004 HPDCMCD, G005 AVGPTMED, G006 AVGPTCDZ, G007 RATMM, G008 For AGES, B01 INTRCPT3, G010 For CTPCP, B02 INTRCPT3, G020 For TERMED, B03 INTRCPT3, G030 For AVGMEDS, B04 INTRCPT3, G040 For SEXCONC, B05 INTRCPT3, G050 For ANYHE, B06 INTRCPT3, G060 For DEMENTIA, B070 INTRCPT3, G070 For TPCVIS slope, P1 For INTRCPT2, B10 INTRCPT3, G100 For AVGMEDS, B11 INTRCPT3, G110 For TSPVIS slope, P2 For INTRCPT2, B20 INTRCPT3, G200 Coefficient Se t Ratio

-2.222 -0.742 0.379 0.001 0.029 0.271 0.161 -.0018 1.619 -0.056 0.291 -0.423 0.038 0.265 0.307 -0.461 0.068 -0.003 0.015

0.164 0.338 0.156 0.032 0.203 0.150 0.048 0.041 0.818 0.008 0.099 0.224 0.010 0.088 0.152 0.210 0.007 0.002 0.007

-13.542 -1.931 2.196 0.224 0.141 1.811 3.333 -0.445 1.979 -6.759 2.937 -1.887 3.633 3.018 2.022 -2.196 9.446 -2.032 1.956

146

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screen Composite Measure 2004 through 2007
(Final Conditional Model) For TIMEM slope, P3 For INTRCPT2, B30 INTRCPT3, G300 FTR, G301 HPDCMCD, G302 AVGPTCDZ, G303

0.261 -0.1625 0.246 0.057

0.097 0.120 0.122 0.026

2.703 -1.360 2.015 2.173

Random Effect Variance Comp. df X2 p Value Level 1 and 2 Variation-Pt (R0) 1.449 4108 7450.129 0.000 Level 3 Variation-PCP (U00) 0.283 162 504.952 0.000 Variation-Time(U30) 0.142 172 509.965 0.010 Level 1 % Variance Coefficient Between PCPs Prob DMScreens (P0) 5.60 Change in gap time (P1) 2.89 Variance-Covariance Components and Correlations Among the Level-2 And Level-3 Random Effects INTRCPT1 TIMEM INTRCPT2,B00 INTRCPT2,B30 Tau(Beta) 0.43860 -0.18650

-0.18650
Tau(Beta)

0.30564 -0.509 1.000


Deviance Statistic # Estimated Parameters

INTRCPT1/INTRCPT2,B00 TIMEM/INTRCPT2,B30

1.000 -0.509
Statistic for covariance components model Random level-2 coefficient Intrcpt1/intrcpt2, B00 TimeM/intrcpt2, B30

42839.573
Reliability Estimate*

27

0.506 0.633 * - reliability estimates reported are based on only 170 of the 171 PCPs that had sufficient data for computation. Fixed effects and variance components are based on all the data.

147

HGLM Results 30-day versus 90-day readmission findings


An HGLM model fitted to examine the log-odds of 30-day readmission rates as a function of total primary care visits, total use of other services and time at level-one; termed, total risk burden, the average number of medication classifications, the average number of specialty visits, and tobacco dependence at level-two; and PCPsex, IPA, years in practice less than 10, average risk profile of assigned study patients, and predicted PCP access score. PCP communication and coordination scores independently and jointly were examined and were not significant in the model. The ICC for this model was very low at two percent and the model fitted explained all of the variation at level-one and level-two and most of the variation at level-three; the estimated variance component of the fitted model at level-one and level-two (R0) was 0.16687 with a chi-square of 4226.5 and a p > 0.500, and the estimated variance component at level-three was .04993 with a chi-square of 207.5 and p value of 0.04.

Significant Laplace parameter estimates from the fitted model included the following. The PCP predicted access score was significantly and negatively associated with the logodds of a 30-day readmission (t-ratio = -2.81, p = 0.03). IPA affiliated PCPs were significantly more likely to have patients with 30-day readmission rates (t-ratio = 2.49, p = 0.01). All of the patient covariates were highly significant in the model beyond the 0.01 level with the exception of tobacco dependence that was marginally significant (tratio = 1.72, p = 0.085).

148

An HGLM model fitted to examine the log-odds of 90-day readmission rates as a function of total primary care visits and time at level-one; total risk burden, the average number of medication classifications, the total number of chronic conditions, a diagnosis of dementia, and poverty at level-two; and PCPsex, internist, years in practice less than 10, average risk profile of assigned study patients and predicted PCP coordination score. PCP access and communication scores were examined independently, and communication and coordination scores were examined jointly and were not significant in the model. Similarly the ICC for this model was very low and the model fitted explained all of the variation at level-one and level-two and most of the variation at levelthree; the estimated variance component of the fitted model at level-one and level-two (R0) was 0.39044 with a chi-square of 3847.9 and a p > 0.500, and the estimated variance component at level-three was .06335 with a chi-square of 211.0 and p value of 0.03.

Significant Laplace parameter estimates from the fitted model included the following. The PCP coordination score was significantly and negatively associated with the logodds of a 90-day readmission (t-ratio = -2.19, p = 0.03). Male PCPs were significantly more likely to have patients with 90-day readmission rates (t-ratio = 2.15, p = 0.03). All of the patient covariates were highly significant in the model beyond the 0.01 level with the exception of total chronic conditions that was marginally significant (t-ratio = 1.65, p = 0.10).

HGLM Results Group practice Sample


As stated previously the assessment of the two components of the Chronic Care Model (CCM), health education resources and electronic medical records were only assessed on

149

the subset of patients and PCPs assigned to the group practice. This data was unknown for the IPA affiliates. Similar to the full-sample longitudinal analyses, models were conducted examining the log-odds for quality measures as a function of time-varying level-one covariates, time invariant patient covariates at level-two, and key PCP predictors and covariates at level-three. These two components of the CCM were modeled as fixed effect parameters at the PCP level (level-three) in the analyses conducted.

HGLM Results - Diabetes Screening Measure Group Practice Model


The hierarchical generalized linear model examining the log-odds of the diabetes screen measure was fit similarly to the full sample population for the same measure. In examining the impact of time however, a significant variance component was not found at level-two or level-three. The predicted probability increased linearly with time. Therefore, time transformed to the midpoint was modeled as a fixed effect at all levels with a random component at level-three. The estimated Intraclass Correlation Coefficient (ICC) for this model was 6.3% and the reliability estimate for Tau Beta was 0.589.

The patient/PCP sample for this group practice model was 4835 patients and 125 PCPs. The group practice patient population had a higher percentage of patients with all measurement time periods and a higher percentage of PCPs had patients over a longer period of time. The results of patient clustering and patient repeated measures are reported in Tables 49 and 50. Almost three-fourths of PCPs link to at least 20 patients (72.0%) and 75.5 percent of patients have information for all four years.

150

Table 49. Level 3 Group Practice Diabetes Screen HGLM PCP Sample with Linked Patients
Number patients linked to PCP Number of PCPs % of PCPs % of linked patients with all time points Number of PCPs % of PCPs

10 > 10 - < 20 20 - < 35 35 - <50 > 50

17 20 29 22 39 125

13.6 16.0 23.2 17.6 31.2 100.00

= > 90 = > 75 = > 50 < 50

75 36 7 7 125 100.00

Table 50. Level 2 Group Practice Diabetes Screen HGLM - Patient Sample Size
Number of time points observed Number of patients % of patients Cumulative frequency of individuals

1 2 3 4

167 306 713 3649

0.03 0.10 0.21 1.00

167 473 1186 4835

The descriptive statistics for the level-one, level-two, and level-three for the group practice model did not differ greatly from the full sample with the following exceptions. A lower proportion of PCPs scored above average on the key predictor of combined communication and coordination at 33 percent (versus 38 percent in the full sample). The overall baseline mean for the screening measure was slightly higher at 73 percent and slightly higher use of health education resources and endocrinology (eight and 11 percent respectively). The proportion of foreign-trained PCPs was less at 30 percent (versus 34 percent in the full sample); the proportion of male PCPs was less at 58 percent (versus 60 percent in the full sample). The proportion of high cost patients was higher, 58 percent as compared with 53 percent in the full sample.

The CCM component variables assessed in this model included the presence of health education (HE) and/or care management (CM) services and the days of electronic health records examined as number of days of EHR operation over the study period and the z151

transformed number of days. The overall proportion of HE/CM was thirty-five percent at the PCP-level and the mean number of days for EHR use was 369 days and with a range of 88 to 577 days. The descriptive statistics for all examined variables are reported in Table 51.
Table 51. Group Practice Diabetes Screen Multilevel Descriptive Statistics

Descriptive Data - Three-Level Model - DM Screens 2004 through 2007 Group Practice Population
Variable TimeM TPCVis TSPVis TEndoVis PCPCS DMScreen N Mean Level one Std Min Max

17157 17157 17157 17157 17157 17157 4835 4835 4835 4835 4835 4835 4835 4835 4835 4835 4835 4835 125 125 125 125 125 125 125 125 125

0.12 6.50 5.66 0.37 0.93 0.73


Level two

1.10 4.39 5.78 1.19 0.25 0.45 0.44 6.80 0.31 0.27 4.11 3.36 0.47 0.29 .50 0.49 6.51 1.44 2.48 0.48 148.82 0.46 0.18 0.49 0.49 0.48 1.38

-1.50 0.00 0.00 0.00 0.00 0.00 0.00 62.00 0.00 0.00 1.00 1.00 0.00 0.00 0.00 0.00 2.00 0.37 63.00 0.00 88.00 0.00 0.00 0.00 0.00 0.00 -4.60

1.50 63.00 77.00 14.00 1.00 1.00 1.00 98.00 1.00 1.00 40.50 50.30 1.00 1.00 1.00 1.00 46.00 12.10 78.95 1.00 577.00 1.00 1.00 1.00 1.00 1.00 2.82

ADDA Ages AnyEndo AnyHE AvgMeds AvgPCVis CTPCP Dementia PtSex SexConc TotChrDz TotRAF AvgAge CMHE EHRDays FTR HCostPts PCPSex IM HPDCMCD PredCMCD

0.27 74.25 0.11 0.08 9.14 6.47 0.33 0.09 0.46 0.60 15.14 2.70
Level three

73.58 0.35 369.61 0.30 0.59 0.58 0.62 0.35 -0.12 152

Descriptive Data - Three-Level Model - DM Screens 2004 through 2007 Group Practice Population PromMHPts z-EHRDays 125 125 0.33 0.09 0.16 1.01 0.00 -1.87 1.00 1.50

The final fitted HGLM for the group practice model is represented by the following three-level HLM equation. At level-two significant modifiers of time included patient sex, total patient risk burden, average number of prescribed medication classifications and any use of endocrinology. At level-three the only significant modifier of time identified was CM/HE, a key predictor in the model.
Level-1 Model Prob DMScreens (Y=1|B) = P log[P/(1-P)] = P0 + P1*(TPCVIS) + P2*(TENDOVIS) + P3*(TIMEM) Level-2 Model P0 = B00 + B01*(AGES) + B02*(PTSEX) + B03*(TOTRAF) + B04*(AVGMEDS) + B05*(AVGPCVIS) + B06*(ANYENDO) + B07*(ANYHE) + B08*(ADDA) + B09*(DEMENTIA) + R0 P1 = B10 P2 = B20 P3 = B30 + B31*(PTSEX) + B32*(TOTRAF) + B33*(AVGMEDS) +B34*(ANYENDO) Level-3 Model B00 = G000 + G001(FTR) + G002(IM) + G003(PCPSEX) + G004(CMHE) + G005(AVGAGE) + G006(HCOSTPTS) + G007(PROMHPTS) + + G008(REFRAT) + G007(PREDCMCD) U00 B01 = G010 B02 = G020 B03 = G030 B04 = G040 B05 = G050 B06 = G060 B07 = G070 B08 = G080 B09 = G090 B10 = G100 B20 = G200 B30 = G300 + G301(CMHE) B31 = G310 B32 = G320 B33 = G330 B34 = G340
NOTE: all continuous variables were modeled as grand mean centered

153

Significant predictors for this fitted HGLM examining the diabetes screening measure for the group practice sample are reported in Table 52. The key PCP predictor of combined communication and coordination was largely significant (p < .001) indicating a positive impact on the log-odds of the DM screen measure (annual A1c and LDL screens) conditional on other predictors and the random effects. However, it was not found to be a modifier of any other covariates in the model including time. The presence of on-site health education and/or care manager resources was a significant predictor of this DM measure (p=0.048) and it was found to be a significant modifier of the influence of time for this measure (p < .001). Although a significant bivariate relationship was found between electronic health records and this DM measure, this key CCM variable had no impact and was therefore not retained in the final model.

Once again at the patient level receipt of any health education services was a significant predictor (p < .001). Although any use of endocrinology was a significant predictor for the measure, as a modifier of time it had a significant negative influence on the measure. Average medication classifications performed similarly. Total risk burden was a negative predictor as were the diagnoses of dementia, anxiety, depression, and drug abuse.

The deviance statistic for this final model was 48183.800 estimating 29 parameters; the deviance statistic for unconditional model was 49355.675. The reliability estimate for Tau Beta in the final conditional model was 0.537 indicating a moderate reliability to discriminate between-physician variation in this measure. The level-three variance component estimate in the final model was 0.19237 (U00) compared with the initial

154

estimate for this variance component from the unconditional model of 0.31175. Although this reduction in variation is significant (34.9 percent), this reduction in variation cannot all be attributed to the level-three predictors because time-varying covariates were included in the model.

Table 52. HGLM: Group Practice Diabetes Screen

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screen Measure Group Assigned Population 2004 through 2007
(Final Conditional Model) Fixed Effect For INTRCPT1, P0 For INTRCPT2, B00 INTRCPT3, G000 FTR, G001 IM, G002 PCPSEX, G003 CMHE, G004 AVGAGE, G005 HCOSTPTS,G006 PROMMHPTS,G007 REFRAT,G008 PredCMCD, G009 For AGES, B01 INTRCPT3, G010 For PTSEX, B02 INTRCPT3, G020 For TOTRAF, B03 INTRCPT3, G030 For AVGMEDS, B04 INTRCPT3, G040 For AVGPCVIS, B05 INTRCPT3, G050 For ANYENDO, B06 INTRCPT3, G060 For ANYHE, B07 INTRCPT3, G070 For ADDA, B08 INTRCPT3, G080 For DEMENTIA, B09 INTRCPT3, G090 For TPCVIS slope, P1 For INTRCPT2, B10 Coefficient Se t Ratio

1.676 0.581 -0.259 -0.392 0.300 -0.051 1.070 1.103 2.231 0.217 -0.041 0.259 -0.165 0.107 -0.056 0.474 0.906 -0.472 -0.352 0.093

0.162 0.141 0.157 0.139 0.150 0.040 0.446 0.598 1.466 0.052 0.006 0.068 0.031 0.012 0.017 0.184 0.214 0.097 0.134 0.008

10.307 4.136 -1.657 -2.827 1.997 -1.284 2.398 1.843 1.533 4.l67 -6.111 3.782 -5.423 8.650 -3.330 2.577 4.232 -4.879 -2.628 11.155

155

Table Summarizing La Place Parameter Estimates for Three-Level Model DM Screen Measure Group Assigned Population 2004 through 2007
(Final Conditional Model) INTRCPT3, G100 For TENDOVIS slope,P2 For INTRCPT2, B20 INTRCPT3, G200 For TIMEM slope, P3 For INTRCPT2, B30 INTRCPT3, G300 CMHE,G301 For PTSEX, B31 INTRCPT3, G310 For TOTRAF, B32 INTRCPT3, G320 For AVGMEDS, B33 INTRCPT3, G330 For ANYENDO, B34 INTRCPT3, G340 Random Effect Level 1 and 2 Variation-Pt (R0) Level 3 Variation-PCP (U00) Variance Comp.

0.333 0.305 0.129 0.085 -0.022 -0.019 -0.242 df 4701 115


Deviance Statistic

0.044 0.032 0.334 0.045 0.013 0.006 0.090 X2 7873.765 380.452


# Estimated Parameters

7.562 9.468 3.809 1.899 -1.727 -2.931 -2.701 p Value 0.000 0.000

1.164 0.19237

Statistic for covariance components model Random level-2 coefficient Intrcpt1/intrcpt2, B00

48183.80
Reliability Estimate*

29

0.537

HGLM Results - DM Measure with Control Group practice Model


A second fitted HGLM model examined the log-odds of a DM screening measure with A1c control and the absence of acute utilization as a function of time-varying covariates, level-two covariates and key level-three predictors including the two CCM components. Electronic health records and the present of health education and care management services were not significant in this model.

156

The final fitted HGLM model was the following: Nijk = Y000 + Y001*FTRk + Y002*SPk + Y003*PCPSEXk + Y004*PREDACSk + Y005*HCostPts k + Y006*RATPOV25k + Y007*PREDCMCD k + Y010*MMjk + Y020*termedjk + Y030**(AvgMedsjk AvgMeds..) + Y040*SexConcjk + Y050*AnyENDOjk + Y060*( ToTCDzjk-ToTCDz..) + Y070*(FirstA1cjk FirstA1c..) + Y100*(TPCVISijk TPCVIS..) + Y200*TimeMijk + Y210*( FirstA1cijk FirstA1c..)* TimeMijk + Y300*TimeM2ijk + Y310*( FirstA1cijk FirstA1c..)* TimeM2ijk + R0jk + U00k The key PCP variable of PredCMCD was a significant predictor of the log-odds for this measure at the 0.05 level. The probability of this DM screening measure with control and the absence of acute utilization for a patient with average values on covariates was 84.8 percent conditional on all model parameters. The interpretation as reflected in this model is the following. A non-Medicaid patient with the average number of prescribed medication classifications, average number of chronic conditions, average baseline A1c, average number of primary visits, average duration of time with the organization and assigned to a non-foreign trained, female, non-Spanish speaking PCP who has the average ratios of high cost patients and lower-income patients, had an 84.8 percent probability of meeting this measure. The majority of variation in this measure was explained by patient risk with a significant predictor being the baseline A1c result.

Cross-Sectional Analyses
Patient Survey - HGLM Results Diabetes Screen Composite Measure
In attempting to fit a three-level HGLM to examine potential predictors for the probability of patient survey respondents receiving A1c, LDL and colorectal screens in 2007 the level-three variation could not be estimated. The potential patient survey

157

sample, linked PCPs and associated sites of care are reported in Tables 53 and 54. One fourth of PCPs (25.2 percent) have fewer than three patient respondents and 88.3 percent of PCPs have fewer than 10 patient respondents. Almost half of the clinics/sites of care have fewer than three PCPs (45.4 percent).
Table 53. Patient Survey Final Sample Nested within Primary Care Physicians (PCPs)
Patients linked to PCP 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 18 PCPs 1 2 23 19 21 9 8 5 5 1 1 1 2 2 1 1 1 Cumulative frequency of PCPs 1 3 26 45 66 75 83 88 93 94 95 96 98 100 101 102 103 Cumulative frequency of patients 1 5 77 153 255 309 365 405 450 460 471 483 509 537 552 568 586

Table 54. Patient Survey Final Sample PCPs Nested within Sites/Clinics
PCPs 1 2 3 4 5 6 Sites 5 10 4 7 4 3 Cumulative frequency Sites 1 15 19 26 30 33 Cumulative frequency PCPs 5 25 37 65 85 103

This sample was further reduced with the diabetes screen composite measure as not all survey respondents were eligible for the measure in 2007 and the age ceiling of 80 years was imposed. This reduced the potential sample to 469 patient respondents and 103 PCPs. The model was therefore reduced to a two-level model and the variance

158

component for Tau Beta at level-two was estimated to be 6.0 percent. However, the reliability estimate for Tau Beta was relatively low at 0.174. The estimated variance component using restricted PQL estimation (recognized for underestimating true variance) was 0.20874 (with a standard deviation of 0.457) and p value = .08. Using a step-wise procedure the covariates found significant in the bivariate analyses for this measure were entered into a two-level model with patient respondents at level-one and PCPs at level-two. The final fitted model estimated the probability of the diabetes composite measure as a function of poor health, regular exercise, education less than eight years, lowest quartile of median household income, total patient risk burden, and age in 2007 at level one; and a function of grand-centered MD RC at level-two, with the inclusion of MD RC as a modifier of patient education less than eight years. Level-1 Model Prob DMScreening Composite(Y=1|B) = P log[P/(1-P)] = B0 + B1*(POORHLTH) + B2*(REXERC) + B3*(EDLT8) + B4*(POV25) + B5*(TOTRAF) + B6*(AGEE) Level-2 Model B0 = G00 + G01*(MDRC) + U0 B1 = G10 B2 = G20 B3 = G30 + G31*(MDRC) B4 = G40 B5 = G50 B6 = G60 Level-1 variance = 1/[P(1-P)] In this fitted model MD RC as a modifier of patients low education level was significant (t-ratio=2.028 and p=0.043) although not significant at the PCP intercept level. The RC of the PCP completely moderated the negative impact of a patient reporting education

159

less than eight years. Regular exercise and age were the most significant predictors of this DM composite measure controlling for other predictors in the model including overall patient disease risk. This fitted model explained all of the variation at level-two reducing variation by 57 percent. The final estimated variance component was 0.089 (pvalue = 0.365).

Logistic Regression Results Patient Survey Discussing End of Life and Total PCP Satisfaction
As the variance components for HGLMs fitted to examine the two dependent variables collected from the patient survey discussing end of life preferences with your PCP and total satisfaction with your PCP could not be estimated, a logistic model using the SAS Genmod procedure with class variables for site and PCP were fitted. These models contained the full sample of patient respondents, PCPs and sites described previously. CCM components were not examined in these models as data was unavailable for the entire sample. A step-wise procedure was used by including significant variables from the bivariate analyses and retaining significant variables.

The probability of total satisfaction with PCP was modeled as a function of z-transformed patient component of RC, continuity with PCP over 2007, poor health, and high MD (PCP scored above average on all care coordination domains RC, trust, PCC, and PACIC). The patient component of RC was highly significant at p <0.0001 and high MD was significant at p < 0.01. Patients self-reported poor health was significantly and negatively associated with total satisfaction with PCP (p < 0.01). Continuity with the PCP in 2007 was not significant but was retained in the model as a control variable.

160

The independent variable of z-transformed MD RC as a key predictor of the study was examined in place of high MD in the model and was significant (p < 0.01) however the AICC from that model was 521.8 (BIC = 543.6) compared with the AICC of 184.9 (BIC = 206.7) of the final model. MD RC and MD Trust variables (z-transformed) were also examined jointly in the model in place of high MD resulting in significance for MD trust (p < 0.01) however insignificant for MD RC (p = 0.211). This model also had a higher goodness of model fit statistics (AICC = 529.2 and BIC = 555.3) indicating an inferior fit to the data.

The probability of end of life discussions with PCP was modeled as a function of ztransformed patient component of RC, care manager support, education less than eight years, living alone, self-identified as Asian, total disease burden, and high MD RC and trust combined. The patient component of RC and care manager support were highly significant at p <0.0001 with this measure. High MD RC and trust combined was significantly associated with the probability of patients having end of life discussions with their PCP (p=0.03). Asian patients and patients with higher total risk burdens were also significantly associated with this measure (p = 0.04 and p = .03 respectively). Education less than eight years was negatively associated with this measure (p < 0.0001). Living alone was not significant (p =0.1870) however was left in the model in accordance with goodness of fit statistics.

Logistic Regression Results Linked to CCM Components

161

As a main research question in this study was the influence of CCM on patient outcomes the CCM data was linked to patient survey respondents. As not all sites/clinics had CCM information, the patient sample was reduced to 408 patients, 81 PCPs and 24 clinics. No variance component could be estimated with HGLM therefore a logistic model was fit using the SAS Genmod procedure with class variables for site/clinic and PCP.

The model fit the log-odds of DM screens with a1c control and the absence of ACSC hospitalizations and ED visits for chronic illness as a function of age, total disease burden, education less than eight years, help completing the survey, care manager support, high MD RC and trust combined, PCP internist, and z-transformed CCM part 3a (self-management resources) and z-transformed CCMF overall scores. Both the overall chronic care model score and the score for self-management support were significant (p = 0.07 and 0.03 respectively). High MD RC and trust was not quite significant (p = 0.13) however the coefficient was in the positive direction. All other variables with the exception of age were negatively and significantly associated with this measure.

A second model linking the patient survey data to the CCM data was fit to estimate the log odds of the DM composite quality measure (A1c, LDL and CR screens in 2007) using the Proc Genmod procedure (n = 390 observations, 24 sites and 81 PCPs). The final model fit the log odds of the DM composite measure as a function of age, regular exercise, poor health, education less than eight years, education less than eight years*ztransformed MDRC, z-transformed MDRC, internist, and z-transformed CCMF overall scores. As in the two-level HGLM model examining this same measure, MDRC reverses

162

the negative impact of education less than eight years (p = 0.05). Additionally in this model the overall score for the Chronic Care Model components is positively associated with the outcome (p = 0.06).

Lastly, the PCP survey data could be matched to only a subset of the patient survey respondents (n = 354); only 57 of the 83 PCPs completing the PCP survey could be linked to patient survey respondents. A final logistic model fit the log odds of the DM screen measure (A1c and LDL screens in 2007) as a function of help completing the survey, education less than eight years, care manager support, highMD3 (PCP above average in RC, trust, and PCC indices), z-transformed coordination domain (key variable on the PCP survey, PCP self-reported care coordination behavior composite), internist, and z-transformed CCM part 3a score (self-management resources). Proc Genmod was used with class variables of site and PCP. In this fitted model, z-transformed PCP care coordination domain was significantly associated with receipt of this quality measure (p = 0.05). Additionally, highMD3 was significantly associated with receipt of the measure (p = 0.02) and the CCM component 3a was significantly associated at the 0.10 level (p = 0.09).

163

Chapter 7. Conclusion and Implications


Assessment of Study Hypotheses
In conclusion an examination of the stated hypotheses finds partial support for all hypotheses related to the longitudinal and cross-sectional analyses with the exception of one hypothesis that could not be tested.

Longitudinal analyses hypotheses: H1-1: Clinic Level clinics with on-site HE/CM and more months of EMR improve the diabetes quality composite measure In the fitted HGLM (group practice sample) estimating the probability of annual DM screens the presence of disease management resources (health education and care management) was a significant predictor. In the fitted HGLM estimating the DM composite measure, annual A1c and LDL screens, A1c and LDL control, and the absence of acute utilization it was not a significant predictor. Although not modeled as a clinic variable in the full longitudinal analyses, receipt of health education services at the patient level was a significant predictor in fitted models examining most of the DM quality measures. The variable for days of electronic health records was significant only in bivariate associations with the DM quality measures and not significant in any fitted HGLM.

164

H1-2: Physician Level - Physician care coordination and communication improve the diabetes quality composite measure and medication adherence measure In all fitted HGLMs examining the diabetes quality measures PCP communication and coordination combined were significant predictors. The medication adherence measures could not be fit due to lack of variation found at either the PCP or the clinic/site level for these measures. Moreover, the PCP and clinic key predictors were not found to be significantly associated with medication adherence to oral diabetes medications or to Ace/Arb medications. Although the literature has identified numerous patient factors associated with medication adherence such as low health literacy, poverty, cognitive limitations, etc.[176, 177], physician factors such as a trusting relationship have also been identified [95, 178-180] . One study has found that increasing age and comorbidity actually are associated with an increase in medication adherence [128]. This may explain the high rates of adherence found in this patient population and the resultant lack of PCP variation.

H1-3: Patient Level - Patient continuity with same PCP improves the diabetes quality composite measure and medication adherence In the majority of fitted HGLMs neither variable for PCP continuity, annual PCP continuity or the entire study period continuity, was a significant predictor. In the DM screening composite measure however PCP continuity across the study period was found to be a significant predictor. This screening measure included the colorectal screen. Colorectal screens have the lowest adherence rate nationally among all population groups. Physician recommendation and patient attitude toward the test result in high

165

adherence while fear of finding cancer and the belief that cancer is fatal result in low adherence [181]. Given the low screening rate in general for CR screens, the significance of the continuity of care with the same PCP over the course of the study supports the existing evidence for influencing adherence to this measure. Cross-sectional analyses Hypotheses: H2-1: Clinic Level - Chronic Care Model (overall score) improves the diabetes quality composite measure In the fitted logistic models estimating the log-odds of the DM quality measures the overall score for CCM was significant at just below the 0.10 levels (p = 0.07 and 0.06). The lack of a finding with higher statistical significance was probably the result of the number of clinics with CCM data (n=24). Although data on CCM components related to operations in 2007 was limited to 24 clinics, variation was found across the clinics. Despite system-wide implementation of many of the CCM components, the clinic interviews revealed variation among PCPs in terms of utilizing the decision-support and clinical information systems to their fullest capacity in managing senior patients with chronic conditions. The descriptive statistics from the PCP survey support this variation in use of the physician portals, physician feedback reports, system-generated registry lists, etc. H2-2: Clinic Level Chronic Care Model moderates physician relational coordination Data sample size and the lack of sufficient PCP, and clinic clustering prohibited the testing of this hypothesis. H2-3: Physician Level - PCP relational coordination and trust improve the diabetes quality composite measure, medication adherence measure, end of life discussion, and overall PCP visit satisfaction

166

RC was found to mediate the negative impact of lower education on the diabetes screen composite measure. RC and trust were found to be significant predictors of end of life discussions and overall PCP satisfaction. This finding supports the general hypothesis that RC and trust may reflect measurable constructs for the mechanism within the CCM infrastructure producing productive interactions between informed, activated patients and prepared PCPs. This finding supports the IOM and CMS positions on the need for patient-centered care. While many patient physiological characteristics may not be moderated by RC and trust, patient vulnerabilities undetected in claims data such as low levels of health literacy, low levels of formal education, limited-English language use, and living alone with a limited support system may be influenced by RC and trust and improve outcomes for senior patients with chronic illness. Most importantly they constitute the requisite ingredients in patient-centered medical homes [47, 182].

H2-4: Longitudinal analyses predictors of PCP coordination and communication as proxy variables for the key variables of interest in cross-sectional analyses correlate with cross-sectional analyses predictors of PCP relational coordination and trust. To examine the degree to which the proxy variables (PCP communication and coordination) used in the longitudinal analyses correlated with the key variables of RC and trust employed in the cross-sectional analyses 79 PCPs were identified in both samples. Statistically significant correlations were found between the variables of PCP coordination and communication, and RC and trust. A slightly higher correlation value was found between PCP communication and RC, and PCP communication and trust. The positive correlation lends support to the proxy variables employed in the longitudinal analyses. The PCP factors of communication and coordination combined facilitate the

167

co-production of quality outcomes for a vulnerable population. The combined influence of high levels of PCP communication and coordination as the best predictor of outcomes indicates that a broader construct of PCP characteristics optimally influence patient outcomes. RC and trust are likely separate constructs and both are important to optimizing patient outcomes for senior patients with chronic conditions.

In summary, this study found a strong association between high levels of PCP communication and coordination, and diabetes quality composite measures fitting hierarchical generalized linear models over time in a patient population with significant disease burden. Over fifty percent of the population had a high cost illness in addition to diabetes (HF, COPD, emphysema, and/or stage III or IV renal disease) and over one fourth had a diagnosis of anxiety, depression, or substance use. This finding supports the concept of medical homes, patient-centered care [183-186], and a proposed new model of geriatric care referred to as guided care [187] for patients with complex, chronic conditions.

The guided care model builds upon the successful components of the Chronic Care Model but employs a nurse to support the work of two to five primary care physicians caring for 50 to 60 older, medically-complex patients in improving quality of life and efficiency of resources [182, 187]. Guided care enhances primary care by working with the primary care physician to provide seven chronic care innovations: disease management, self-management, case management, lifestyle modification, transitional care, caregiver education and support, and geriatric evaluation and management [187].

168

Guided care has been implemented within the primary care physician office as an extension of the primary care physician.

The longitudinal analyses in this study also found a significant association between the receipt of health education services and diabetes quality composite measures. The availability and receipt of self-management resources are key components of the CCM and have been demonstrated previously to be effective at improving quality [54, 188192]. The effectiveness of such resources apart from the active involvement of a primary care physician however, has not been successful as evidenced by the two recent CMS demonstration programs [60, 61]. The variable for days of electronic health records was not significant in any of the fitted models however EHR was significantly associated with the receipt of diabetes screening measures in the bivariate relationships. Moreover, the organization implemented EHR across all sites within an 18-month period of time during the later half of the study period. EHR with the embedded physician decision-support and clinical information systems has been previously identified as positively impacting patient outcomes [54, 58, 102, 171]

Although the cross-sectional models examining PCP Relational Coordination (RC) and trust were not robust, an association between RC and the diabetes screen composite was found. In the fitted HGLM PCP RC completely moderated the negative impact on this quality measure from low formal education of the patient (p=0.04). In the fitted logistic model examining overall PCP satisfaction all care coordination indices (jointly) were important to this measure RC, trust, PCC, and PACIC (p < 0.01). Both RC and trust

169

were significantly associated with the probability of patients having end of life discussions with their PCP (p=0.03).

Lastly, the three models that were fit with the CCM component scores found support for the CCM as an infrastructure that influences quality outcome measures. Additionally, the CCM infrastructure appears to support RC and trust by providing the capacity for improved patient-centered care [39, 193]. In the first model examining DM screens with A1c control and the absence of acute utilization, the overall chronic care model score and the score for self-management support were significant (p = 0.07 and 0.03 respectively). In this fitted model the variable indicating high RC and trust did not reach statistical significance however the coefficient was positive. In the second model examining the DM screen composite measure (A1c, LDL, and CR screens), the overall CCM score was significantly associated with receipt of this measure at the 0.06 level. In this fitted model MDRC negated the negative impact of a patients low level of education on the receipt of the DM screen composite measure. In the third model examining DM screens (A1c and LDL) with the inclusion of the care coordination domain from the PCP survey, CCM part 3a (self-management resources) was marginally significant (p = 0.09) and PCP care coordination behavior was significantly associated with DM screens (p = 0.05). The variable highMD3 indicating above average scores on RC, trust, and PCC was significantly associated with this quality measure (p = 0.02). This model illustrates the complexity inherent in previous studies assessing the impact of the CCM on patient outcomes. The CCM infrastructure, physician behavior, relational coordination, and trust each play a role in influencing patient outcomes.

170

Finally, this exploratory study of a large multispecialty medical group practice serving a senior population identified with multiple chronic conditions provides support for the increased capacity for patient-centered care that is realized within the CCM infrastructure. The CCM components reorganize physician practices resulting in increased capacity for care coordination via decision-support, physician feedback, electronic medical records with electronic prescribing, patient registries, selfmanagement resources, etc. however physicians must maximize the CCM tools [194-196] and develop ongoing partnerships with patients in order to achieve the desired outcomes of physician practice reorganization. As CMS embarks on the implementation of patientcentered medical homes for Fee-For-Service beneficiaries the continued focus on redesigning physician practices in accordance with the principles of the CCM and perhaps even the guided care model appears appropriate. Assessing and reinforcing the relational factors of physician-patient partnership such as relational coordination and trust will be critical to its success. As Berenson has stated, a house is not a home and the infrastructure alone does not build relationships [182].

Benefits of HGLM Analyses in this Study


When data is nested within a hierarchical structure as was the case in this study, aggregation bias can occur when a variable takes on different meanings and therefore may have different effects at different organizational levels. As an example the proportion of patients with mental health diagnoses may have an effect on quality outcomes at the physician level or at the clinic level above and beyond the effect of the individual patient with a mental health diagnosis. At the physician and clinic level the

171

ratio of patients with mental health diagnoses may be a proxy for the additional unmeasured challenges in attaining adherence to a particular quality measure and/or processes in place to address such high risk patients. As an example in the fitted model estimating the log odds of the DM screen measure for the group practice population a a diagnosis of anxiety, depression, drug abuse and dementia had statistically significant negative coefficients however the contextual ratio of mental health patients to study patients at the PCP level had a positive coefficient that did not quite reach statistical significance. The use of hierarchical linear models permits the simultaneous examination of individual and unit variables by facilitating the decomposition of any observed relationships between variables such as quality and patients with mental health diagnoses, into separate level-one and level-two components. Additionally, dependence is assessed among individuals belonging to the same unit physician and/or clinic and models incorporate a random effect for each unit. The variability in these random effects is taken into account in estimating standard errors.

In this study the estimated Intra-class Correlation Coefficients (ICCs) ranged from a low of two percent (PCP-level in the patient survey-linked cross-sectional analysis examining the log odds of the DM composite measure) to a high of eight percent (PCP-level in the longitudinal fitted HGLM examining the log odds of annual DM screens). These estimated ICCs are relatively low although sufficient to indicate the non-independence in the data and to use predictors at the PCP level to reduce estimated variation. All fitted HGLMs reduced level-three variation (PCP-level) by 12.4 to 67.5 percent. In the longitudinal models incorporating time-varying predictors such as annual visit rates, total

172

reduction in variance cannot be attributed to the level-three predictors because the incorporation of time-varying predictors reduces variation at the individual level (leveltwo), and the PCP-level (level-three). Lastly key predictors were able to be modeled as interaction terms across levels such as the influence of health education/care management resources at the PCP-level as well as a modifier of time duration in the fitted HGLM estimating the DM screening composite measure.

Study Limitations
There are several limitations to this proposed study that must be acknowledged. This study examined clinic and PCP predictors of quality and adherence among elderly patients with multi-morbidity within a single medical practice organization limiting the ability to generalize the findings. More importantly the study organization was a highperforming organization with many of the components of the CCM in place reducing the variability found upon assessment. The variation identified was largely driven by the degree to which the infrastructure was being maximized at a given clinic versus access to the components such as electronic medical records, decision-support, e-prescribing, physician feedback reports, disease management resources, etc.

As the study design was exploratory without a specific intervention and a comparison population, no causal inferences can be derived. Longitudinal analysis over four years strengthens the validity of the associations found however further experimental research into the key predictors would be warranted before drawing definitive conclusions.

173

Although the key predictors employed in the longitudinal data analyses were imperfect proxy variables for the key variables of interest (RC and trust), statistically significant correlations were found between PCP communication/coordination, and the RC and trust measures.

Additionally, the patient survey response rate was very low at just under 20 percent. This produced a very low potential sample for fitting the cross-sectional analytical models. Numerous explanations for the lower than ideal response rate are hypothesized. The largest barrier to patient response was likely the outreach to a senior population by a third-party researcher via the mail. Although one of the medical directors prepared a cover letter for the survey packet, I was an unknown entity and seniors are probably more reluctant to participate in research. Additionally, seniors may have been hesitant about participation as a result of the request for their permission to link 2007 claims data. Finally, the response rate was likely impacted by the frequent surveying of this same population by the study organization as part of its ongoing quality improvement initiatives. However, given the response rate and possibility for non-comparability few differences were found between respondents and non-respondents as measured by available claims data.

Lastly, the absence of a strong finding of association between physician continuity and patient outcomes must be mentioned. This finding is not contrary to the strong empirical evidence for the benefit of continuity of care with a primary care physician and its relative importance in the current development of the patient-centered medical home

174

[182] . It is likely the result of the intentional study requirement of a high PCP ratio for sample eligibility. A primary care physician ratio of at least 0.60 was required which necessitated the removal of patients frequently changing PCPs from the eligible data sample. As the retained study sample had very high primary care ratios the threshold for the benefit of continuity of care may have been reached. Patients may need to change primary care physicians for many reasons and the change alone may not impact quality of care outcomes whereas the frequent changing of PCPs or the lack of a PCP altogether have been demonstrated to negatively impact the quality of care.

Recommendations for Future Study


This study confirmed the intricacies and complexities of the medical environment in ambulatory care and the challenges inherent in attempting to study them. One recommendation for future study would be qualitative work on the primary care teams including patients. Too often the dynamics of clinic operations are examined as processes done for patients rather than as collaborative activities performed in partnership with patients. Within this particular study organization the PCP is supported by two dedicated nursing assistants who respond to patient calls and run interference for the primary care physicians throughout the day. Future study within this organization would need to include these individuals as part of the effective care team with the PCP and the patient and the care manager as appropriate.

Future study into the best strategies for obtaining patient feedback is critical to continued health care system innovation designed to meet the needs of an aging population. The

175

current IRB regulations protecting HIPAA however create challenges that researchers must successfully overcome. To identify predictors that modify the vulnerabilities patients bring to the medical encounter requires information obtained directly from patients and not available in claims data. As an example in the cross-sectional analyses PCP Relational Coordination was found to completely mediate the negative impact on a diabetes quality composite measure of patient low levels of formal education. Patientcentered care attributes of physicians such as RC, trust, communication, and coordination are likely to mediate patient vulnerabilities that cannot be captured in claims such as lower levels of education, poor health literacy, lack of home support, etc.

Lastly, increased use of hierarchical linear models and generalized hierarchical linear models in health services research is advocated especially within studies examining organizational predictors of outcomes. The naturally nested structure of health care and the urgent need to identify effective processes at various levels of this complex structure should encourage greater use of HLM and HGLM approaches to address such research questions. Health services research currently lags behind the fields of education and social science in this regard.

Health Policy Implications


This study attempted to disentangle key predictors of successful medical encounters in the management of elderly patients with multimorbidity. The study did find significant associations between PCP communication and coordination scores and diabetes quality composite measures over time in a population with significant co-morbidity.

176

Additionally, the study found positive associations for patient receipt of health education services and for the availability of on-site health education and care management services. Given the growing prevalence of diabetes in the elderly and the high cost associated with the frequent complications of renal disease and heart disease [197, 198], this study provides some evidence in support of reimbursement for care coordination infrastructure within primary care. The study also supports the current emphasis on the creation of patient-centered medical homes within an infrastructure of the Chronic Care Models components [182]. Reimbursement to primary care physicians for care coordination may permit the additional time required by physicians to develop and nurture on-going partnerships with patients within the foundation of RC and trust.

A second policy finding highlighted by this study is the need for non-English language support and cultural training for primary care physicians. The association found between Latino patients and PCPs with lower than average scores for Relational Coordination and trust may support other empirical evidence suggesting that Latino patients report lower levels of PCP communication and satisfaction, lower receipts of recommended preventive services [199-201]. The impact of race/ethnicity could not be examined in this study within the longitudinal analyses as this data was not collected by the study organization. It must also be noted that the association between Latino patients and PCPs with lower RC and trust scores is found within a single organization with good performance overall. Recent empirical evidence on hospital quality has found that minority Medicare patients receive care from lower performing providers hospitals and physicians [202-206]. This

177

may further suggest a need for ongoing technical training and supportive resources to lower performing hospitals and providers. Literature suggests that physician-patient relationships are strengthened when patients perceive a shared identity with the provider. Although shared identity might be race concordant care other factors such as patientcentered communication have been found to be important to a patients sense of shared identity with their provider [97].

Finally, the study found that physician relational coordination and trust play an important role in end of life discussions with patients. The high cost in the last year of life among Medicare patients has been well documented by the Dartmouth group and others [207, 208]. The role of PCP partnership including Relational Coordination and trust is critical to the facilitation of meaningful discussions with patients in primary care during which patient preferences and options can be fully explored, and before an emergent medical crisis. Training and support for primary care physicians who treat the vast majority of medically complex seniors may be an important strategy to the increased uptake of the Medicare hospice benefit and to death with dignity for many more patients. The key to the current national health care reform debate on this subject is the need for Medicare patients to have an ongoing partnership with a primary care physician based upon Relational Coordination and trust.

178

References
1. Institute of Medicine: Crossing the Quality Chasm: A New Health System for the 21st Century. Washington D.C.: National Academy of Sciences, 2001. 2. MedPac: Increasing the Value of Medicare. Washington, DC: Medicare Payment Advisory Commission, 2006. 3. Cubanski J, Voris M, Kitchman M, Neuman T, Potetz L: The Medicare ChartBook, The Third Edition. The Henry J. Kaiser Family Foundation, 2005. 4. Kane RL, Priester R, Totten AM: Meeting the Challenge of Chronic Illness. Baltimore: The John Hopkins University Press, 2005. 5. A Portrait of the Chronically Ill in America. Robert Wood Johnson Foundation, 2001. 6. Coleman EA, Williams MV: Executing high-quality care transitions: a call to do it right. J Hosp Med 2007; 2(5): 287-90. 7. Epstein AM: Revisiting readmissions--changing the incentives for shared accountability. N Engl J Med 2009; 360(14): 1457-9. 8. Naylor MD: Transitional care for older adults: a cost-effective model. LDI Issue Brief 2004; 9(6): 1-4. 9. Sherman FT: Rehospitalizations: packaging discharge and transition services to prevent "bounce backs". Geriatrics 2009; 64(5): 8-9. 10. Schoen C, Osborn R, How SK, Doty MM, Peugh J: In chronic condition: experiences of patients with complex health care needs, in eight countries, 2008. Health Aff (Millwood) 2009; 28(1): w1-16. 11. MedPac: Promoting Greater Efficiency in Medicare. Medicare Payment Advisory Commission, 2007. 12. Parchman ML, Noel PH, Lee S: Primary care attributes, health care system hassles, and chronic illness. Med Care 2005; 43(11): 1123-9. 13. Wenger NS, Solomon DH, Roth CP, et al.: The quality of medical care provided to vulnerable community-dwelling older patients. Ann Intern Med 2003; 139(9): 740-7. 14. Starfield B: Primary care in the United States. Int J Health Serv 1986; 16(2): 179-98. 15. Starfield BH, Simborg DW, Horn SD, Yourtee SA: Continuity and coordination in primary care: their achievement and utility. Med Care 1976; 14(7): 625-36. 16. Mohler PJ, Mohler NB: Improving chronic illness care: lessons learned in a private practice. Fam Pract Manag 2005; 12(10): 50-6. 17. Austin B, Wagner E, Hindmarsh M, Davis C: Elements of Effective Chronic Care: A Model for Optimizing Outcomes for the Chronically Ill. Epilepsy Behav 2000; 1(4): S15-S20. 18. Bodenheimer T: Primary care in the United States. Innovations in primary care in the United States. Bmj 2003; 326(7393): 796-9. 19. Bodenheimer T, Wagner EH, Grumbach K: Improving primary care for patients with chronic illness. Jama 2002; 288(14): 1775-9. 20. Solberg LI, Crain AL, Sperl-Hillen JM, Hroscikoski MC, Engebretson KI, O'Connor PJ: Care quality and implementation of the chronic care model: a quantitative study. Ann Fam Med 2006; 4(4): 310-6. 21. Wagner EH, Austin BT, Davis C, Hindmarsh M, Schaefer J, Bonomi A: Improving chronic illness care: translating evidence into action. Health Aff (Millwood) 2001; 20(6): 64-78.

179

22. Gittell JH: Relationships between service providers and their impact on customers. Journal of Management Studies 2002; 41(4): 299-311. 23. Gittell JH: Coordinating mechanisms in care provider groups: relational coordination as a mediator and input uncertainty as a moderator of performance effects. Management Science 2002; 48(11): 1408-1426. 24. Gittell JH: Coordination networks within and across organizations: a multi-level framework. Journal of Management Studies 2004; 41(1): 127-153. 25. Gittell JH: Relational coordination: coordinating work through relationships of shared goals, shared knowledge and mutual respect. In: Kyriakidou, Ozbilgin M, eds. Relational Perspectives in Organizational Studies: A Research Companion Edward Elgar Publishers, 2006. 26. Gittell JH, Fairfield KM, Bierbaum B, et al.: Impact of relational coordination on quality of care, postoperative pain and functioning, and length of stay: a nine-hospital study of surgical patients. Med Care 2000; 38(8): 807-19. 27. Lewin D, Cross R: The strength of weak ties you can trust: the mediating role of trust in knowledge transfer. . Management Science 2004; 50(11): 1477-1490. 28. Mayer RC, Davis JH, Schoorman FD: An integrative model of organizational trust. Academy of Management Review 1995; 20(3): 709-734. 29. Thom DH, Hall MA, Pawlson LG: Measuring patients' trust in physicians when assessing quality of care. Health Aff (Millwood) 2004; 23(4): 124-32. 30. Szasz TS, Hollender MH: A contribution to the philosophy of medicine; the basic models of the doctor-patient relationship. AMA Arch Intern Med 1956; 97(5): 585-92. 31. Cooper-Patrick L, Gallo JJ, Gonzales JJ, et al.: Race, gender, and partnership in the patientphysician relationship. Jama 1999; 282(6): 583-9. 32. van Ryn M: Research on the provider contribution to race/ethnicity disparities in medical care. Med Care 2002; 40(1 Suppl): I140-51. 33. Institute of Medicine: Unequal Treatment. Washington, D.C.: National Academy Press, 2002. (Brian D. Smedley AYS, and Alan R. Nelson, ed. 34. Weinberg DB, Lusenhop RW, Gittell JH, Kautz CM: Coordination between formal providers and informal caregivers. Health Care Manage Rev 2007; 32(2): 140-149. 35. Lings P, Evans P, Seamark D, et al.: The doctor-patient relationship in US primary care. J R Soc Med 2003; 96(4): 180-4. 36. Zickmund SL, Blasiole JA, Brase V, Arnold RM: Congestive heart failure patients report conflict with their physicians. J Card Fail 2006; 12(7): 546-53. 37. Parchman ML, Burge SK: The patient-physician relationship, primary care attributes, and preventive services. Fam Med 2004; 36(1): 22-7. 38. Manderbacka K: Exploring gender and socioeconomic differences in treatment of coronary heart disease. Eur J Public Health 2005; 15(6): 634-9. 39. Piette JD, Schillinger D, Potter MB, Heisler M: Dimensions of patient-provider communication and diabetes self-care in an ethnically diverse population. J Gen Intern Med 2003; 18(8): 624-33. 40. Beck RS, Daughtridge R, Sloane PD: Physician-patient communication in the primary care office: a systematic review. J Am Board Fam Pract 2002; 15(1): 25-38. 41. Wagner EH: Chronic disease management: what will it take to improve care for chronic illness? Eff Clin Pract 1998; 1(1): 2-4.

180

42. Sipkoff M: Health plans are ill-prepared for looming diabetes epidemic. Manag Care 2006; 15(5): 24-5, 28-30, 36. 43. Galbraith J: Organization design: an information processing view: R. D. Irwin, Inc., 1972. (Lorsch JW, ed. Organizational Planning: Cases and Concepts). 44. Tushman M, Nadler D: Information processing as an integrating concept in organizational design. Academy of Management Review 1978; 3(3): 613-24. 45. Argote L: Input uncertainty and organizational coordination in hospital emergency units. Administrative Science Quarterly 1982; 27: 420-434. 46. Weinberg DB, Lusenhop RW, Gittell JH, Kautz CM: Coordination between formal providers and informal caregivers. in-press 2007. 47. Bonds DE, Camacho F, Bell RA, Duren-Winfield VT, Anderson RT, Goff DC: The association of patient trust and self-care among patients with diabetes mellitus. BMC Fam Pract 2004; 5: 26. 48. Pearson SD, Raeke LH: Patients' trust in physicians: many theories, few measures, and little data. J Gen Intern Med 2000; 15(7): 509-13. 49. Wade JC: The patient/physician relationship: one doctor's view. Health Aff (Millwood) 1995; 14(4): 209-12. 50. McLeod A, Shearer M, Taylor M: How to use a case manager--a partnership approach. Aust Fam Physician 2005; 34(1-2): 69-71. 51. Bova C, Fennie KP, Watrous E, Dieckhaus K, Williams AB: The health care relationship (HCR) trust scale: development and psychometric evaluation. Res Nurs Health 2006; 29(5): 477-88. 52. Safran DG, Taira DA, Rogers WH, Kosinski M, Ware JE, Tarlov AR: Linking primary care performance to outcomes of care. J Fam Pract 1998; 47(3): 213-20. 53. Bodenheimer T, Wagner EH, Grumbach K: Improving primary care for patients with chronic illness: the chronic care model, Part 2. Jama 2002; 288(15): 1909-14. 54. Tsai AC, Morton SC, Mangione CM, Keeler EB: A meta-analysis of interventions to improve care for chronic illnesses. Am J Manag Care 2005; 11(8): 478-88. 55. Hroscikoski MC, Solberg LI, Sperl-Hillen JM, Harper PG, McGrail MP, Crabtree BF: Challenges of change: a qualitative study of chronic care model implementation. Ann Fam Med 2006; 4(4): 317-26. 56. Pearson ML, Wu S, Schaefer J, et al.: Assessing the implementation of the chronic care model in quality improvement collaboratives. Health Serv Res 2005; 40(4): 978-96. 57. Asch SM, Baker DW, Keesey JW, et al.: Does the collaborative model improve care for chronic heart failure? Med Care 2005; 43(7): 667-75. 58. Coleman K, Austin BT, Brach C, Wagner EH: Evidence on the Chronic Care Model in the new millennium. Health Aff (Millwood) 2009; 28(1): 75-85. 59. Von Korff M, J G, Schaefer J, S C, Wagner E: Collaborative management of chronic illness. Annals of Internal Medicine 1997; 127(12): 1097-1102. 60. Peikes D, Chen A, Schore J, Brown R: Effects of care coordination on hospitalization, quality of care, and health care expenditures among Medicare beneficiaries: 15 randomized trials. JAMA 2009; 301(6): 603-18. 61. McCall N, Cromwell J, Urato C, Rabiner D: Evaluation of Phase I of the Medicare Health Support Pilot Program Under Traditional Fee-For-Service Medicare: 18-Month Interim Analysis. 2008.

181

62. Schore J, Brown R, Cheh V, Schneider B: Costs and consequences of case management for Medicare beneficiaries. Final Report to Health Care Financing Administration submitted by Mathematica Policy Research, Inc., 1997. 63. Brown R, Peikes D, Chen A, Ng J, Schore J, Soh C: The evaluation of the Medicare Coordinated Care Demonstration: findings for the first two years.: Report to Congress issued by Mathematica Policy Research, Inc., 2007. 64. Flocke SA: Measuring attributes of primary care: development of a new instrument. J Fam Pract 1997; 45(1): 64-74. 65. Safran DG, Karp M, Coltin K, et al.: Measuring patients' experiences with individual primary care physicians. Results of a statewide demonstration project. J Gen Intern Med 2006; 21(1): 13-21. 66. Safran DG, Kosinski M, Tarlov AR, et al.: The Primary Care Assessment Survey: tests of data quality and measurement performance. Med Care 1998; 36(5): 728-39. 67. Glasgow RE, Wagner EH, Schaefer J, Mahoney LD, Reid RJ, Greene SM: Development and validation of the Patient Assessment of Chronic Illness Care (PACIC). Med Care 2005; 43(5): 436-44. 68. Cumming RB, Cameron BA: A comparative analysis of claims-based methods of health risk assessment for commericial populations. A research study sponsored by the Society of Actuaries, 2002. 69. Wickizer TM, Franklin G, Fulton-Kehoe D, Turner JA, Mootz R, Smith-Weller T: Patient satisfaction, treatment experience, and disability outcomes in a population-based cohort of injured workers in Washington State: implications for quality improvement. Health Serv Res 2004; 39(4 Pt 1): 727-48. 70. Williams B, Coyle J, Healy D: The meaning of patient satisfaction: an explanation of high reported levels. Social Science and Medicine 1998; 47(9): 1351-1359. 71. Williams B: Patient satisfaction: a valid concept? Social Science and Medicine 1994; 38(4): 509-516. 72. Dickinson LM, Basu A: Multilevel Modeling and Practice-Based Research. Ann Fam Med 2005; 3(supplement 1): S52-S60. 73. Bardenheier BH, Shefer A, Barker L, Winston CA, Sionean CK: Public health application comparing multilevel analysis with logistic regression: immunization coverage among long-term care facility residents. Ann Epidemiol 2005; 15(10): 749-55. 74. McGuigan WM, Katzev AR, Pratt CC: Multi-level determinants of retention in a homevisiting child abuse prevention program. Child Abuse Negl 2003; 27(4): 363-80. 75. Gibbons RD, Hedeker D: Random effects probit and logistic regression models for threelevel data. Biometrics 1997; 53(4): 1527-37. 76. Hung DY, Glasgow RE, Dickinson LM, et al.: The chronic care model and relationships to patient health status and health-related quality of life. Am J Prev Med 2008; 35(5 Suppl): S398-406. 77. Hung DY, Shelley DR: Multilevel analysis of the chronic care model and 5A services for treating tobacco use in urban primary care clinics. Health Serv Res 2009; 44(1): 103-27. 78. O'Connell AA, McCoach DB: Applications of hierarchical linear models for evaluations of health interventions: demystifying the methods and interpretations of multilevel models. Eval Health Prof 2004; 27(2): 119-51. 79. Tan A, Freeman JL, Freeman DH, Jr.: Evaluating health care performance: strengths and limitations of multilevel analysis. Biom J 2007; 49(5): 707-18. 182

80. Shadish WR, Cook TD, Campbell DT: Experimental and Quasi-Experimental Designs for Generalized Causal Inference. Boston: Houghton Mifflin Company, 2002. 81. Smalarz A: Physician group cultural dimensions and quality performance indicators: not all is equal. Health Care Manage Rev 2006; 31(3): 179-87. 82. Daaleman TP: Reorganizing medicare for older adults with chronic illness. J Am Board Fam Med 2006; 19(3): 303-9. 83. Pham HH, Ginsburg PB: Unhealthy trends: the future of physician services. Health Aff (Millwood) 2007; 26(6): 1586-98. 84. Fortin M, Bravo G, Hudon C, Vanasse A, Lapointe L: Prevalence of multimorbidity among adults seen in family practice. Ann Fam Med 2005; 3(3): 223-8. 85. Fortin M, Lapointe L, Hudon C, Vanasse A: Multimorbidity is common to family practice: is it commonly researched? Can Fam Physician 2005; 51: 244-5. 86. Klabunde CN, Warren JL, Legler JM: Assessing comorbidity using claims data: an overview. Med Care 2002; 40(8 Suppl): IV-26-35. 87. Baldwin LM, Klabunde CN, Green P, Barlow W, Wright G: In search of the perfect comorbidity measure for use with administrative claims data: does it exist? Med Care 2006; 44(8): 745-53. 88. Elixhauser A, Steiner C, Harris DR, Coffey RM: Comorbidity measures for use with administrative data. Med Care 1998; 36(1): 8-27. 89. Longford NT: Random coefficient models: Oxford University Press, USA, 1993. 90. Raudenbush SW, Bryk AS: Hierarchical Linear Models, Applications and Data Analysis Methods, Vol. 1. Thousand Oaks: Sage Publications, 2002. 91. Snijders T, Bosker R: Multilevel Analysis, An Introduction to Basic and Advanced Multilevel Modeling. Thousand Oaks, CA: Sage Publications, 1999. 92. Wong GY, Mason WM: The hierarchical logistic regression model for multilevel analysis. Journal of the American Statistical Association 1985: 513-524. 93. Raudenbush S, Bryk A, Congdon R, duToit M: HLM6, Hierarchical Linear and Nonlinear Modeling. Lincolnwood, IL: Scientific Software International, Inc., 2004. 94. Diaz RE: Comparison of PQL and Laplace 6 estimates of hierarchical linear models when comparing groups of small incident rates in cluster randomised trails. Computational Statistics & Data Analysis 2007; 51: 2871-2888. 95. Piette JD, Heisler M, Krein S, Kerr EA: The role of patient-physician trust in moderating medication nonadherence due to cost pressures. Arch Intern Med 2005; 165(15): 1749-55. 96. Schouten BC, Meeuwesen L: Cultural differences in medical communication: a review of the literature. Patient Educ Couns 2006; 64(1-3): 21-34. 97. Street RL, Jr., O'Malley KJ, Cooper LA, Haidet P: Understanding concordance in patientphysician relationships: personal and ethnic dimensions of shared identity. Ann Fam Med 2008; 6(3): 198-205. 98. Thom DH, Ribisl KM, Stewart AL, Luke DA: Further validation and reliability testing of the Trust in Physician Scale. The Stanford Trust Study Physicians. Med Care 1999; 37(5): 5107. 99. Chumbler NR, Vogel WB, Garel M, Qin H, Kobb R, Ryan P: Health services utilization of a care coordination/home-telehealth program for veterans with diabetes: a matched-cohort study. J Ambul Care Manage 2005; 28(3): 230-40.

183

100. Counsell SR, Callahan CM, Buttar AB, Clark DO, Frank KI: Geriatric Resources for Assessment and Care of Elders (GRACE): a new model of primary care for low-income seniors. J Am Geriatr Soc 2006; 54(7): 1136-41. 101. Himelhoch S, Weller WE, Wu AW, Anderson GF, Cooper LA: Chronic medical illness, depression, and use of acute medical services among Medicare beneficiaries. Med Care 2004; 42(6): 512-21. 102. MacPhail LH, Neuwirth EB, Bellows J: Coordination of diabetes care in four delivery models using an electronic health record. Med Care 2009; 47(9): 993-9. 103. Ouwens M, Wollersheim H, Hermens R, Hulscher M, Grol R: Integrated care programmes for chronically ill patients: a review of systematic reviews. Int J Qual Health Care 2005; 17(2): 141-6. 104. Rothman AA, Wagner EH: Chronic illness management: what is the role of primary care? Ann Intern Med 2003; 138(3): 256-61. 105. Starfield B: Primary care : balancing health needs, services, and technology, Rev. ed. New York: Oxford University Press, 1998. 106. Stock RD, Reece D, Cesario L: Developing a comprehensive interdisciplinary senior healthcare practice. J Am Geriatr Soc 2004; 52(12): 2128-33. 107. Wolff JL, Starfield B, Anderson G: Prevalence, expenditures, and complications of multiple chronic conditions in the elderly. Arch Intern Med 2002; 162(20): 2269-76. 108. Callahan EJ, Bertakis KD, Azari R, Robbins JA, Helms LJ, Chang DW: The influence of patient age on primary care resident physician-patient interaction. J Am Geriatr Soc 2000; 48(1): 30-5. 109. Kong MC, Camacho FT, Feldman SR, Anderson RT, Balkrishnan R: Correlates of patient satisfaction with physician visit: differences between elderly and non-elderly survey respondents. Health Qual Life Outcomes 2007; 5: 62. 110. Wallenhammar LM, Nyfjall M, Lindberg M, Meding B: Health-related quality of life and hand eczema--a comparison of two instruments, including factor analysis. J Invest Dermatol 2004; 122(6): 1381-9. 111. Wang WC, Cunningham EG: Comparison of alternative estimation methods in confirmatory factor analyses of the General Health Questionnaire. Psychol Rep 2005; 97(1): 3-10. 112. Hatcher L: A Step-by-Step Approach to Using SAS for Factor Analysis and Structural Equation Modeling. Cary, NC: SAS Institute, Inc., 1994. 113. Devellis RF: Scale Development Theory and Applications, Second ed, Vol. 26. Thousand Oaks: Sage Publications, 2003. 114. Spector PE: Summated Rating Scale Construction: An Introduction. Newbury Park, CA: Sage, 1992. 115. Larsson SC, Orsini N, Wolk A: Diabetes mellitus and risk of colorectal cancer: a metaanalysis. J Natl Cancer Inst 2005; 97(22): 1679-87. 116. Rousseau MC, Parent ME, Pollak MN, Siemiatycki J: Diabetes mellitus and cancer risk in a population-based case-control study among men from Montreal, Canada. Int J Cancer 2006; 118(8): 2105-9. 117. Schiel R, Beltschikow W, Steiner T, Stein G: Diabetes, insulin, and risk of cancer. Methods Find Exp Clin Pharmacol 2006; 28(3): 169-75.

184

118. Laditka JN, Laditka SB, Mastanduno MP: Hospital utilization for ambulatory care sensitive conditions: health outcome disparities associated with race and ethnicity. Soc Sci Med 2003; 57(8): 1429-41. 119. Rosenblatt RA, Wright GE, Baldwin LM, et al.: The effect of the doctor-patient relationship on emergency department use among the elderly. Am J Public Health 2000; 90(1): 97-102. 120. Benbassat J, Taragin M: Hospital readmissions as a measure of quality of health care: advantages and limitations. Arch Intern Med 2000; 160(8): 1074-81. 121. Maurer PP, Ballmer PE: Hospital readmissions--are they predictable and avoidable? Swiss Med Wkly 2004; 134(41-42): 606-11. 122. Anderson GF, Steinberg EP: Hospital readmissions in the Medicare population. N Engl J Med 1984; 311(21): 1349-53. 123. Anderson GF, Steinberg EP: Predicting hospital readmissions in the Medicare population. Inquiry 1985; 22(3): 251-8. 124. Jencks SF, Williams MV, Coleman EA: Rehospitalizations among patients in the Medicare fee-for-service program. N Engl J Med 2009; 360(14): 1418-28. 125. Jiang HJ, Andrews R, Stryer D, Friedman B: Racial/ethnic disparities in potentially preventable readmissions: the case of diabetes. Am J Public Health 2005; 95(9): 1561-7. 126. Balkrishnan R, Bhosle MJ, Camacho FT, Anderson RT: Predictors of medication adherence and associated health care costs in an older population with overactive bladder syndrome: a longitudinal cohort study. J Urol 2006; 175(3 Pt 1): 1067-71; discussion 1071-2. 127. Balkrishnan R, Rajagopalan R, Camacho FT, Huston SA, Murray FT, Anderson RT: Predictors of medication adherence and associated health care costs in an older population with type 2 diabetes mellitus: a longitudinal cohort study. Clin Ther 2003; 25(11): 2958-71. 128. Rozenfeld Y, Hunt JS, Plauschinat C, Wong KS: Oral antidiabetic medication adherence and glycemic control in managed care. Am J Manag Care 2008; 14(2): 71-5. 129. Hess LM, Raebel MA, Conner DA, Malone DC: Measurement of adherence in pharmacy administrative databases: a proposal for standard definitions and preferred measures. Ann Pharmacother 2006; 40(7-8): 1280-88. 130. Cooke CE, Fatodu H: Physician conformity and patient adherence to ACE inhibitors and ARBs in patients with diabetes, with and without renal disease and hypertension, in a medicaid managed care organization. J Manag Care Pharm 2006; 12(8): 649-55. 131. HEDIS & Quality Measurement. vol 2008. 1100 13th Street, NW, Suite 1000, Washington, DC 20005: National Committee for Quality Assurance 132. Alam T, Weintraub N, Weinreb J: What is the proper use of hemoglobin A1c monitoring in the elderly? J Am Med Dir Assoc 2006; 7(3 Suppl): S60-4, 59. 133. McCall N: Investigation of increasing rates of hospitalization for ambulatory care sensitive conditions among Medicare Fee-for-Service beneficiaries. RTI International, 2004; 141. 134. McCall N, Harlow J, Dayhoff D: Rates of Hospitalization for Ambulatory Care Sensitive Conditions in the Medicare+Choice Population.(Statistical Data Included) Health Care Finan Review 2001; 3. 135. Oster A, Bindman AB: Emergency department visits for ambulatory care sensitive conditions: insights into preventable hospitalizations. Med Care 2003; 41(2): 198-207. 136. AHRQ: Adult diabetes clinical practice guidelines., 2005. 137. McGlynn EA: Six challenges in measuring the quality of health care. Health affairs 1997; 16(3): 7.

185

138. Goldstein H: Multilevel models. Encyclopedia of biostatistics. Wiley, London, UK 1998: 27252731. 139. Goldstein H: Multilevel statistical models. 2003. 140. Leeuw Jde, Meijer E: Handbook of Multilevel Analysis. New York, NY: Springer, 2008. 141. Rice N, Jones A: Multilevel models and health economics. Health Economics 1997; 6(6): 561-575. 142. Mold JW, Fryer GE, Roberts AM: When do older patients change primary care physicians. Journal of the American Board of Family Medicine, vol 17, 2004; 453-460. 143. Nutting PA, Goodwin MA, Flocke SA, Zyzanski SJ, Stange KC: Continuity of primary care: to whom does it matter and when? Ann Fam Med 2003; 1(3): 149-55. 144. Saultz JW: Defining and measuring interpersonal continuity of care. Ann Fam Med 2003; 1(3): 134-43. 145. Stokes T, Tarrant C, Mainous AG, 3rd, Schers H, Freeman G, Baker R: Continuity of care: is the personal doctor still important? A survey of general practitioners and family physicians in England and Wales, the United States, and The Netherlands. Ann Fam Med 2005; 3(4): 353-9. 146. Johansson P, Dahlstrom U, Brostrom A: Consequences and predictors of depression in patients with chronic heart failure: implications for nursing care and future research. Prog Cardiovasc Nurs 2006; 21(4): 202-11. 147. Shoor S, Lorig KR: Self-care and the doctor-patient relationship. Med Care 2002; 40(4 Suppl): II40-44. 148. Zrinyi M, Juhasz M, Balla J, et al.: Dietary self-efficacy: determinant of compliance behaviours and biochemical outcomes in haemodialysis patients. Nephrol Dial Transplant 2003; 18(9): 1869-73. 149. Lorig K, Stewart A, Ritter P, Gonzalez V, Laurent D, Lynch J: Outcome Measures for Health Education and other Health Care Interventions. Thousand Oaks, CA.: Sage Publications, Inc., 1996. 150. Morisky DE, Ang A, Krousel-Wood M, Ward HJ: Predictive validity of a medication adherence measure in an outpatient setting. J Clin Hypertens (Greenwich) 2008; 10(5): 348-54. 151. Morisky DE, Green LW, Levine DM: Concurrent and predictive validity of a self-reported measure of medication adherence. Med Care 1986; 24(1): 67-74. 152. Roth MT, Ivey JL: Self-reported medication use in community-residing older adults: A pilot study. Am J Geriatr Pharmacother 2005; 3(3): 196-204. 153. Shalansky SJ, Levy AR, Ignaszewski AP: Self-reported Morisky score for identifying nonadherence with cardiovascular medications. Ann Pharmacother 2004; 38(9): 1363-8. 154. Wagner EH, Grothaus LC, Sandhu N, et al.: Chronic care clinics for diabetes in primary care: a system-wide randomized trial. Diabetes Care 2001; 24(4): 695-700. 155. Baker DW, Gazmararian JA, Williams MV, et al.: Functional health literacy and the risk of hospital admission among Medicare managed care enrollees. Am J Public Health 2002; 92(8): 1278-83. 156. Gazmararian JA, Williams MV, Peel J, Baker DW: Health literacy and knowledge of chronic disease. Patient Educ Couns 2003; 51(3): 267-75. 157. Schillinger D, Bindman A, Wang F, Stewart A, Piette J: Functional health literacy and the quality of physician-patient communication among diabetes patients. Patient Educ Couns 2004; 52(3): 315-23. 186

158. Schillinger D, Grumbach K, Piette J, et al.: Association of health literacy with diabetes outcomes. Jama 2002; 288(4): 475-82. 159. Benjamins MR: Religion and functional health among the elderly: is there a relationship and is it constant? J Aging Health 2004; 16(3): 355-74. 160. Benjamins MR: Religious influences on trust in physicians and the health care system. Int J Psychiatry Med 2006; 36(1): 69-83. 161. Benjamins MR: Religious influences on preventive health care use in a nationally representative sample of middle-age women. J Behav Med 2006; 29(1): 1-16. 162. Koenig HG, Bearon LB, Dayringer R: Physician perspectives on the role of religion in the physician-older patient relationship. J Fam Pract 1989; 28(4): 441-8. 163. Mull CS, Cox CL, Sullivan JA: Religion's role in the health and well-being of well elders. Public Health Nurs 1987; 4(3): 151-9. 164. Daly JM, Hartz AJ, Xu Y, et al.: An assessment of attitudes, behaviors, and outcomes of patients with type 2 diabetes. J Am Board Fam Med 2009; 22(3): 280-90. 165. Fossum E, Gleim GW, Kjeldsen SE, et al.: The effect of baseline physical activity on cardiovascular outcomes and new-onset diabetes in patients treated for hypertension and left ventricular hypertrophy: the LIFE study. J Intern Med 2007; 262(4): 439-48. 166. Piette JD, Wagner TH, Potter MB, Schillinger D: Health insurance status, cost-related medication underuse, and outcomes among diabetes patients in three systems of care. Med Care 2004; 42(2): 102-9. 167. Rodriguez HP, Rogers WH, Marshall RE, Safran DG: The effects of primary care physician visit continuity on patients' experiences with care. J Gen Intern Med 2007; 22(6): 787-93. 168. Schafer JL, Graham J: Missing Data: Our View of the State of the Art. Psychological Methods 2002; 7(2): 147-177. 169. Dugan E, Trachtenberg F, Hall MA: Development of abbreviated measures to assess patient trust in a physician, a health insurer, and the medical profession. BMC Health Serv Res 2005; 5: 64. 170. Kelly PA, O'Malley KJ, Kallen MA, Ford ME: Integrating validity theory with use of measurement instruments in clinical settings. Health Services Research 2005; 40(5 (part II)): 1605-1619. 171. Dorr DA, Wilcox A, Burns L, Brunker CP, Narus SP, Clayton PD: Implementing a multidisease chronic care model in primary care using people and technology. Dis Manag 2006; 9(1): 1-15. 172. Sperl-Hillen JM, Solberg LI, Hroscikoski MC, Crain AL, Engebretson KI, O'Connor PJ: Do all components of the chronic care model contribute equally to quality improvement? Jt Comm J Qual Saf 2004; 30(6): 303-9. 173. Matza LS, Park J, Coyne KS, Skinner EP, Malley KG, Wolever RQ: Derivation and validation of the ASK-12 adherence barrier survey. Ann Pharmacother 2009; 43(10): 162130. 174. Singer JD, Willet JB: Applied Longitudinal Data Analysis, Modeling Change and Event Occurence. New York, NY: Oxford University Press, 2003. 175. Wu YW, Wooldridge PJ: The impact of centering first-level predictors on individual and contextual effects in multilevel data analysis. Nurs Res 2005; 54(3): 212-6. 176. Bull SA, Hu XH, Hunkeler EM, et al.: Discontinuation of use and switching of antidepressants: influence of patient-physician communication. JAMA 2002; 288(11): 1403-9. 187

177. Clifford S, Barber N, Horne R: Understanding different beliefs held by adherers, unintentional nonadherers, and intentional nonadherers: application of the NecessityConcerns Framework. J Psychosom Res 2008; 64(1): 41-6. 178. Berry LL, Parish JT, Janakiraman R, et al.: Patients' commitment to their primary physician and why it matters. Ann Fam Med 2008; 6(1): 6-13. 179. Donohue JM, Huskamp HA, Wilson IB, Weissman J: Whom do older adults trust most to provide information about prescription drugs? Am J Geriatr Pharmacother 2009; 7(2): 10516. 180. Gajdosik J, Brukkerova D, Kriska M, Svorenova A: A comparison of the opinion of general practitioners and their patients on compliance with pharmacotherapy. Bratisl Lek Listy 2009; 110(6): 350-3. 181. Subramanian S, Klosterman M, Amonkar MM, Hunt TL: Adherence with colorectal cancer screening guidelines: a review. Preventive Medicine 2004; 38(5): 536-550. 182. Berenson R, Hammons T, Gans D, et al.: A house is not a home: keeping patients at the center of practice redesign. Health affairs 2008; 27(5): 1219-1230. 183. Bayliss EA, Edwards AE, Steiner JF, Main DS: Processes of care desired by elderly patients with multimorbidities. Fam Pract 2008; 25(4): 287-93. 184. Berenson RA, Hammons T, Gans DN, et al.: A house is not a home: keeping patients at the center of practice redesign. Health Aff (Millwood) 2008; 27(5): 1219-30. 185. Cooley WC, McAllister JW, Sherrieb K, Clark RE: The Medical Home Index: development and validation of a new practice-level measure of implementation of the Medical Home model. Ambul Pediatr 2003; 3(4): 173-80. 186. Fiscella K, Meldrum S, Franks P, et al.: Patient trust: is it related to patient-centered behavior of primary care physicians? Med Care 2004; 42(11): 1049-55. 187. Boyd CM, Boult C, Shadmi E, et al.: Guided care for multimorbid older adults. Gerontologist 2007; 47(5): 697-704. 188. Baker DW, Asch SM, Keesey JW, et al.: Differences in education, knowledge, selfmanagement activities, and health outcomes for patients with heart failure cared for under the chronic disease model: the improving chronic illness care evaluation. J Card Fail 2005; 11(6): 405-13. 189. Coleman MT, Newton KS: Supporting self-management in patients with chronic illness. Am Fam Physician 2005; 72(8): 1503-10. 190. Fenton JJ, Levine MD, Mahoney LD, Heagerty PJ, Wagner EH: Bringing geriatricians to the front lines: evaluation of a quality improvement intervention in primary care. J Am Board Fam Med 2006; 19(4): 331-9. 191. Schillinger D, Handley M, Wang F, Hammer H: Effects of self-management support on structure, process, and outcomes among vulnerable patients with diabetes: a three-arm practical clinical trial. Diabetes Care 2009; 32(4): 559-66. 192. Williams GC, Patrick H, Niemiec CP, et al.: Reducing the health risks of diabetes: how selfdetermination theory may help improve medication adherence and quality of life. Diabetes Educ 2009; 35(3): 484-92. 193. Heisler M, Vijan S, Anderson RM, Ubel PA, Bernstein SJ, Hofer TP: When do patients and their physicians agree on diabetes treatment goals and strategies, and what difference does it make? J Gen Intern Med 2003; 18(11): 893-902. 194. Nutting PA, Dickinson WP, Dickinson LM, et al.: Use of chronic care model elements is associated with higher-quality care for diabetes. Ann Fam Med 2007; 5(1): 14-20. 188

195. Greene J, Yedidia MJ: Provider behaviors contributing to patient self-management of chronic illness among underserved populations. J Health Care Poor Underserved 2005; 16(4): 808-24. 196. Heisler M, Cole I, Weir D, Kerr EA, Hayward RA: Does physician communication influence older patients' diabetes self-management and glycemic control? Results from the Health and Retirement Study (HRS). J Gerontol A Biol Sci Med Sci 2007; 62(12): 143542. 197. Sloan FA, Bethel MA, Ruiz D, Jr., Shea AM, Feinglos MN: The growing burden of diabetes mellitus in the US elderly population. Arch Intern Med 2008; 168(2): 192-9; discussion 199. 198. Schneider KM, O'Donnell BE, Dean D: Prevalence of multiple chronic conditions in the United States' Medicare population. Health Qual Life Outcomes 2009; 7: 82. 199. Doescher MP, Saver BG, Franks P, Fiscella K: Racial and ethnic disparities in perceptions of physician style and trust. Archives of Family Medicine 2000; 9(10): 1156. 200. Morales LS, Cunningham WE, Brown JA, Liu H, Hays RD: Are Latinos less satisfied with communication by health care providers? Journal of General Internal Medicine 1999; 14(7): 409-417. 201. Saha S, Arbelaez JJ, Cooper LA: Patient-physician relationships and racial disparities in the quality of health care. American Journal of Public Health 2003; 93(10): 1713. 202. Fine JM, Fine MJ, Galusha D, Petrillo M, Meehan TP: Patient and hospital characteristics associated with recommended processes of care for elderly patients hospitalized with pneumonia: results from the medicare quality indicator system pneumonia module. Arch Intern Med 2002; 162(7): 827-33. 203. Gaskin DJ, Spencer CS, Richard P, Anderson GF, Powe NR, Laveist TA: Do hospitals provide lower-quality care to minorities than to whites? Health Aff (Millwood) 2008; 27(2): 518-27. 204. Hausmann LR, Ibrahim SA, Mehrotra A, et al.: Racial and ethnic disparities in pneumonia treatment and mortality. Med Care 2009; 47(9): 1009-17. 205. Jha AK, Orav EJ, Zheng J, Epstein AM: The characteristics and performance of hospitals that care for elderly Hispanic Americans. Health Aff (Millwood) 2008; 27(2): 528-37. 206. Mortensen EM, Cornell J, Whittle J: Racial variations in processes of care for patients with community-acquired pneumonia. BMC Health Serv Res 2004; 4(1): 20. 207. Kronman AC, Ash AS, Freund KM, Hanchate A, Emanuel EJ: Can primary care visits reduce hospital utilization among Medicare beneficiaries at the end of life? J Gen Intern Med 2008; 23(9): 1330-5. 208. Wennberg JE, Fisher ES, Stukel TA, Skinner JS, Sharp SM, Bronner KK: Use of hospitals, physician visits, and hospice care during last six months of life among cohorts loyal to highly respected hospitals in the United States. BMJ 2004; 328(7440): 607.

189

APPENDICES

190

Appendix A. The Chronic Care Model

The Chronic Care Model


Community
Resources & Policies
Self-Mgmt Support

Health System
Organization of Health Care & Safety
Delivery System Design

Decision Support

Clinical Information Systems

Informed, Active Patient

Productive Interactions

Prepared, Proactive Team

Functional and Clinical Outcomes

Relational Coordination
Communication Frequent Timely Accurate Problem solving Relationships Shared goals Shared knowledge Mutual respect

Trust

Patient

Productive Interactions

Primary Doctor

Quality Outcomes and Adherence

191

Appendix B.

Model I Distributions for PCP Key Domains

35 30 25

Mean Median

-0.03505 -0.04749

Std Deviation 0.766459 Skewness -0.14634 Kurtosis 0.725239

Percent
20 15 10 5 0 -3.3 -2.7 -2.1 -1.5 -0.9 -0.3 0.3 0.9 1.5 2.1 2.7 3.3

Predicted z-score Access

35 30 25

Mean Median

-0.08043 -0.064

Std Deviation 0.873396 Skewness -1.58816 Kurtosis 5.679241

Percent 20
15 10 5 0 -5.1 -4.5 -3.9 -3.3 -2.7 -2.1 -1.5 -0.9 -0.3 0.3 0.9 1.5 2.1

Predicted z-score Communication

192

50

Mean Median Std Deviation Skewness Kurtosis

-0.06651 -0.00874 0.818653 -1.65347 7.178108

40

Percent
20

30

10

0 -5.6 -4.8 -4.0 -3.2 -2.4 -1.6 -0.8 0 0.8 1.6 2.4

Predicated z-score Coordination

50

Mean Median

-0.14694 0.00898

40

Std Deviation 1.619322 Skewness -1.70782 Kurtosis 6.782733

30

Percent
20

10

0 -10.5 -9.0 -7.5 -6.0 -4.5 -3.0 -1.5 0 1.5 3.0 4.5

Sum of Predicted z-scores, Communication + Coordination


193

Appendix C. Correlation between different constructions of PCP Domains


PCP Domain Communication Constructed Variables

Variable predCM avgCM zCM

Label Predicted Communication (HLM) Average raw score across all years z-transformed score of annual ztransformed communication scores

N 179 179 179

Mean -0.080 4.407 0

StdDev 0.873 0.216 1.000

Median -0.064 4.145 0.111

Min -4.501 3.150 -5.886

Max 1.521 4.817 1.956

predCM avgCM zCM

Pearson Correlation Coefficients, N=179 Prob > Rho under H0: Rho=0 predCM avgCM 1.000 0.994 < .0001 0.994 1.000 < .0001 0.966 0.974 < .0001 < .0001

zCM 0.966 < .0001 0.974 < .0001 1.000

predCM avgCM zCM

Spearman Correlation Coefficients, N=179 Prob > Rho under H0: Rho=0 predCM avgCM 1.000 0.999 < .0001 0.999 1.000 < .0001 0.969 0.971 < .0001 < .0001

zCM 0.969 < .0001 0.971 < .0001 1.000

194

PCP Domain Coordination Constructed Variables


Variable predCD avgCD zCD Label Predicted Coordination (HLM) Average raw score across all years z-transformed score of annual ztransformed coordination scores N 179 179 179 Mean StdDev Median -0.067 0.819 -0.009 4.204 0.201 4.222 0 1.000 -0.055 Min Max -4.863 1.768 2.783 4.667 -7.443 2.408

predCD avgCD zCD

Pearson Correlation Coefficients, N=179 Prob > Rho under H0: Rho=0 predCD avgCD 1.000 0.991 < .0001 0.991 1.000 < .0001 0.950 0.967 < .0001 < .0001

zCD 0.950 < .0001 0.967 < .0001 1.000

predCD avgCM zCM

Spearman Correlation Coefficients, N=179 Prob > Rho under H0: Rho=0 predCD avgCD 1.000 0.997 < .0001 0.997 1.000 < .0001 0.936 0.939 < .0001 < .0001

zCD 0.936 < .0001 0.939 < .0001 1.000

195

PCP Domain Access Constructed Variables


Variable predACS avgACS zACS Label Predicted Access (HLM) Average raw score across all years z-transformed score of annual ztransformed access scores N 179 179 179 Mean StdDev Median -0.035 0.766 -0.047 3.935 0.222 3.933 0 1.000 -0.211 Min Max -2.700 2.169 2.967 4.597 -4.353 3.011

predACS avgACS zACS

Pearson Correlation Coefficients, N=179 Prob > Rho under H0: Rho=0 predACS avgACS 1.000 0.991 < .0001 0.991 1.000 < .0001 0.969 0.979 < .0001 < .0001

zACS 0.969 < .0001 0.979 < .0001 1.000

predACS avgACS zACS

Spearman Correlation Coefficients, N=179 Prob > Rho under H0: Rho=0 predACS avgACS 1.000 0.997 < .0001 0.997 1.000 < .0001 0.970 0.975 < .0001 < .0001

zACS 0.970 < .0001 0.975 < .0001 1.000

196

Appendix D. Correlation Matrix PCP Domain Scores


Correlation between the individual Key Predictors of PCP Communication, Coordination, and Access
Variable predCD predCM predACS N 179 179 179 Mean -0.067 -0.080 -0.035 StdDev 0.819 0.873 0.767 Median -0.009 -0.064 -0.047 Min -4.864 -4.508 -2.700 Max 1.768 1.521 2.169

predCD predCM predACS

Pearson Correlation Coefficients, N=179 Prob > Rho under H0: Rho=0 predCD predCM 1.000 0.832 < .0001 0.832 1.000 < .0001 0.804 0.550 < .0001 < .0001

predACS 0.804 < .0001 0.717 < .0001 1.000

Spearman Correlation Coefficients, N=179 Prob > Rho under H0: Rho=0 predCD predCM predACS 1.000 0.754 0.761 predCD < .0001 < .0001 0.754 1.000 0.654 predCM < .0001 < .0001 0.761 0.421 1.000 predACS < .0001 < .0001

197

Appendix E. Patient Survey Packet Note: the font-size used for actual documents was 13-point (font-size shrunk for dissertation margin requirements)
[Organization Letterhead] Date Dear Patient: As a patient of XX Medical Group you are invited to take part in a study that is looking at how doctors and senior patients work as a team to improve the health and quality of life of seniors living with chronic illness. This study is being done by a doctoral student at Brandeis University, Marian Ryan who has worked for many years in health care in southern California. More information about the study is contained in the enclosed patient consent form. Ms. Ryan has not been granted your name or any other private health information about you. We have agreed to send her study materials to our patients meeting her study criteria and you were selected. The focus of her study is persons who are 65 years of age and older living with diabetes and at least one additional chronic condition. You have the right to choose not to participate in this study. However, we hope after reading the enclosed materials that you will want to participate. If you do wish to participate in the study please read and sign the enclosed consent form. Complete the patient survey and mail the consent form and the survey in the prepaid envelope provided. Everyone who completes and returns the survey within the first four weeks will be entered into a drawing for a $50.00 grocery gift certificate. A total of 20 winners will be chosen. If you have any questions about this study you can call the researcher, Marian Ryan at 949-290-7697. Materials are being mailed in English and Spanish however if another language is needed please call Marian. You may also call Marian if you wish to participate in the study and do the survey over the telephone. You can leave a message and your telephone number with her and she will call you back. Your responses to this survey will greatly help make this research study a success! This study will inform health policy to improve systems and processes that will help seniors get better health care. It is also hoped that this study might reduce health inequalities. Sincerely, [Signed by the Medical Director]

198

Patient Informed Consent Study Title: Examining the Factors Associated with Patient Outcomes among Seniors Living with Multiple Chronic Diseases What is the purpose of the study? This study is looking at care coordination and the role that your primary care doctor plays in helping you live well with chronic illnesses. Who is doing this study? Hello, my name is Marian Ryan and I am a third-year doctoral student at the Heller School, Brandeis University in Waltham, Massachusetts. You were selected by chance from among patients with chronic illness. XX has agreed to allow me to invite your participation in my study; however, you get to decide whether you want to participate. I am the researcher for this study and I would like to tell you a little about myself. I worked for many years in the healthcare system in southern California before returning to school for my doctoral degree. I have a degree in public health and worked as a respiratory therapist and a health educator. In addition to my enjoyable years of working directly with patients, I have personally faced some of the problems in the health care system today. These experiences fueled my desire to return to school. My father died at a young age of lung cancer and leukemia as they were not diagnosed early enough. Many of my friends with chronic disease have been challenged in getting all the knowledge, skills and resources they need to enjoy a good qualify of life. My passion for changing health policy to improve care for seniors and others such as the poor and those with limited English led to my doctoral fellowship. I am thrilled that XX supports this approved study and allowed me to invite your participation. Many doctors and organizations are doing a good job in patient care. I hope to find out more about factors that help and hinder your care. A faculty member at the college will be supervising my work. Who is funding this study? This study is being partially funded by a grant awarded by the Jewish Healthcare Foundation. I am also funding a large part of this study myself. I believe with your help that changes in our health care delivery system and policy can improve the quality of care for seniors and others living with multiple chronic illnesses. Do I have to participate? Your participation is completely voluntary. You can choose to not answer any question. You can decide at any time that you do not want to continue in the study. I hope however that you will find the study interesting and want to continue. Your responses about this important subject may help change health policy in the future. What are my rights if I participate? Everything you tell me will be kept private as it relates to you. All reports from this study will not identify you in any way. Your information will not be shared with your doctor or your medical group. No one other than me will know if you chose to participate in this study. It will not affect your benefits or medical care in any way. What risks or discomforts might occur if I participate? There are no known risks to you for participating in this study. You will need to complete the enclosed written survey which will take about 20 minutes of your time. The survey asks questions about your health and actions taken by your primary doctor. You will mail the survey back to me in the prepaid envelope. I promise to keep all of your information protected. You would also need to sign this consent form. I will link your medical service claims for 2007 to your survey responses in order to meet the main objective of this study. I am hoping to identify factors that only you can tell me that are associated with improved health outcomes for seniors. Too often studies are done on large databases that do not have a real patient voice. I want to hear your voice! You have a chance to win one of 20 (twenty) $50 Wal-Mart gift certificates if you return the survey within the first four-week timeframe.

199

This study has been approved by the Brandeis Universitys Review Board for the Protection of Human Subjects. If you have specific questions about this study you can call me, the researcher, Marian Ryan at 949-290-7697. If you have questions about your rights as a person in a study you can call Lorrie Clark from the Board. You can call her at 781-736-7596. Print name:_________________________________________________________ Your signature: __________________________________ [yes___] [no___] I give my consent to participate in this study. Date: ___________

200

Patient Survey Your responses to this survey will greatly help to make this research study a success! Please answer each question by circling the answer that is best for you. Think about the health care you received during the past year from your primary doctor. Your answers will be known only to the researcher. They will not be shared with your doctor or XX, or be published. Your specific answers will be kept private. With your permission your answers will be linked to your health care claims as part of meeting the studys goal of finding patterns of care related to health outcomes. Coordination of Care 1. 2. 3. 4. 5. 6. 7. 8. 9. When you have a medical need how often does your doctor communicate with you? Often4 Very Often5 Never1 Rarely2 Sometimes3 Does your doctor communicate with you in a timely way? Often4 Very Often5 Never1 Rarely2 Sometimes3 Does your doctor communicate with you accurately and honestly? Often4 Very Often5 Never1 Rarely2 Sometimes3 When problems arise does your doctor work with you to solve the problem? Often4 Very Often5 Never1 Rarely2 Sometimes3 Does your doctor understand you and the role you play in your own treatment? Very little2 Some3 A lot4 Completely5 Not at all1 Does your doctor respect you? Very little2 Not at all1 Some3 A lot4 Completely5

Does your doctor share your goals for treatment or your plan of care? Very little2 Some3 A lot4 Completely5 Not at all1 When you have a medical need how often do you communicate with your doctor? Often4 Very Often5 Never1 Rarely2 Sometimes3 Do you communicate with your doctor in a timely way? Often4 Very Often5 Never1 Rarely2 Sometimes3

10. Do you communicate with your doctor accurately and honestly? Often4 Very Often5 Never1 Rarely2 Sometimes3 11. When problems arise do you work with your doctor to resolve them? Often4 Very Often5 Never1 Rarely2 Sometimes3 12. Do you understand your doctor and the role he or she plays in your treatment? Very little2 Some3 A lot4 Completely5 Not at all1 13. Do you respect your doctor? Very little2 Not at all1 Some3 A lot4 Completely5 Completely5

14. Do you share your doctors goals for your treatment? Very little2 Some3 A lot4 Not at all1

201

Components of Primary Care - Coordination 15. Your doctor knows when you are due for a check-up. Often4 Very Often5 Never1 Rarely2 Sometimes3 16. Your doctor keeps track of all your health care. Often4 Never1 Rarely2 Sometimes3 Very Often5

17. Your doctor always follows up on problems you have had, either by phone or at the next visit. Often4 Very Often5 Never1 Rarely2 Sometimes3 18. Your doctor always follows up with you about visits to other doctors or health providers. Often4 Very Often5 Never1 Rarely2 Sometimes3 19. Your doctor helps you understand your lab tests, x-rays or information from visits to other doctors. Often4 Very Often5 Never1 Rarely2 Sometimes3 20. Your doctor communicates with the other doctors or health providers that you see. Often4 Very Often5 Never1 Rarely2 Sometimes3 21. How much does your doctor know about you as a person including values and beliefs that may be related to your medical care? A Little2 Some3 A Lot4 Most Everything5 Almost Nothing1 22. How much have you told your doctor about your wishes for end of life medical care? A Little2 Some3 A Lot4 Most Everything5 Almost Nothing1

Information and Support Chronic Illness Care 23. How often did your doctor ask you about your preferences about treatment to manage your conditions? Often4 Very Often5 Never1 Rarely2 Sometimes3 24. How often did your doctor review with you all of the medicine you are taking? Often4 Very Often5 Never1 Rarely2 Sometimes3 25. How often did a care manager from XX call you? (A care manager is not your doctors nurse but a special nurse who works with some patients to provide education and support.) Often4 Very Often5 Never1 Rarely2 Sometimes3 26. How often did you leave your doctors office without being able to express your feelings? Often4 Very Often5 Never1 Rarely2 Sometimes3 27. How often were you contacted after a visit to see how things were going? Often4 Very Often5 Never1 Rarely2 Sometimes3 28. How often were you encouraged to attend programs in the community that could help you? Often4 Very Often5 Never1 Rarely2 Sometimes3 29. How often were you referred to a dietitian, health educator or a counselor? Often4 Very Often5 Never1 Rarely2 Sometimes3 30. How often were you told how your visits with other doctors such as the eye doctor helped your treatment? Often4 Very Often5 Never1 Rarely2 Sometimes3

202

31. How often were you asked how your visits with other doctors were going? Often4 Very Often5 Never1 Rarely2 Sometimes3 32. How often were you totally satisfied with your doctors visit? Often4 Very Often5 Never1 Rarely2 Sometimes3

Trust 33. Sometimes your doctor cares more about what is convenient for him or her than about your medical needs. Disagree2 Neutral3 Agree4 Strongly Agree5 Strongly Disagree1 34. Your doctor is extremely thorough and careful. Disagree2 Neutral3 Strongly Disagree1 Agree4 Strongly Agree5

35. You completely trust your doctors decisions about which medical treatments are best for you. Disagree2 Neutral3 Agree4 Strongly Agree5 Strongly Disagree1 36. Your doctor is honest in telling you about all of the treatment options available for you. Disagree2 Neutral3 Agree4 Strongly Agree5 Strongly Disagree1 37. All in all, you have complete trust in your doctor. Disagree2 Neutral3 Strongly Disagree1 Needs Assessment 38. A fall is when your body goes to the ground without being pushed. In the past year did your doctor ask you about falling or any problems with balance or walking? Yes2 No1 39. Over the past year did your doctor ask you about depression? Yes2 No1 Yes2 No1 Agree4 Strongly Agree5

40. Over the past year did your doctor ask you about how much alcohol you drink? 41. Do you ever forget to take your medicine? Yes2 No1 Yes2 No1 Yes2

42. Are you careless at times about taking your medicine?

43. When you feel better, do your sometimes stop taking your medicine?

No1 No1

44. Sometimes if you feel worse when you take your medicine, do you stop taking it? Yes2 Self-Efficacy

45. Living with a chronic illness means doing tasks and activities to manage your condition. How confident are you that you can do all the things needed to manage your diabetes on a regular basis? Not Confident at all Completely Confident 1 2 3 4 5 6 7 8 9 10

203

46. How confident are you that you can do all the things to manage your other chronic condition(s)? Not Confident at all Completely Confident 1 2 3 4 5 6 7 8 9 10 47. Judge when the changes in your condition(s) mean that you should call or see your doctor? Not Confident at all Completely Confident 1 2 3 4 5 6 7 8 9 10 48. Reduce the emotional pain caused by your health condition(s) so that it does not affect your life? Not Confident at all Completely Confident 1 2 3 4 5 6 7 8 9 10 49. Do things other than just taking medicine to improve your health and well-being? Not Confident at all Completely Confident 1 2 3 4 5 6 7 8 9 10 Demographics 50. In general, would you say your health is: Poor1 Fair2 Good3

Very Good4

Excellent5

51. How often do you walk or do some other kind of exercise? Never1 Rarely2 Sometimes3 Often4 Very Often5 52. How do you describe yourself? African-American1 Asian-American2 Caucasian3 Latino/Hispanic4 Other5: ________

53. What language do you feel the most comfortable speaking? Chinese1 English2 Spanish3 Vietnamese4 54. How many years of school were you able to complete? Less than 8yrs1 8-11 yrs2 High School3 Some College4

Other5: ____________

Bachelors Degree or higher5

55. How do you find reading the health information that you receive? Very Difficult1 Difficult2 Average3 Easy4 56. Did you have help filling out this survey? 57. What is your living situation? Live alone-children close1 Live with spouse/friend3 Yes2 No1

Very Easy5

Live alone-children not nearby2 Live with children4 Yes2 No1 ______ years Live alone-no children5

58. Does your primary doctor speak your language?

59. How many years has this doctor been your primary doctor?

204

60. Do you consider yourself a religious or spiritual person? Not at all religious/spiritual1 Slightly2 Moderately3

Very religious/spiritual4

61. How often do you find that you cannot afford to pay for your medicine? Never1 Rarely2 Sometimes3 Often4 Very Often5 I appreciate very much your time in completing this survey! Please return this survey with your signed consent form. Good luck in the drawing and thank you very much!

205

Appendix F. Correlation Matrix MD RC, trust, PCC, and PACIC

Variable zMDRC zTrust zPCC zPACIC MD RC Trust PCC PACIC

N 103 103 103 103 103 103 103 103

Simple Statistics Mean StdDev 0 1.0 0 1.0 0 1.0 0 1.0 4.119 0.384 4.069 0.322 3.991 0.411 2.584 0.525

Min -3.098 -3.114 -3.221 -2.445 2.929 3.067 2.667 1.300

Max 2.106 2.269 2.319 2.950 4.929 4.800 4.944 4.133

zMDRC zTrust zPACIC zPCC

Pearson Correlation Coefficients, N=103 Prob > Rho under H0: Rho=0 zMDRC zTrust zPACIC 1.000 0.763 0.564 < .0001 < .0001 0.763 1.000 0.559 < .0001 < .0001 0.564 0.559 1.000 < .0001 < .0001 0.759 0.730 0.677 < .0001 < .0001 < .0001 Spearman Correlation Coefficients, N=103 Prob > Rho under H0: Rho=0 zMDRC zTrust zPACIC 1.000 0.746 0.510 < .0001 < .0001 0.746 1.000 0.539 < .0001 < .0001 0.510 0.539 1.000 < .0001 < .0001 0.704 0.675 0.637 < .0001 < .0001 < .0001

zPCC 0.759 < .0001 0.730 < .0001 0.677 < .0001 1.000

zMDRC zTrust zPACIC zPCC

zPCC 0.704 < .0001 0.675 < .0001 0.637 < .0001 1.000

206

Appendix G. PCP Survey Packet


Note: distributed all documents with 12-point font; shrunk for dissertation to 10-point to not lose survey formatting in meeting dissertation margin requirements) Date Dear Dr. : As a primary care physician of XX I am hoping that you will be willing to participate in my doctoral dissertation study. My research seeks to examine facilitators and barriers to physician-initiated care coordination for senior patients with multimorbidity. Although I am a doctoral fellow at Brandeis University outside Boston I worked for many years in health care. I directed disease management and care management programs before returning to school and recognized the importance of the relationship between the primary care physician and the patient as the primary facilitator of improvement in chronic illness care. I also acknowledge that many challenges arise in meeting the goals for optimal chronic illness care even in well-designed systems such as XX. Please click on the following link to access the consent and survey. Please read the physician consent form and check the box indicating your consent if you wish to complete the survey. Complete the physician survey on-line and all your responses will be contained within a secured database. I am the only person with access to your individual responses. All physicians who complete the on-line survey before March 31, 2008 will be given $10 IN CASH as a small token of my appreciation and a chance to win one of six $50 bills! The survey should take only about 20 minutes to complete. If you have any questions about this study you can call me, Marian Ryan, the researcher at 949-290-7697. If you have questions about your rights as a study participant you can call the Brandeis Research Review Board at 781-736-7596. If you do not consent please indicate and you will exit the survey and not get reminders to complete it. My research study can not be successful without your responses! It is hoped that this study might inform policy on the importance of payment to primary care physicians for care coordination. Your voice is important in this regard. Sincerely, Marian Ryan, RRT, MPH, MA, CHES AHRQ/JHF Doctoral Fellow Heller School of Social Policy and Management Brandeis University Waltham, MA 02459 myran@brandeis.edu

207

Physician Informed Consent Study Title: Care Coordination for Senior Patients with Multiple Chronic Diseases: Examining the Association between Organizational and Relational Factors, and Patient Outcomes What is the purpose of the study? This study is looking at the association between practice characteristics (the Chronic Care Model elements, physician care coordination behavior, relational factors between the physician and patient) and patient outcomes and adherence among seniors with multiple chronic conditions. Who is doing this study? Marian Ryan, a third-year doctoral student at the Heller School for Social Policy and Management, Brandeis University in Waltham, Massachusetts is the Principal Investigator (PI) for this study. You have been selected to participate because one or more of your patient panel met her study criteria that included: 1) senior patient (65 years of age or older), 2) identified diabetes and one additional chronic condition, and 3) served by XX between the years of 2003 and 2007. Although XX has consented to actively participate in this research study, you have the right to decide whether you want to participate. Marian Ryan (PI) would like to share some information about her previous experience and interests. I have enjoyed a 20-year clinical and administrative career that has included neonatal medicine, public health, patient education, quality improvement, and disease and care management programs, prior to accepting the doctoral fellowship at Brandeis in health policy. I have master degrees in public health and social policy following my entry into healthcare as a neonatal respiratory therapist working at Memorial Medical Center of Long Beach in the early 1980s. I am very interested in research that can be readily translated into practice improvements in non-academic healthcare settings. Throughout my career I have consistently demonstrated the highest level of honesty and integrity. A faculty member at the college will be supervising my work. Who is funding this study? This study is being partially funded by a grant awarded by the Jewish Healthcare Foundation. As the study costs have exceeded grant funding the PI is also funding a large portion of this study. Do I have to participate? Your participation is completely voluntary. You can choose to not answer any question. The PI hopes however that you will find the study interesting and want to continue. Your participation in this study will assist in the identification of tangible organizational and/or relational factors that may influence changes in care processes or policy to support your ability to provide the highest level of care for senior patients with multi-morbidity. What are my rights if I participate? Your survey responses will be maintained as confidential and the Principal Investigator will not disclose any information to the medical group or management. All reports from this study will not identify you in any way. No one other than the PI will know if you chose to participate in this study. You have the right to terminate your participation at any time. You can choose to not answer any question. What risks or discomforts might occur if I participate? There are no known risks to you for participating in this study. You will need to complete the enclosed written survey which will take about 20 minutes of your time. The survey asks questions about the transfer of information across healthcare settings, your style of practice, care coordination behavior and some general information about you. The survey will be returned either via secured e-mail or prepaid mail directly to the Principal Investigator and be kept confidential. The Principal Investigator will link your survey responses with the 2007 claim history from the subset of your paneled patients participating in the study. The purpose for linking physician and patient survey data directly to patient claim history for 2007 is to meet the primary research objectives. The primary goal of this study is to identify factors that facilitate and/or impede health outcomes among senior patients with multi-morbidity. Too often research is conducted on large administrative databases without the additional information that can be provided directly by physicians and patients.

208

You will be given $10 in cash and a chance to win one of six $50 bills for returning the survey before the end of April, 2008. This study has been approved by the Brandeis Universitys Review Board for the Protection of Human Subjects. If you have specific questions about this study you can call me, the researcher, Marian Ryan at 949-290-7697. If you have questions about your rights as a person in a study you can call Lorrie Clark from the Board. You can call her at 781-736-7596. Print name:_________________________________________________________ Your signature: __________________________________ [yes___] [no___] I give my consent to participate in this study. Date: ___________

209

Physician Survey Your responses to this survey will greatly make this research study a success! Please answer each question by circling the response that is best for you. In answering all questions consider only your interactions with your senior patients with multi-morbidity during the past year. The term patient will be used for the sake of brevity. Your answers will be known only to the researcher and will not be shared with XX, or be published. Your specific responses will be kept confidential. With your permission your survey responses will be linked to medical claims over the past year for your senior patients meeting the studys criteria. This is necessary in meeting the studys goal of finding patterns of care related to health outcomes among senior patients with multi-morbidity. Transfer of Information How often did you: 1. Find that the information you needed about your patient was available to you at the time of the first visit? Never Rarely Sometimes Fairly Often Very Often Always 2. 3. 4. 5. 6. Find that the information you needed about your patient was available to you at the time of the follow-up visits? Never Rarely Sometimes Fairly Often Very Often Always Find that information was timely following a hospitalization of one of your patients? Never Rarely Sometimes Fairly Often Very Often Always Find that information was accurate following a hospitalization of one of your patients? Never Rarely Sometimes Fairly Often Very Often Always Obtain feedback from behavioral health providers on patients referred? Never Rarely Sometimes Fairly Often Very Often Obtain feedback from specialists on patients referred? Never Rarely Sometimes Fairly Often Very Often Always Always

Communication with Patients How often did you: 7. Ask patients about their fears and/or concerns? Never Rarely Sometimes Fairly Often Very Often 8. 9.

Always

Find language discordance to be a barrier to meeting your goals during a patient visit? Never Rarely Sometimes Fairly Often Very Often Always Find health literacy to be a barrier to meeting your goals during a patient visit? Never Rarely Sometimes Fairly Often Very Often Always

10. Assess recall of your care instructions by asking patient to tell you what they will do? Never Rarely Sometimes Fairly Often Very Often Always 11. Problem-solve with your patients the barriers they face in managing their chronic condition(s)? Never Rarely Sometimes Fairly Often Very Often Always 12. Ask about their preferences for treatment options? Never Rarely Sometimes Fairly Often 13. Ask about their preferences for end of life care? Never Rarely Sometimes Fairly Often Very Often Always

Very Often

Always

210

14. For how many patients have you discussed end of life care wishes? None Few Some Many Style of Practice How often did you: 15. Screen or assess your patients for depression? Never Rarely Sometimes Fairly Often

Most

All

Very Often

Always Always Always Always Always

16. Assess all health habits such as nutrition, exercise, smoking and drinking? Never Rarely Sometimes Fairly Often Very Often 17. Assess patients living situation related to their health status? Never Rarely Sometimes Fairly Often 18. Refer patients to Disease Management or the dietician within XX? Never Rarely Sometimes Fairly Often 19. Refer patients to XX Care Management program? Never Rarely Sometimes Fairly Often Very Often Very Often Very Often

20. Refer patients to available community resources to assist with support and/or education? Never Rarely Sometimes Fairly Often Very Often Always 21. Look at patient log books or records kept of previous glucose or B/P readings or weights? Never Rarely Sometimes Fairly Often Very Often Always 22. Initiate follow-up with the specialists seeing your patients? Never Rarely Sometimes Fairly Often Very Often Always

23. Make sure that the information from specialists and other providers was put into the patients medical chart? Never Rarely Sometimes Fairly Often Very Often Always 24. Use the physician portal with patient intervention reports to help you manage your patients? Never Rarely Sometimes Fairly Often Very Often Always 25. Ensure that advance directives are recorded for your patients? Never Rarely Sometimes Fairly Often Very Often Always

Medication Management How often did you: 26. Review all the medications that your patients are taking with your patients? Never Rarely Sometimes Fairly Often Very Often 27. Ask about over the counter medications and alternative medicine? Never Rarely Sometimes Fairly Often Very Often

Always Always

28. Simplify the medication regime of your patients to increase patient adherence to your treatment plan? Never Rarely Sometimes Fairly Often Very Often Always 29. Assess how well medications are working and ask about side effects? Never Rarely Sometimes Fairly Often Very Often

Always

211

Organizational Support/Tools 30. The hospital in which my patients are hospitalized impacts my ability to adequately coordinate my patients follow-up care. Strongly Disagree Disagree Neutral Agree Strongly Agree 31. My senior patients with multi-morbidity and myself are a team working together to ensure positive health outcomes. Strongly Disagree Disagree Neutral Agree Strongly Agree 32. Reminder systems are important to ensure that timely follow-up care is provided. Strongly Disagree Disagree Neutral Agree Strongly Agree 33. HCPs pharmacy reports are not helpful to me for the medication management of my patients. Strongly Disagree Disagree Neutral Agree Strongly Agree 34. XXs care managers have assisted me with care coordination and improved my ability to improve outcomes for my patients. Strongly Disagree Disagree Neutral Agree Strongly Agree 35. DM health educators and the dieticians within XX give me feedback that helps me and guides my patients care. Strongly Disagree Disagree Neutral Agree Strongly Agree 36. Inadequate training and support for chronic illness management prevents me from achieving the best possible patient care outcomes. Strongly Disagree Disagree Neutral Agree Strongly Agree 37. Adequate time is scheduled for patients with multi-morbidity to provide care coordination, education and follow-up. Strongly Disagree Disagree Neutral Agree Strongly Agree Health Plan Relationships 38. The patient profiles sent or faxed to me from our various health plans were helpful to me in managing my patients. Strongly Disagree Disagree Neutral Agree Strongly Agree 39. Case Management or care coordination done by our contracted health plans was more of a hindrance than a help in managing my patients. Strongly Disagree Disagree Neutral Agree Strongly Agree 40. Health plans did not keep the primary care physician informed about patients enrolled in disease management or other special programs. Strongly Disagree Disagree Neutral Agree Strongly Agree Demographics 41. What is your specialty? Internal Medicine Family Practice Geriatrics Other > 15 yrs

42. How many years have you practiced medicine? < 3 yrs 3 to 5 yrs >5 to 10 yrs

>10 to 15 yrs

212

43. How many years have your worked for XX or the group assumed by HCP? < 3 yrs 3 to 5 yrs >5 to 10 yrs >10 to 15 yrs > 15 yrs 44. Where did you obtain your medical training? Northeast Southeast South Midwest 45. What languages do you speak other than English? Chinese Vietnamese Korean Spanish 46. What is your age? < 35 yrs 47. What is your gender? 35 to 40 yrs Male >40 to 50 yrs Female > 50% Northwest Southwest

Other: __________________ >50 to 60 yrs > 60 yrs

48. What percentage of your patients are ethnic minorities? <10% 10 to 25% >25 to 35% >35 to 50%

49. What percentage of your patients is enrolled currently in XX Care Management program? <5% 5 to 10% >10 to 20% >20 to 33% > 33% 50. For what percentage of your patients are you primarily working with the care-giver rather than the patient? <5% 5 to 10% >10 to 20% >20 to 33% > 33% 51. What is working today in medicine or specifically within HCP to assist you with coordinating the care of your senior patients with multiple chronic illnesses and being the captain of the ship?

52. What are your greatest challenges to coordinating the care of your senior patients with multiple chronic illnesses?

Thank you so much for your time and participation. Your voice is critical to this study. Good luck with the drawing!

Refers to your senior patients with multi-morbidity over the past year (2007)

213

Appendix H. ACIC with introduction


Wagners Assessment Chronic Illness Care Framework (ACIC version 3.5)
Copyright 2000 MacColl Institute for Healthcare Innovation, Group Health Cooperative This ACIC has been completed via an in-person interview process to facilitate discussion and understanding of all characteristics comprising each component within the Chronic Care Model. XX Clinic HCP Region Clinic Administrator Estimated Clinic Demographics Urgent Care On-site Interview Date

Other Participants

Titles

This ACIC was completed with consideration of the three high frequency chronic illnesses within the elder population diabetes, coronary disease/CHF, and COPD. Each factor was evaluated strictly from the perspective of this facility not the overall organization and specifically for the 2007 year.

214

Wagners Assessment Chronic Illness Care Framework (ACIC)


Copyright 2000 MacColl Institute for Healthcare Innovation, Group Health Cooperative Part 1: Organization of the Healthcare Delivery System. Chronic illness management programs can be more effective if overall system (organization) in which care is provided is oriented and led in a manner that allows for a focus on chronic illness care. Components Overall organizational leadership in Chronic Illness Care Level D does not exist or there is little interest. Level C is reflected in vision statements and business plans, but no dedicated resources. 3 4 5 exist but are not actively reviewed at least annually. Level B is reflected by senior leadership and specific dedicated resources (dollars and personnel). 6 7 8 are measured and reviewed at least annually. Level A is part of the systems long planning strategy adequate resources and specific people are held accountable for results. 9 10 11 are measured, reviewed and are incorporated into plans for improvement. 9 10 11 includes a proven improvement strategy and uses it proactively in meeting organizational goals. 9 10 11 are used to motivate and empower providers to support patient care goals. 9 10 11

Score Organizational Goals for Chronic Care

0 1 2 do not exist or are limited to one condition.

Score Improvement strategy for Chronic Illness Care

0 1 2 is ad hoc and not organized or supported consistently. 0 1 2 are not used to influence clinical performance goals.

3 4 5 utilizes ad hoc approaches for targeted problems as they emerge. 3 4 5 are used to influence utilization and costs of chronic illness care. 3 4 5

6 7 8 utilizes a proven improvement strategy for targeted problems. 6 7 8 are used to support patient care goals.

Score Incentives and regulations for Chronic Illness Care

Score Senior Leaders

discourage enrollment of the chronically ill.

do not make improvements to chronic illness care a priority. 3 4 5 neither encourages nor discourages patient self-management or system changes. 3 4 5

encourage improvement efforts in chronic care.

visibly participate in improvement efforts in chronic care. 9 10 11 are specifically designed to promote better chronic illness care. 9 10 11

Score Benefits

0 1 2 discourage patient self-management or system changes.

6 7 8 encourage patient self-management or system changes.

Score

Total Health Care Organization Score: ______

Average Score (Health Care Organization Score/6): ______

215

Part 2: Community Linkages. Linkages between the provider practice and community resources play important roles in the management of chronic illness. Components Linking Patients to Outside Resources Level D is not done systematically. Level C is limited to a list of identified community resources in an accessible format. Level B is done by a designated person, responsible for ensuring providers and patients make maximum use of community resources. 6 7 8 are formed to develop supportive programs and policies. Level A is done through active coordination between the facility, community service agencies and the patients. 9 10 11 are actively sought to develop formal supportive programs and policies across your system. 9 10 11 currently coordinate chronic illness guidelines, measures and resources at the practice level for most chronic illnesses. 10 11

Score Partnerships with Community Organizations

0 1 do not exist.

3 4 5 are being considered but have not yet been implemented.

Score Health Plans

0 1 2 do not coordinate chronic illness guidelines, measures or care resources at the practice level.

Score

4 5 would consider some degree of coordination of guidelines, measures or care resources at the practice level but have not yet implemented changes. 3 4 5

7 8 currently coordinate guidelines, measures or care resources in one or two chronic illness areas.

Total Community Linkage Score: ______

Average Score (Community Linkage Score/3): ______

216

Part 3: Practice Level. Four components at the provider practice level (individual clinic) Part 3a: Self-Management Support. Effective self-management support can help patients and families cope with challenges of living with and treating chronic illness and reduce complications and symptoms. Components Assessment and Documentation of Self-Management Needs and Activities Level D are not done. Level C are expected to be done. Level B are completed in a standardized manner. Level A are regularly assessed and recorded in a standardized form AND linked to a treatment plan available to practice and patients. 9 10 11 is provided by clinical educators, affiliated with each practice and trained in patient empowerment, problem-solving and see most patients with chronic illness. 9 10 11 is an integral part of care includes systematic assessment and routine involvement in peer support, groups or mentoring programs. 9 10 11 are readily available and an integral part of routine care.

Score Self-Management Support

0 1 2 is limited to the distribution of information (pamphlets, booklets, etc).

3 4 5 is available by referral to selfmanagement classes or educators.

6 7 8 is provided by trained clinical educators who are designated to do self-mgt support, affiliated with each practice, and see patients on referral. 6 7 8 is encouraged, and peer support and mentoring programs are available.

Score Addressing Concerns of Patients and Families

0 1 2 is not consistently done.

3 4 5 is provided for specific patients and families via referral.

Score Effective Behavior Change Interventions and Peer Support Score

0 1 2 are not available.

3 4 5 are limited to the distribution of pamphlets, booklets or written information.

6 7 8 are available only by referral to specialized centers staffed by trained personnel.

10

11

Total Self-Management Support Score: ______ Average Score (Self-Management Support Score/4): ______

217

Part 3b: Decision Support. Assure that providers have access to evidence-based information, decision-support to care for patients including: evidence-based practice guidelines or protocols, specialty consultation, provider education and activating patients to make provider teams aware of effective strategies. Components Evidence-Based Guidelines Level D are not available. Level C are available but are not integrated into care delivery. Level B are available and supported by provider education. Level A are available, supported by provider education and integrated via reminders and other proven provider behavior change methods. 9 10 11 includes specialist leadership and specialist involvement in improving the care of primary care patients. 9 10 11 includes training all practice teams in chronic illness care methods population-based and self-mgt support. 9 10 11 includes specific materials developed for patients that describe their role in achieving guideline adherence. 9 10 11

Score Involvement of Specialists in Improving Primary Care

0 1 2 is primarily through traditional referral.

Score Provider Education for Chronic Illness Care

0 1 is provided sporadically.

3 4 5 is achieved through specialist leadership to enhance the capacity of the facility to implement guidelines 3 4 5 is provided systematically through traditional methods.

6 7 8 includes specialist leadership and designated specialists provide primary care team training. 6 7 8 is provided using optimal methods (e.g. academic detailing site visit education). 6 7 8 is done through specific patient education materials for each guideline. 6 7 8

Score Informing Patients about Guidelines

0 1 are not done.

3 4 5 happens on request of through system publications.

Score

Total Decision-Support Score: ______

Average Score (Decision-Support Score/4): ______

218

Part 3c: Delivery System Design. Effective chronic illness management may require changes to the organization of the practice to impact the provision of care. Components Practice Team Functioning (Care Team for HCP) Level D are not addressed. Level C is assured by the availability of individuals with appropriate training in key elements of chronic illness care. Level B is assured by regular team meetings to address guidelines, role and accountability, and problems in chronic care. Level A is assured by teams who meet regularly and have clearly defined roles including patient self-mgt education, proactive f/u, resource coordination, etc. in chronic care. 9 10 11 is guaranteed by the appointment of a team leader who assures that roles and accountabilities for chronic care are clearly defined. 9 10 11 includes organization of care that facilitates the patient seeing multiple providers in a single visit. 9 10 11 is customized to patient needs (phone, in-person, email) and assures guideline f/u.

Score Practice Team Leadership

0 1 2 is not recognized within the facility.

3 4 5 is assumed by the organization to reside in specific organizational roles.

6 7 8 is assured by the appointment of a team leader but the role in chronic illness is not defined. 6 7 8 is flexible and can accommodate innovations such as customized visit length or group visits. 6 7 8 is assured by the care team by monitoring patient utilization. 6 7 8 are an option for interested patients.

Score Appointment System

0 1 2 can be used to schedule acute care visits, f/u and preventive visits. 0 1 2 is scheduled by patients or providers in an ad hoc fashion.

3 4 5 assures scheduled follow-up with chronically ill patients. 3 4 5 is scheduled by the practice in accordance with guidelines (e.g. at least two visits/year for diabetes). 3 4 5 are occasionally used for complicated patients.

Score Follow-up

9 10 11 are used for all patients and include regular assessment, preventive care, and self-mgt support. 1 2 3 4 5 6 7 8 9 10 11 Score 0 is not a priority. depends upon between PCPs, Continuity of is a high priority and specialists and Care and all chronic written others is a priority disease interventions communication between PCPs, but not includes active specialists, case implemented coordination managers or disease systematically. between primary management care, specialists and educators. other team members. 1 2 3 4 5 6 7 8 9 Score 0 10 11 Total Delivery System Design Score: ______ Average Score (Delivery System Design Score/6): _____ Score Planned Chronic Illness Care Visits

0 1 are not used.

219

Part 3d: Clinical Information Systems. Assure that providers have timely, useful information about individual patients and populations of patients with chronic conditions. Components Registry (a list of patients with specific conditions) Level D is not available. Level C includes name, diagnosis, contact information and date of last contact paper or database. 3 4 5 includes notification of chronic illness but does not give information about needed services at next visit. 3 4 is provided at infrequent intervals and is delivered impersonally. 5 Level B allows queries to sort sub-populations by clinical priorities (e.g. all persons with diabetes plus coronary artery disease). 6 7 8 includes indications of needed service for populations of pts. via periodic reporting (e.g. diabetes pts needing flu shot) 6 7 8 occurs at frequent enough intervals to monitor performance and is specific to the care teams population. 6 7 8 can be obtained upon request but is not routinely available. 6 7 8 are established collaboratively and include selfmanagement as well as clinical goals. 6 7 8 Level A is tied to guidelines which provide prompts and reminders about needed services. 9 10 11 includes specific information for the team about guideline adherence at the time of the individual patient visit. 9 10 11 is timely, specific to the care team, routine and delivered by a respected opinion leader to improve team performance. 9 10 11 is provided routinely to providers to help them deliver planned care. 9 10 11 are established collaboratively and include self-mgt and clinical mgt. Followup occurs and guides care at every visit. 9 10 11

Score Reminders to Providers

0 1 2 are not available.

Score Feedback

0 1 2 are not available or is non-specific to the care team.

Score Information about Relevant Subgroups of Patients Needing Services Score Patient Treatment Plans

0 1 2 is not available.

3 4 5 can be obtained only with special efforts or additional programming. 3 4 are achieved through a standardized approach. 5

0 1 2 are not expected.

Score

Total Clinical Information System Score: _____ Average Score (Clinical Information System Score/5):_____

220

Integration of Chronic Care Model Components. Combines all the elements of the Chronic Care Model Components Informing Patients about Guidelines (repeat of 3b-d) Score Information Systems/Registries Level D are not done. Level C happens on request of through system publications. 3 4 5 include results of patient assessments (e.g. functional status, readiness to change), but no goals. 3 4 5 provide sporadic feedback at joint meetings between the clinic and community program. 3 4 5 uses data from information systems to plan care. Level B is done through specific patient education materials for each guideline. 6 7 8 include results of patient assessments and self-mgt goals developed using input from the care team, PCP and patient. 6 7 8 provide regular feedback to the clinic using formal mechanisms (e.g. Internet progress report) about patients progress. 6 7 8 uses data from information systems to proactively plan population-based care including the development of self-mgt programs and partnerships with community resources. 6 7 8 is ensured by assigning responsibilities to specific staff (e.g. case manager). 6 7 8 are provided to all pts. to help them develop effective self-mgt skills and know when to see PCP. 6 7 8 Level A includes specific materials developed for patients that describe their role in achieving guideline adherence. 9 10 11 include results of pt. assessments, self-mgt goals developed collaboratively and prompt reminders to patient and/or PCP about f/u to re-evaluate goals. 9 10 11 provide regular feedback to the clinic that requires input from patients that is then used to modify programs to better meet the needs to patients. 9 10 11 all of level B plus include a built-in evaluation plan to determine success over time.

0 1 2 do not include patient self-mgt goals.

Score Community Programs

0 1 2 do not provide feedback to the clinic about the patients progress in their programs.

Score Organizational Planning for Chronic Illness Care

0 1 2 do not include a population-based approach.

Score Routine f/u for appointments, patient assessments and goal planning Score Guidelines for Chronic Illness Care

0 1 2 are not ensured.

3 4 5 is sporadically done, usually for appointments only.

0 1 2 are not shared with patients.

3 4 5 are given to patients who express interest in self-mgt of their condition. 3 4 5

Score

9 10 11 is ensured by Level B plus specific person uses registry and other prompts to coordinate with patients and entire care team. 9 10 11 are reviewed by the care team with the patient to develop selfmgt plan to meet guidelines and patients goals and readiness to change. 9 10 11

Total Integration Score: ______

Average Score (Integration Score/6): ______

221

Appendix I. Correlation Matrix PredCMCD and RC and Trust


Correlation between Key Predictor of Communication/Coordination used as proxy variable in the Longitudinal Analyses, and RC and Trust in the Cross-sectional Analyses
Variable predCMCD zRC zTrust avgCMCD RC Trust N 84 84 84 84 84 84 Mean 0.119 -0.018 -0.021 4.338 4.113 4.063 StdDev 1.222 0.961 1.017 0.136 0.369 0.327 Median 0.162 0.036 -0.129 4.344 4.133 4.028 Min -3.148 -3.098 -3.114 3.966 2.929 3.067 Max 2.376 2.106 2.269 4.586 4.929 4.800

PredCMCD zRC zTrust

Pearson Correlation Coefficients, N=84 Prob > Rho under H0: Rho=0 PredCMCD zRC 0.337 1.000 0.002 0.337 1.000 0.002 0.314 0.717 0.004 < .0001

zTrust 0.314 0.004 0.717 < .0001 1.000

Spearman Correlation Coefficients, N=84 Prob > Rho under H0: Rho=0 PredCMCD zRC 0.311 1.000 PredCMCD 0.004 0.311 1.000 zRC 0.004 0.255 0.654 zTrust 0.019 < .0001 Pearson Correlation Coefficients, N=84 Prob > Rho under H0: Rho=0 AvgCMCD zRC 0.332 1.000 0.002 0.332 1.000 0.002 0.316 0.717 0.003 < .0001

zTrust 0.255 0.019 0.654 < .0001 1.000

AvgCMCD zRC zTrust

zTrust 0.316 0.003 0.717 < .0001 1.000

222

Spearman Correlation Coefficients, N=84 Prob > Rho under H0: Rho=0 AvgCMCD zRC 0.303 1.000 AvgCMCD 0.005 0.303 1.000 zRC 0.005 0.258 0.654 zTrust 0.018 < .0001

zTrust 0.258 0.018 0.654 < .0001 1.000

223

You might also like